var bibbase_data = {"data":"\"Loading..\"\n\n
\n\n \n\n \n\n \n \n\n \n\n \n \n\n \n\n \n
\n generated by\n \n \"bibbase.org\"\n\n \n
\n \n\n
\n\n \n\n\n
\n\n Excellent! Next you can\n create a new website with this list, or\n embed it in an existing web page by copying & pasting\n any of the following snippets.\n\n
\n JavaScript\n (easiest)\n
\n \n <script src=\"https://bibbase.org/show?bib=http%3A%2F%2Fweb.stanford.edu%2Fgroup%2Ffullergroup%2Fcgi-bin%2Ffullerlabsite%2FFuller_bibtexfile.bib&jsonp=1&jsonp=1\"></script>\n \n
\n\n PHP\n
\n \n <?php\n $contents = file_get_contents(\"https://bibbase.org/show?bib=http%3A%2F%2Fweb.stanford.edu%2Fgroup%2Ffullergroup%2Fcgi-bin%2Ffullerlabsite%2FFuller_bibtexfile.bib&jsonp=1\");\n print_r($contents);\n ?>\n \n
\n\n iFrame\n (not recommended)\n
\n \n <iframe src=\"https://bibbase.org/show?bib=http%3A%2F%2Fweb.stanford.edu%2Fgroup%2Ffullergroup%2Fcgi-bin%2Ffullerlabsite%2FFuller_bibtexfile.bib&jsonp=1\"></iframe>\n \n
\n\n

\n For more details see the documention.\n

\n
\n
\n\n
\n\n This is a preview! To use this list on your own web site\n or create a new web site from it,\n create a free account. The file will be added\n and you will be able to edit it in the File Manager.\n We will show you instructions once you've created your account.\n
\n\n
\n\n

To the site owner:

\n\n

Action required! Mendeley is changing its\n API. In order to keep using Mendeley with BibBase past April\n 14th, you need to:\n

    \n
  1. renew the authorization for BibBase on Mendeley, and
  2. \n
  3. update the BibBase URL\n in your page the same way you did when you initially set up\n this page.\n
  4. \n
\n

\n\n

\n \n \n Fix it now\n

\n
\n\n
\n\n\n
\n \n \n
\n
\n  \n 2022\n \n \n (10)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Instability and symmetry breaking of surfactant films over an air bubble.\n \n \n \n\n\n \n Shi, X.; Fuller, G. G.; and Shaqfeh, E. S.\n\n\n \n\n\n\n Journal of Fluid Mechanics,A26. 2022.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{shi2022surfactant,\n  title  ={Instability and symmetry breaking of surfactant films over an air bubble},\n  author={Shi, Xingyi and Fuller, Gerald G. and Shaqfeh, Eric S.G.},\n  journal={Journal of Fluid Mechanics},\n  pages={A26},\n  year={2022}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Kitchen flows: Making science more accessible, affordable, and curiosity driven.\n \n \n \n\n\n \n Fuller, G. G.; Lisicki, M.; Mathijssen, A. J. T. M.; Mossige, E. J. L.; Pasquino, R.; Prakash, V. N.; and Ramos, L.\n\n\n \n\n\n\n Physics of Fluids,110401. 2022.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller2022kitchen,\n  title  ={Kitchen flows: Making science more accessible, affordable, and curiosity driven},\n  author={Fuller,Gerald G.  and Lisicki,Maciej  and Mathijssen,Arnold J. T. M.  and Mossige,Endre J. L.  and Pasquino,Rossana  and Prakash,Vivek N.  and Ramos,Laurence},\n  journal={Physics of Fluids},\n  pages={110401},\n  year={2022}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Effect of Recombinant Human Lubricin on Model Tear Film Stability.\n \n \n \n\n\n \n Cui, K. W.; Xia, V. X.; Cirera-Salinas, D.; Myung, D.; and Fuller, G. G.\n\n\n \n\n\n\n Translational Vision Science & Technology,9-9. 2022.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{cui2022lubricin,\n  title  ={Effect of Recombinant Human Lubricin on Model Tear Film Stability},\n  author={Cui, Kiara W. and Xia, Vincent X. and Cirera-Salinas, Daniel and Myung, David and Fuller, Gerald G.},\n  journal={Translational Vision Science & Technology},\n  pages={9-9},\n  year={2022}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Tear Film Stability as a Function of Tunable Mucin Concentration Attached to Supported Lipid Bilayers.\n \n \n \n\n\n \n Cui, K. W.; Myung, D. J.; and Fuller, G. G.\n\n\n \n\n\n\n The Journal of Physical Chemistry B,6338-6344. 2022.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{cui2022tear,\n  title  ={Tear Film Stability as a Function of Tunable Mucin Concentration Attached to Supported Lipid Bilayers},\n  author={Cui, Kiara W. and Myung, David J. and Fuller, Gerald G.},\n  journal={The Journal of Physical Chemistry B},\n  pages={6338-6344},\n  year={2022}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Influence of salt on the formation and separation of droplet interface bilayers.\n \n \n \n\n\n \n Huang, Y. (.; Chandran Suja, V.; Amirthalingam, L.; and Fuller, G. G.\n\n\n \n\n\n\n Physics of Fluids,067107. 2022.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{huang2022salt,\n  title  ={Influence of salt on the formation and separation of droplet interface bilayers},\n  author={Huang,Yaoqi (黄尧奇) and Chandran Suja,Vineeth  and Amirthalingam,Layaa  and Fuller,Gerald G. },\n  journal={Physics of Fluids},\n  pages={067107},\n  year={2022}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Dewetting characteristics of contact lenses coated with wetting agents.\n \n \n \n\n\n \n Chandran Suja, V.; Verma, A.; Mossige, E.; Cui, K.; Xia, V.; Zhang, Y.; Sinha, D.; Joslin, S.; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science,24-32. 2022.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chandran2022dewetting,\n  title  ={Dewetting characteristics of contact lenses coated with wetting agents},\n  author={V. {Chandran Suja} and A. Verma and E.J.L. Mossige and K.W. Cui and V. Xia and Y. Zhang and D. Sinha and S. Joslin and G.G. Fuller},\n  journal={Journal of Colloid and Interface Science},\n  pages={24-32},\n  year={2022}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n A Mucin-Deficient Ocular Surface Mimetic Platform for Interrogating Drug Effects on Biolubrication, Antiadhesion Properties, and Barrier Functionality.\n \n \n \n\n\n \n Madl, A. C.; Liu, C.; Cirera-Salinas, D.; Fuller, G. G.; and Myung, D.\n\n\n \n\n\n\n ACS Applied Materials & Interfaces,18016-18030. 2022.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{madl2022mucin,\n  title  ={A Mucin-Deficient Ocular Surface Mimetic Platform for Interrogating Drug Effects on Biolubrication, Antiadhesion Properties, and Barrier Functionality},\n  author={Madl, Amy C. and Liu, Chunzi and Cirera-Salinas, Daniel and Fuller, Gerald G. and Myung, David},\n  journal={ACS Applied Materials \\& Interfaces},\n  pages={18016-18030},\n  year={2022}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n A shape stability model for 3D printable biopolymer-bound soil composite.\n \n \n \n\n\n \n Biggerstaff, A.; Lepech, M.; Fuller, G. G.; and Loftus, D.\n\n\n \n\n\n\n Construction and Building Materials,126337. 2022.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{biggerstaff2022soil,\n  title  ={A shape stability model for 3D printable biopolymer-bound soil composite},\n  author={Biggerstaff, Adrian and Lepech, Michael and Fuller, Gerald G. and Loftus, David},\n  journal={Construction and Building Materials},\n  pages={126337},\n  year={2022}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Foaming and antifoaming in non-aqueous liquids.\n \n \n \n\n\n \n Calhoun, S. G. K.; Suja, V. C.; and Fuller, G. G.\n\n\n \n\n\n\n Current Opinion in Colloid & Interface Science,101558. 2022.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{calhoun2022foaming,\n  title  ={Foaming and antifoaming in non-aqueous liquids},\n  author={Calhoun, Suzanne G. K. and Suja, Vineeth Chandran and Fuller, Gerald G.},\n  journal={Current Opinion in Colloid & Interface Science},\n  pages={101558},\n  year={2022}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Systematic characterization of effect of flow rates and buffer compositions on double emulsion droplet volumes and stability.\n \n \n \n\n\n \n Calhoun, S. G. K.; Brower, K. K.; Suja, V. C.; Kim, G.; Wang, N.; McCully, A. L.; Kusumaatmaja, H.; Fuller, G. G.; and Fordyce, P. M.\n\n\n \n\n\n\n Lab Chip,2315-2330. 2022.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{calhoun2022droplet,\n  title  ={Systematic characterization of effect of flow rates and buffer compositions on double emulsion droplet volumes and stability},\n  author={Calhoun, Suzanne G. K. and Brower, Kara K. and Suja, Vineeth Chandran and Kim, Gaeun and Wang, Ningning and McCully, Alexandra L. and Kusumaatmaja, Halim and Fuller, Gerald G. and Fordyce, Polly M.},\n  journal={Lab Chip},\n  pages={2315-2330},\n  year={2022}\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2021\n \n \n (15)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Dewetting Characteristics of Contact Lenses Coated with Wetting Agents.\n \n \n \n\n\n \n Chandran Suja, V.; Verma, A.; Mossige, E. J.; Cui, K.; Xia, V.; Zhang, Y.; Sinha, D.; Joslin, S.; and Fuller, G. G\n\n\n \n\n\n\n arXiv e-prints,arXiv–2112. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chandran2021dewetting,\n  title={Dewetting Characteristics of Contact Lenses Coated with Wetting Agents},\n  author={Chandran Suja, Vineeth and Verma, Archana and Mossige, Endre Joachim and Cui, Kiara and Xia, Vincent and Zhang, Yong and Sinha, Dola and Joslin, Scott and Fuller, Gerald G},\n  journal={arXiv e-prints},\n  pages={arXiv--2112},\n  year={2021}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Interfacial Assembly of Graphene Oxide: From Super Elastic Interfaces to Liquid-in-Liquid Printing.\n \n \n \n\n\n \n Kamkar, M.; Erfanian, E.; Bazazi, P.; Ghaffarkhah, A.; Sharif, F.; Xie, G.; Kannan, A.; Arjmand, M.; Hejazi, S H.; Russell, T. P; and others\n\n\n \n\n\n\n Advanced Materials Interfaces,2101659. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kamkar2021interfacial,\n  title={Interfacial Assembly of Graphene Oxide: From Super Elastic Interfaces to Liquid-in-Liquid Printing},\n  author={Kamkar, Milad and Erfanian, Elnaz and Bazazi, Parisa and Ghaffarkhah, Ahmadreza and Sharif, Farbod and Xie, Ganhua and Kannan, Aadithya and Arjmand, Mohammad and Hejazi, S Hossein and Russell, Thomas P and others},\n  journal={Advanced Materials Interfaces},\n  pages={2101659},\n  year={2021},\n  publisher={Wiley Online Library}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Foaming and antifoaming in non-aqueous liquids.\n \n \n \n\n\n \n Calhoun, S. G.; Suja, V C.; and Fuller, G. G\n\n\n \n\n\n\n Current Opinion in Colloid & Interface Science,101558. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{calhoun2021foaming,\n  title={Foaming and antifoaming in non-aqueous liquids},\n  author={Calhoun, Suzanne GK and Suja, V Chandran and Fuller, Gerald G},\n  journal={Current Opinion in Colloid \\& Interface Science},\n  pages={101558},\n  year={2021},\n  publisher={Elsevier}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Miscible antifoams: Leveraging evaporative solutocapillary flows for a novel antifoam mechanism.\n \n \n \n\n\n \n Calhoun, S.; Chadran Suja, V.; Nguyen, L.; and Fuller, G.\n\n\n \n\n\n\n Bulletin of the American Physical Society, 66. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{calhoun2021miscible,\n  title={Miscible antifoams: Leveraging evaporative solutocapillary flows for a novel antifoam mechanism},\n  author={Calhoun, Suzanne and Chadran Suja, Vineeth and Nguyen, Lien and Fuller, Gerald},\n  journal={Bulletin of the American Physical Society},\n  volume={66},\n  year={2021},\n  publisher={APS}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Determining the yield stress of a Biopolymer-bound Soil Composite for extrusion-based 3D printing applications.\n \n \n \n\n\n \n Biggerstaff, A.; Fuller, G.; Lepech, M.; and Loftus, D.\n\n\n \n\n\n\n Construction and Building Materials, 305: 124730. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{biggerstaff2021determining,\n  title={Determining the yield stress of a Biopolymer-bound Soil Composite for extrusion-based 3D printing applications},\n  author={Biggerstaff, Adrian and Fuller, Gerald and Lepech, Michael and Loftus, David},\n  journal={Construction and Building Materials},\n  volume={305},\n  pages={124730},\n  year={2021},\n  publisher={Elsevier}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Platform and methods for dynamic thin film measurements using hyperspectral imaging.\n \n \n \n\n\n \n Fuller, G. G; and Suja, V. C.\n\n\n \n\n\n\n August 26 2021.\n US Patent App. 17/175,799\n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@misc{fuller2021platform,\n  title={Platform and methods for dynamic thin film measurements using hyperspectral imaging},\n  author={Fuller, Gerald G and Suja, Vineeth Chandran},\n  year={2021},\n  month=aug # "~26",\n  publisher={Google Patents},\n  note={US Patent App. 17/175,799}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Mucin-Like Glycoproteins Modulate Interfacial Properties of a Mimetic Ocular Epithelial Surface.\n \n \n \n\n\n \n Liu, C.; Madl, A. C; Cirera-Salinas, D.; Kress, W.; Straube, F.; Myung, D.; and Fuller, G. G\n\n\n \n\n\n\n Advanced Science, 8(16): 2100841. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{liu2021mucin,\n  title={Mucin-Like Glycoproteins Modulate Interfacial Properties of a Mimetic Ocular Epithelial Surface},\n  author={Liu, Chunzi and Madl, Amy C and Cirera-Salinas, Daniel and Kress, Wolfgang and Straube, Frank and Myung, David and Fuller, Gerald G},\n  journal={Advanced Science},\n  volume={8},\n  number={16},\n  pages={2100841},\n  year={2021},\n  publisher={Wiley Online Library}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Flowering in bursting bubbles with viscoelastic interfaces.\n \n \n \n\n\n \n Tammaro, D.; Suja, V. C.; Kannan, A.; Gala, L. D.; Di Maio, E.; Fuller, G. G; and Maffettone, P. L.\n\n\n \n\n\n\n Proceedings of the National Academy of Sciences, 118(30). 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{tammaro2021flowering,\n  title={Flowering in bursting bubbles with viscoelastic interfaces},\n  author={Tammaro, Daniele and Suja, Vinny Chandran and Kannan, Aadithya and Gala, Luigi Davide and Di Maio, Ernesto and Fuller, Gerald G and Maffettone, Pier Luca},\n  journal={Proceedings of the National Academy of Sciences},\n  volume={118},\n  number={30},\n  year={2021},\n  publisher={National Acad Sciences}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Instability and symmetry breaking in binary evaporating thin films over a solid spherical dome.\n \n \n \n\n\n \n Shi, X.; Rodrı́guez-Hakim, Mariana; Shaqfeh, E. S.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Fluid Mechanics, 915. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{shi2021instability,\n  title={Instability and symmetry breaking in binary evaporating thin films over a solid spherical dome},\n  author={Shi, Xingyi and Rodr{\\'\\i}guez-Hakim, Mariana and Shaqfeh, Eric SG and Fuller, Gerald G},\n  journal={Journal of Fluid Mechanics},\n  volume={915},\n  year={2021},\n  publisher={Cambridge University Press}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n In-use interfacial stability of monoclonal antibody formulations diluted in saline iv bags.\n \n \n \n\n\n \n Kannan, A.; Shieh, I. C; Hristov, P.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Pharmaceutical Sciences, 110(4): 1687–1692. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kannan2021use,\n  title={In-use interfacial stability of monoclonal antibody formulations diluted in saline iv bags},\n  author={Kannan, Aadithya and Shieh, Ian C and Hristov, Petar and Fuller, Gerald G},\n  journal={Journal of Pharmaceutical Sciences},\n  volume={110},\n  number={4},\n  pages={1687--1692},\n  year={2021},\n  publisher={Elsevier}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Adsorption and Aggregation of Monoclonal Antibodies at Silicone Oil–Water Interfaces.\n \n \n \n\n\n \n Kannan, A.; Shieh, I. C; Negulescu, P. G; Chandran Suja, V.; and Fuller, G. G\n\n\n \n\n\n\n Molecular Pharmaceutics, 18(4): 1656–1665. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kannan2021adsorption,\n  title={Adsorption and Aggregation of Monoclonal Antibodies at Silicone Oil--Water Interfaces},\n  author={Kannan, Aadithya and Shieh, Ian C and Negulescu, Patrick G and Chandran Suja, Vineeth and Fuller, Gerald G},\n  journal={Molecular Pharmaceutics},\n  volume={18},\n  number={4},\n  pages={1656--1665},\n  year={2021},\n  publisher={ACS Publications}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Dynamics of freely suspended drops translating through miscible environments.\n \n \n \n\n\n \n Mossige, E. J.; Chandran Suja, V.; Walls, D. J; and Fuller, G. G\n\n\n \n\n\n\n Physics of Fluids, 33(3): 033106. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{mossige2021dynamics,\n  title={Dynamics of freely suspended drops translating through miscible environments},\n  author={Mossige, Endre Joachim and Chandran Suja, Vineeth and Walls, Daniel J and Fuller, Gerald G},\n  journal={Physics of Fluids},\n  volume={33},\n  number={3},\n  pages={033106},\n  year={2021},\n  publisher={AIP Publishing LLC}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Interfacial and Cohesive Properties of Corneal Epithelium.\n \n \n \n\n\n \n Liu, C.; and Fuller, G. G\n\n\n \n\n\n\n Biophysical Journal, 120(3): 169a. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{liu2021interfacial,\n  title={Interfacial and Cohesive Properties of Corneal Epithelium},\n  author={Liu, Chunzi and Fuller, Gerald G},\n  journal={Biophysical Journal},\n  volume={120},\n  number={3},\n  pages={169a},\n  year={2021},\n  publisher={Elsevier}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Engineering Insulin Cold Chain Resilience to Improve Global Access.\n \n \n \n\n\n \n Maikawa, C. L; Mann, J. L; Kannan, A.; Meis, C. M; Ou, B. S; Smith, A. A.; Fuller, G. G; Maahs, D. M; and Appel, E. A\n\n\n \n\n\n\n bioRxiv. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{maikawa2021engineering,\n  title={Engineering Insulin Cold Chain Resilience to Improve Global Access},\n  author={Maikawa, Caitlin L and Mann, Joseph L and Kannan, Aadithya and Meis, Catherine M and Ou, Ben S and Smith, Anton AA and Fuller, Gerald G and Maahs, David M and Appel, Eric A},\n  journal={bioRxiv},\n  year={2021},\n  publisher={Cold Spring Harbor Laboratory}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Axisymmetry breaking, chaos, and symmetry recovery in bubble film thickness profiles due to evaporation-induced Marangoni flows.\n \n \n \n\n\n \n Chandran Suja, V; Hadidi, A; Kannan, A; and Fuller, G.\n\n\n \n\n\n\n Physics of Fluids, 33(1): 012112. 2021.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chandran2021axisymmetry,\n  title={Axisymmetry breaking, chaos, and symmetry recovery in bubble film thickness profiles due to evaporation-induced Marangoni flows},\n  author={Chandran Suja, V and Hadidi, A and Kannan, A and Fuller, GG},\n  journal={Physics of Fluids},\n  volume={33},\n  number={1},\n  pages={012112},\n  year={2021},\n  publisher={AIP Publishing LLC}\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2020\n \n \n (23)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Instability and Symmetry Breaking in Binary Evaporating Thin Films over a Curved Substrate.\n \n \n \n\n\n \n Shi, X.; Rodriguez-Hakim, M.; Fuller, G.; and Shaqfeh, E.\n\n\n \n\n\n\n In 2020 Virtual AIChE Annual Meeting, 2020. AIChE\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@inproceedings{shi2020instability,\n  title={Instability and Symmetry Breaking in Binary Evaporating Thin Films over a Curved Substrate},\n  author={Shi, Xingyi and Rodriguez-Hakim, Mariana and Fuller, Gerald and Shaqfeh, Eric},\n  booktitle={2020 Virtual AIChE Annual Meeting},\n  year={2020},\n  organization={AIChE}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n In Vitro Monitoring of Drainage and Dewetting Dynamics over Contact Lenses Coated with Wetting Agents.\n \n \n \n\n\n \n Verma, A.; Suja, V. C.; Zhang, Y.; Joslin, S.; Mossige, E. J.; Sinha, D.; Scales, C. W; and Fuller, G.\n\n\n \n\n\n\n In 2020 Virtual AIChE Annual Meeting, 2020. AIChE\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@inproceedings{verma2020vitro,\n  title={In Vitro Monitoring of Drainage and Dewetting Dynamics over Contact Lenses Coated with Wetting Agents},\n  author={Verma, Archana and Suja, Vineeth Chandran and Zhang, Yong and Joslin, Scott and Mossige, Endre Joachim and Sinha, Dola and Scales, Charles W and Fuller, Gerald},\n  booktitle={2020 Virtual AIChE Annual Meeting},\n  year={2020},\n  organization={AIChE}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Stability of binary evaporating thin films.\n \n \n \n\n\n \n Shi, X.; Shaqfeh, E.; Rodriguez-Hakim, M.; and Fuller, G.\n\n\n \n\n\n\n Bulletin of the American Physical Society. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{shi2020stability,\n  title={Stability of binary evaporating thin films},\n  author={Shi, Xingyi and Shaqfeh, Eric and Rodriguez-Hakim, Mariana and Fuller, Gerald},\n  journal={Bulletin of the American Physical Society},\n  year={2020},\n  publisher={APS}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Freely Suspended Drops Rising in Miscible Environments.\n \n \n \n\n\n \n Mossige, E. J.; Chandran Suja, V.; Walls, D.; and Fuller, G.\n\n\n \n\n\n\n Bulletin of the American Physical Society. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{mossige2020freely,\n  title={Freely Suspended Drops Rising in Miscible Environments},\n  author={Mossige, Endre Joachim and Chandran Suja, Vinny and Walls, Daniel and Fuller, Gerald},\n  journal={Bulletin of the American Physical Society},\n  year={2020},\n  publisher={APS}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Rinsing and Mixing Flows of Miscible Liquids.\n \n \n \n\n\n \n Fuller, G.\n\n\n \n\n\n\n Bulletin of the American Physical Society. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller2020rinsing,\n  title={Rinsing and Mixing Flows of Miscible Liquids},\n  author={Fuller, Gerald},\n  journal={Bulletin of the American Physical Society},\n  year={2020},\n  publisher={APS}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Understanding the adsorption and potential tear film stability properties of recombinant human lubricin and bovine submaxillary mucins in an in vitro tear film model.\n \n \n \n\n\n \n Rabiah, N. I; Sato, Y.; Kannan, A.; Kress, W.; Straube, F.; and Fuller, G. G\n\n\n \n\n\n\n Colloids and Surfaces B: Biointerfaces, 195: 111257. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{rabiah2020understanding,\n  title={Understanding the adsorption and potential tear film stability properties of recombinant human lubricin and bovine submaxillary mucins in an in vitro tear film model},\n  author={Rabiah, Noelle I and Sato, Yasunori and Kannan, Aadithya and Kress, Wolfgang and Straube, Frank and Fuller, Gerald G},\n  journal={Colloids and Surfaces B: Biointerfaces},\n  volume={195},\n  pages={111257},\n  year={2020},\n  publisher={Elsevier}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Surface Energy and Separation Mechanics of Droplet Interface Phospholipid Bilayers.\n \n \n \n\n\n \n Huang, Y.; Suja, V. C.; Tajuelo, J.; and Fuller, G. G\n\n\n \n\n\n\n arXiv preprint arXiv:2010.13237. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{huang2020surface,\n  title={Surface Energy and Separation Mechanics of Droplet Interface Phospholipid Bilayers},\n  author={Huang, Yaoqi and Suja, Vineeth Chandran and Tajuelo, Javier and Fuller, Gerald G},\n  journal={arXiv preprint arXiv:2010.13237},\n  year={2020}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n In-use interfacial stability of monoclonal antibody formulations diluted in saline iv bags.\n \n \n \n\n\n \n Kannan, A.; Shieh, I. C; Hristov, P.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Pharmaceutical Sciences. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kannan2020use,\n  title={In-use interfacial stability of monoclonal antibody formulations diluted in saline iv bags},\n  author={Kannan, Aadithya and Shieh, Ian C and Hristov, Petar and Fuller, Gerald G},\n  journal={Journal of Pharmaceutical Sciences},\n  year={2020},\n  publisher={Elsevier}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n The Froth Thickens.\n \n \n \n\n\n \n Fuller, G.; and Suja, V. C.\n\n\n \n\n\n\n Physics, 13: 162. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller2020froth,\n  title={The Froth Thickens},\n  author={Fuller, Gerald and Suja, Vinny Chandran},\n  journal={Physics},\n  volume={13},\n  pages={162},\n  year={2020},\n  publisher={APS}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Single bubble and drop techniques for characterizing foams and emulsions.\n \n \n \n\n\n \n Suja, V C.; Rodrı́guez-Hakim, Mariana; Tajuelo, J.; and Fuller, G. G\n\n\n \n\n\n\n Advances in Colloid and Interface Science,102295. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{suja2020single,\n  title={Single bubble and drop techniques for characterizing foams and emulsions},\n  author={Suja, V Chandran and Rodr{\\'\\i}guez-Hakim, Mariana and Tajuelo, Javier and Fuller, Gerald G},\n  journal={Advances in Colloid and Interface Science},\n  pages={102295},\n  year={2020},\n  publisher={Elsevier}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Asphaltene-induced spontaneous emulsification: Effects of interfacial co-adsorption and viscoelasticity.\n \n \n \n\n\n \n Rodrı́guez-Hakim, Mariana; Anand, S.; Tajuelo, J.; Yao, Z.; Kannan, A.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Rheology, 64(4): 799–816. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{rodriguez2020asphaltene,\n  title={Asphaltene-induced spontaneous emulsification: Effects of interfacial co-adsorption and viscoelasticity},\n  author={Rodr{\\'\\i}guez-Hakim, Mariana and Anand, Satyam and Tajuelo, Javier and Yao, Zhen and Kannan, Aadithya and Fuller, Gerald G},\n  journal={Journal of Rheology},\n  volume={64},\n  number={4},\n  pages={799--816},\n  year={2020},\n  publisher={The Society of Rheology}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Evaporation-induced Rayleigh–Taylor instabilities in polymer solutions.\n \n \n \n\n\n \n Mossige, E.; Chandran Suja, V; Islamov, M; Wheeler, S.; and Fuller, G. G\n\n\n \n\n\n\n Philosophical Transactions of the Royal Society A, 378(2174): 20190533. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{mossige2020evaporation,\n  title={Evaporation-induced Rayleigh--Taylor instabilities in polymer solutions},\n  author={Mossige, EJ and Chandran Suja, V and Islamov, M and Wheeler, SF and Fuller, Gerald G},\n  journal={Philosophical Transactions of the Royal Society A},\n  volume={378},\n  number={2174},\n  pages={20190533},\n  year={2020},\n  publisher={The Royal Society Publishing}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Bubble coalescence in worm-like micellar solutions.\n \n \n \n\n\n \n Chandran Suja, V.; Kannan, A.; Kubicka, B.; Hadidi, A.; and Fuller, G. G\n\n\n \n\n\n\n arXiv e-prints,arXiv–2006. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chandran2020bubble,\n  title={Bubble coalescence in worm-like micellar solutions},\n  author={Chandran Suja, Vineeth and Kannan, Aadithya and Kubicka, Bruce and Hadidi, Alex and Fuller, Gerald G},\n  journal={arXiv e-prints},\n  pages={arXiv--2006},\n  year={2020}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Tuning corneal epithelial cell adhesive strength with varying crosslinker content in silicone hydrogel materials.\n \n \n \n\n\n \n Liu, C.; Scales, C. W; and Fuller, G. G\n\n\n \n\n\n\n Translational Vision Science & Technology, 9(6): 3–3. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{liu2020tuning,\n  title={Tuning corneal epithelial cell adhesive strength with varying crosslinker content in silicone hydrogel materials},\n  author={Liu, Chunzi and Scales, Charles W and Fuller, Gerald G},\n  journal={Translational Vision Science \\& Technology},\n  volume={9},\n  number={6},\n  pages={3--3},\n  year={2020},\n  publisher={The Association for Research in Vision and Ophthalmology}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Polymeric-nanofluids stabilized emulsions: Interfacial versus bulk rheology.\n \n \n \n\n\n \n Kamkar, M.; Bazazi, P.; Kannan, A.; Suja, V. C.; Hejazi, S. H.; Fuller, G. G; and Sundararaj, U.\n\n\n \n\n\n\n Journal of Colloid and Interface Science. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kamkar2020polymeric,\n  title={Polymeric-nanofluids stabilized emulsions: Interfacial versus bulk rheology},\n  author={Kamkar, Milad and Bazazi, Parisa and Kannan, Aadithya and Suja, Vineeth Chandran and Hejazi, Seyed Hossein and Fuller, Gerald G and Sundararaj, Uttandaraman},\n  journal={Journal of Colloid and Interface Science},\n  year={2020},\n  publisher={Elsevier}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Foam stability in filtered lubricants containing antifoams.\n \n \n \n \n\n\n \n Suja, V C.; Kar, A; Cates, W; Remmert, S.; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science. 2020.\n \n\n\n\n
\n\n\n\n \n \n \"FoamPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{suja2020foam,\n  title={Foam stability in filtered lubricants containing antifoams},\n  author={Suja, V Chandran and Kar, A and Cates, W and Remmert, SM and Fuller, GG},\n  journal={Journal of Colloid and Interface Science},\n  year={2020},\n  publisher={Elsevier},\n  doi={10.1016/j.jcis.2020.01.103},\n  URL={https://www.sciencedirect.com/science/article/pii/S002197972030117X},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Viscoelastic interfaces comprising of cellulose nanocrystals and lauroyl ethyl arginate for enhanced foam stability.\n \n \n \n\n\n \n Czakaj, A.; Kannan, A.; Wiśniewska, A.; Grześ, G.; Krzan, M.; Warszyński, P.; and Fuller, G. G\n\n\n \n\n\n\n Soft Matter, 16(16): 3981–3990. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{czakaj2020viscoelastic,\n  title={Viscoelastic interfaces comprising of cellulose nanocrystals and lauroyl ethyl arginate for enhanced foam stability},\n  author={Czakaj, Agnieszka and Kannan, Aadithya and Wi{\\'s}niewska, Agnieszka and Grze{\\'s}, Gabriela and Krzan, Marcel and Warszy{\\'n}ski, Piotr and Fuller, Gerald G},\n  journal={Soft Matter},\n  volume={16},\n  number={16},\n  pages={3981--3990},\n  year={2020},\n  publisher={The Royal Society of Chemistry}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Surfactant-laden bubble dynamics under porous polymer films.\n \n \n \n\n\n \n Kannan, A.; Hristov, P.; Li, J.; Zawala, J.; Gao, P.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Colloid and Interface Science. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kannan2020surfactant,\n  title={Surfactant-laden bubble dynamics under porous polymer films},\n  author={Kannan, Aadithya and Hristov, Petar and Li, Jin and Zawala, Jan and Gao, Ping and Fuller, Gerald G},\n  journal={Journal of Colloid and Interface Science},\n  year={2020},\n  publisher={Elsevier}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Symmetry breaking and chaos in evaporation driven Marangoni flows over bubbles.\n \n \n \n\n\n \n Suja, V. C.; Hadidi, A.; Kannan, A.; and Fuller, G. G\n\n\n \n\n\n\n arXiv preprint arXiv:2004.09752. 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{suja2020symmetry,\n  title={Symmetry breaking and chaos in evaporation driven Marangoni flows over bubbles},\n  author={Suja, Vineeth Chandran and Hadidi, Alex and Kannan, Aadithya and Fuller, Gerald G},\n  journal={arXiv preprint arXiv:2004.09752},\n  year={2020}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Oscillatory spontaneous dimpling in evaporating curved thin films.\n \n \n \n \n\n\n \n Shi, X.; Fuller, G. G; and Shaqfeh, E. S.\n\n\n \n\n\n\n Journal of Fluid Mechanics. 2020.\n \n\n\n\n
\n\n\n\n \n \n \"OscillatoryPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{shi2020oscillatory,\n  title={Oscillatory spontaneous dimpling in evaporating curved thin films},\n  author={Shi, Xingyi and Fuller, Gerald G and Shaqfeh, Eric SG},\n  journal={Journal of Fluid Mechanics},\n  year={2020},\n  publisher={Cambridge University Press},\n  doi={10.1017/jfm.2020.92},\n  URL={https://www.cambridge.org/core/journals/journal-of-fluid-mechanics/article/oscillatory-spontaneous-dimpling-in-evaporating-curved-thin-films/6F46169EB5052CA9F94396AFF8980EEB},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Perpendicular alignment of lymphatic endothelial cells in response to spatial gradients in wall shear stress.\n \n \n \n \n\n\n \n Michalaki, E.; Surya, V. N; Fuller, G. G; and Dunn, A. R\n\n\n \n\n\n\n Communications Biology. 2020.\n \n\n\n\n
\n\n\n\n \n \n \"PerpendicularPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{michalaki2020perpendicular,\n  title={Perpendicular alignment of lymphatic endothelial cells in response to spatial gradients in wall shear stress},\n  author={Michalaki, Eleftheria and Surya, Vinay N and Fuller, Gerald G and Dunn, Alexander R},\n  journal={Communications Biology},\n  year={2020},\n  publisher={Nature Publishing Group},\n  doi={10.1038/s42003-019-0732-8},\n  URL={https://www.nature.com/articles/s42003-019-0732-8},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Mechanical and microstructural insights of Vibrio cholerae and Escherichia coli dual-species biofilm at the air-liquid interface.\n \n \n \n \n\n\n \n Abriat, C.; Enriquez, K.; Virgilio, N.; Cegelski, L.; Fuller, G. G; Daigle, F.; and Heuzey, M.\n\n\n \n\n\n\n Colloids and Surfaces B: Biointerfaces. 2020.\n \n\n\n\n
\n\n\n\n \n \n \"MechanicalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{abriat2020mechanical,\n  title={Mechanical and microstructural insights of Vibrio cholerae and Escherichia coli dual-species biofilm at the air-liquid interface},\n  author={Abriat, Cl{\\'e}mence and Enriquez, Kyle and Virgilio, Nick and Cegelski, Lynette and Fuller, Gerald G and Daigle, France and Heuzey, Marie-Claude},\n  journal={Colloids and Surfaces B: Biointerfaces},\n  year={2020},\n  publisher={Elsevier},\n  doi={10.1016/j.colsurfb.2020.110786},\n  URL={https://www.sciencedirect.com/science/article/pii/S0927776520300163},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Hyperspectral imaging for dynamic thin film interferometry.\n \n \n \n\n\n \n Chandran Suja, V.; Sentmanat, J.; Hofmann, G.; Scales, C.; and Fuller, G. G\n\n\n \n\n\n\n Scientific Reports, 10(1). 2020.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chandran2020hyperspectral,\n  title={Hyperspectral imaging for dynamic thin film interferometry.},\n  author={Chandran Suja, Vineeth and Sentmanat, John and Hofmann, Gregory and Scales, Charles and Fuller, Gerald G},\n  journal={Scientific Reports},\n  volume={10},\n  number={1},\n  year={2020}\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2019\n \n \n (11)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Binding partner-and force-promoted changes in $α$E-catenin conformation probed by native cysteine labeling.\n \n \n \n \n\n\n \n Terekhova, K.; Pokutta, S.; Kee, Y. S; Li, J.; Tajkhorshid, E.; Fuller, G.; Dunn, A. R; and Weis, W. I\n\n\n \n\n\n\n Scientific reports. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"BindingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{terekhova2019binding,\n  title={Binding partner-and force-promoted changes in $\\alpha$E-catenin conformation probed by native cysteine labeling},\n  author={Terekhova, Ksenia and Pokutta, Sabine and Kee, Yee S and Li, Jing and Tajkhorshid, Emad and Fuller, Gerald and Dunn, Alexander R and Weis, William I},\n  journal={Scientific reports},\n  year={2019},\n  publisher={Nature Publishing Group},\n  doi={10.1038/s41598-019-51816-3},\n  URL={https://www.nature.com/articles/s41598-019-51816-3},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Mechanical Properties of Solidifying Assemblies of Nanoparticle Surfactants at the Oil–Water Interface.\n \n \n \n \n\n\n \n Toor, A.; Forth, J.; Bochner de Araujo, S.; Merola, M. C.; Jiang, Y.; Liu, X.; Chai, Y.; Hou, H.; Ashby, P. D; Fuller, G. G; and others\n\n\n \n\n\n\n Langmuir. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"MechanicalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{toor2019mechanical,\n  title={Mechanical Properties of Solidifying Assemblies of Nanoparticle Surfactants at the Oil--Water Interface},\n  author={Toor, Anju and Forth, Joe and Bochner de Araujo, Simone and Merola, Maria Consiglia and Jiang, Yufeng and Liu, Xubo and Chai, Yu and Hou, Honghao and Ashby, Paul D and Fuller, Gerald G and others},\n  journal={Langmuir},\n  year={2019},\n  publisher={ACS Publications},\n  doi={10.1021/acs.langmuir.9b01575},\n  URL={https://pubs.acs.org/doi/abs/10.1021/acs.langmuir.9b01575},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Ablation of water drops suspended in asphaltene/heptol solutions due to spontaneous emulsification.\n \n \n \n \n\n\n \n de Araujo, S B.; Reyssat, M; Monteux, C; and Fuller, G.\n\n\n \n\n\n\n Science advances. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"AblationPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{de2019ablation,\n  title={Ablation of water drops suspended in asphaltene/heptol solutions due to spontaneous emulsification},\n  author={de Araujo, S Bochner and Reyssat, M and Monteux, C and Fuller, GG},\n  journal={Science advances},\n  year={2019},\n  publisher={American Association for the Advancement of Science},\n  doi={10.1126/sciadv.aax8227},\n  URL={https://advances.sciencemag.org/content/5/10/eaax8227.abstract},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Carbon compositional analysis of hydrogel contact lenses by solid-state NMR spectroscopy.\n \n \n \n \n\n\n \n Rabiah, N. I; Romaniuk, J. A.; Fuller, G. G; Scales, C. W; and Cegelski, L.\n\n\n \n\n\n\n Solid state nuclear magnetic resonance. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"CarbonPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{rabiah2019carbon,\n  title={Carbon compositional analysis of hydrogel contact lenses by solid-state NMR spectroscopy},\n  author={Rabiah, Noelle I and Romaniuk, Joseph AH and Fuller, Gerald G and Scales, Charles W and Cegelski, Lynette},\n  journal={Solid state nuclear magnetic resonance},\n  year={2019},\n  publisher={Elsevier},\n  doi={10.1016/j.ssnmr.2019.07.003},\n  URL={https://www.sciencedirect.com/science/article/pii/S0926204019300608},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Spreading of rinsing liquids across a horizontal rotating substrate.\n \n \n \n \n\n\n \n Walls, D. J; Ylitalo, A. S; Mui, D. S.; Frostad, J. M; and Fuller, G. G\n\n\n \n\n\n\n Physical Review Fluids. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"SpreadingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{walls2019spreading,\n  title={Spreading of rinsing liquids across a horizontal rotating substrate},\n  author={Walls, Daniel J and Ylitalo, Andrew S and Mui, David SL and Frostad, John M and Fuller, Gerald G},\n  journal={Physical Review Fluids},\n  year={2019},\n  publisher={APS},\n  doi={10.1103/PhysRevFluids.4.084102},\n  URL={https://journals.aps.org/prfluids/abstract/10.1103/PhysRevFluids.4.084102},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Evolution of rivulets during spreading of an impinging water jet on a rotating, precoated substrate.\n \n \n \n \n\n\n \n Ylitalo, A. S; Walls, D. J; Mui, D. S.; Frostad, J. M; and Fuller, G. G\n\n\n \n\n\n\n Physics of Fluids. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"EvolutionPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{ylitalo2019evolution,\n  title={Evolution of rivulets during spreading of an impinging water jet on a rotating, precoated substrate},\n  author={Ylitalo, Andrew S and Walls, Daniel J and Mui, David SL and Frostad, John M and Fuller, Gerald G},\n  journal={Physics of Fluids},\n  year={2019},\n  publisher={AIP Publishing},\n  doi={10.1063/1.5109806},\n  URL={https://aip.scitation.org/doi/full/10.1063/1.5109806},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Unraveling Escherichia coli’s Cloak: Identification of Phosphoethanolamine Cellulose, Its Functions, and Applications.\n \n \n \n \n\n\n \n Jeffries, J.; Fuller, G. G; and Cegelski, L.\n\n\n \n\n\n\n Microbiology Insights. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"UnravelingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{jeffries2019unraveling,\n  title={Unraveling Escherichia coli’s Cloak: Identification of Phosphoethanolamine Cellulose, Its Functions, and Applications},\n  author={Jeffries, Jamie and Fuller, Gerald G and Cegelski, Lynette},\n  journal={Microbiology Insights},\n  year={2019},\n  publisher={SAGE Publications Sage UK},\n  doi={10.1177/1178636119865234},\n  URL={https://journals.sagepub.com/doi/full/10.1177/1178636119865234},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The influence of protein deposition on contact lens tear film stability.\n \n \n \n \n\n\n \n Rabiah, N. I; Scales, C. W; and Fuller, G. G\n\n\n \n\n\n\n Colloids and Surfaces B: Biointerfaces. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{rabiah2019influence,\n  title={The influence of protein deposition on contact lens tear film stability},\n  author={Rabiah, Noelle I and Scales, Charles W and Fuller, Gerald G},\n  journal={Colloids and Surfaces B: Biointerfaces},\n  year={2019},\n  publisher={Elsevier},\n  doi={10.1016/j.colsurfb.2019.04.051},\n  URL={https://www.sciencedirect.com/science/article/pii/S0927776519302826},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Linking aggregation and interfacial properties in monoclonal antibody-surfactant formulations.\n \n \n \n \n\n\n \n Kannan, A.; Shieh, I. C; and Fuller, G. G\n\n\n \n\n\n\n Journal of Colloid and Interface Science. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"LinkingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kannan2019linking,\n  title={Linking aggregation and interfacial properties in monoclonal antibody-surfactant formulations},\n  author={Kannan, Aadithya and Shieh, Ian C and Fuller, Gerald G},\n  journal={Journal of Colloid and Interface Science},\n  year={2019},\n  publisher={Elsevier},\n  doi={10.1016/j.jcis.2019.04.060},\n  URL={https://www.sciencedirect.com/science/article/pii/S0021979719304904},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Evaporation-driven solutocapillary flow of thin liquid films over curved substrates.\n \n \n \n \n\n\n \n Rodriguez-Hakim, M.; Barakat, J. M; Shi, X.; Shaqfeh, E. S.; and Fuller, G. G\n\n\n \n\n\n\n Physical Review Fluids, 4(3): 034002. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"Evaporation-drivenPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{rodriguez2019evaporation,\n  title={Evaporation-driven solutocapillary flow of thin liquid films over curved substrates},\n  author={Rodriguez-Hakim, Mariana and Barakat, Joseph M and Shi, Xingyi and Shaqfeh, Eric SG and Fuller, Gerald G},\n  journal={Physical Review Fluids},\n  volume={4},\n  number={3},\n  pages={034002},\n  year={2019},\n  publisher={APS},\n  doi={10.1103/PhysRevFluids.4.034002},\n  URL={https://journals.aps.org/prfluids/abstract/10.1103/PhysRevFluids.4.034002},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Lymphatic endothelial cell calcium pulses are sensitive to spatial gradients in wall shear stress.\n \n \n \n \n\n\n \n Surya, V. N; Michalaki, E.; Fuller, G. G; and Dunn, A. R\n\n\n \n\n\n\n Molecular biology of the cell,mbc–E18. 2019.\n \n\n\n\n
\n\n\n\n \n \n \"LymphaticPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{surya2019lymphatic,\n  title={Lymphatic endothelial cell calcium pulses are sensitive to spatial gradients in wall shear stress},\n  author={Surya, Vinay N and Michalaki, Eleftheria and Fuller, Gerald G and Dunn, Alexander R},\n  journal={Molecular biology of the cell},\n  pages={mbc--E18},\n  year={2019},\n  publisher={Am Soc Cell Biol},\n  doi={10.1091/mbc.E18-10-0618},\n  URL={https://www.molbiolcell.org/doi/full/10.1091/mbc.E18-10-0618},\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2018\n \n \n (5)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Influence of interfacial elasticity on liquid entrainment in thin foam films.\n \n \n \n \n\n\n \n Lin, G.; Frostad, J. M; and Fuller, G. G\n\n\n \n\n\n\n Physical Review Fluids, 3(11): 114001. 2018.\n \n\n\n\n
\n\n\n\n \n \n \"InfluencePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lin2018influence,\n  title={Influence of interfacial elasticity on liquid entrainment in thin foam films},\n  author={Lin, Gigi and Frostad, John M and Fuller, Gerald G},\n  journal={Physical Review Fluids},\n  volume={3},\n  number={11},\n  pages={114001},\n  year={2018},\n  publisher={APS},\n  doi={10.1103/PhysRevFluids.3.114001},\n  URL={https://journals.aps.org/prfluids/abstract/10.1103/PhysRevFluids.3.114001},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Phosphoethanolamine cellulose enhances curli-mediated adhesion of uropathogenic Escherichia coli to bladder epithelial cells.\n \n \n \n \n\n\n \n Hollenbeck, E. C; Antonoplis, A.; Chai, C.; Thongsomboon, W.; Fuller, G. G; and Cegelski, L.\n\n\n \n\n\n\n Proceedings of the National Academy of Sciences, 115(40): 10106–10111. 2018.\n \n\n\n\n
\n\n\n\n \n \n \"PhosphoethanolaminePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hollenbeck2018phosphoethanolamine,\n  title={Phosphoethanolamine cellulose enhances curli-mediated adhesion of uropathogenic Escherichia coli to bladder epithelial cells},\n  author={Hollenbeck, Emily C and Antonoplis, Alexandra and Chai, Chew and Thongsomboon, Wiriya and Fuller, Gerald G and Cegelski, Lynette},\n  journal={Proceedings of the National Academy of Sciences},\n  volume={115},\n  number={40},\n  pages={10106--10111},\n  year={2018},\n  doi={10.1073/pnas.1801564115},\n  URL={https://www.pnas.org/content/115/40/10106},\n  publisher={National Acad Sciences},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Evaporation-induced foam stabilization in lubricating oils.\n \n \n \n \n\n\n \n Chandran Suja, V.; Kar, A.; Cates, W.; Remmert, S. M.; Savage, P. D.; and Fuller, G. G.\n\n\n \n\n\n\n Proceedings of the National Academy of Sciences, 115(31): 7919–7924. 2018.\n \n\n\n\n
\n\n\n\n \n \n \"Evaporation-inducedPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{Suja2018Evaporation,\n\ttitle={Evaporation-induced foam stabilization in lubricating oils},\n\tauthor={Chandran Suja, V. and Kar, A. and Cates, W. and Remmert, S. M. and Savage, P. D. and Fuller, G. G.},\n\tjournal={Proceedings of the National Academy of Sciences},\n\tvolume={115},\n\tnumber={31},\n\tpages={7919--7924},\n\tyear={2018},\n\tdoi={10.1073/pnas.1805645115},\n\tpublisher={National Academy of Sciences},\n\tURL={http://www.pnas.org/content/115/31/7919}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The shape evolution of liquid droplets in miscible environments.\n \n \n \n \n\n\n \n Walls, D. J; Meiburg, E.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Fluid Mechanics, 852: 422–452. 2018.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{walls2018shape,\n  title={The shape evolution of liquid droplets in miscible environments},\n  author={Walls, Daniel J and Meiburg, Eckart and Fuller, Gerald G},\n  journal={Journal of Fluid Mechanics},\n  volume={852},\n  pages={422--452},\n  year={2018},\n  publisher={Cambridge University Press},\nURL={https://www.cambridge.org/core/journals/journal-of-fluid-mechanics/article/shape-evolution-of-liquid-droplets-in-miscible-environments/65E77D6FF3CE35EEB89C11DBE21601D7},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Monoclonal Antibody Interfaces: Dilatation Mechanics and Bubble Coalescence.\n \n \n \n \n\n\n \n Kannan, A.; Shieh, I. C.; Leiske, D. L.; and Fuller, G. G.\n\n\n \n\n\n\n Langmuir, 34(2): 630-638. 2018.\n \n\n\n\n
\n\n\n\n \n \n \"MonoclonalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{doi:10.1021/acs.langmuir.7b03790,\nauthor = {Kannan, Aadithya and Shieh, Ian C. and Leiske, Danielle L. and Fuller, Gerald G.},\ntitle = {Monoclonal Antibody Interfaces: Dilatation Mechanics and Bubble Coalescence},\njournal = {Langmuir},\nvolume = {34},\nnumber = {2},\npages = {630-638},\nyear = {2018},\ndoi = {10.1021/acs.langmuir.7b03790},\nURL = {https://doi.org/10.1021/acs.langmuir.7b03790},\neprint = { https://doi.org/10.1021/acs.langmuir.7b03790}\n}\n\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2017\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Droplet coalescence and spontaneous emulsification in the presence of asphaltene adsorption.\n \n \n \n \n\n\n \n Bochner de Araujo, S.; Merola, M. C; Vlassopoulos, D.; and Fuller, G. G\n\n\n \n\n\n\n Langmuir. 2017.\n \n\n\n\n
\n\n\n\n \n \n \"DropletPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{bochner2017droplet,\n  title={Droplet coalescence and spontaneous emulsification in the presence of asphaltene adsorption},\n  author={Bochner de Araujo, Simone and Merola, Maria C and Vlassopoulos, Dimitris and Fuller, Gerald G},\n  journal={Langmuir},\n  year={2017},\n  publisher={ACS Publications},\nURL={http://pubs.acs.org/doi/abs/10.1021/acs.langmuir.7b02638},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n DACH1 stimulates shear stress-guided endothelial cell migration and coronary artery growth through the CXCL12–CXCR4 signaling axis.\n \n \n \n \n\n\n \n Chang, A. H; Raftrey, B. C; D'Amato, G.; Surya, V. N; Poduri, A.; Chen, H. I; Goldstone, A. B; Woo, J.; Fuller, G. G; Dunn, A. R; and others\n\n\n \n\n\n\n Genes & Development. 2017.\n \n\n\n\n
\n\n\n\n \n \n \"DACH1Paper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chang2017dach1,\n  title={DACH1 stimulates shear stress-guided endothelial cell migration and coronary artery growth through the CXCL12--CXCR4 signaling axis},\n  author={Chang, Andrew H and Raftrey, Brian C and D'Amato, Gaetano and Surya, Vinay N and Poduri, Aruna and Chen, Heidi I and Goldstone, Andrew B and Woo, Joseph and Fuller, Gerald G and Dunn, Alexander R and others},\n  journal={Genes \\& Development},\n  year={2017},\n  publisher={Cold Spring Harbor Lab},\nURL={http://genesdev.cshlp.org/content/early/2017/08/04/gad.301549.117.full.pdf+html}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Interfacial mechanisms for stability of surfactant-laden films.\n \n \n \n \n\n\n \n Bhamla, M S.; Chai, C.; Alvarez-Valenzuela, M. A; Tajuelo, J.; and Fuller, G. G\n\n\n \n\n\n\n PLOS ONE, 12(5): e0175753. 2017.\n \n\n\n\n
\n\n\n\n \n \n \"InterfacialPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{bhamla2017interfacial,\n  title={Interfacial mechanisms for stability of surfactant-laden films},\n  author={Bhamla, M Saad and Chai, Chew and Alvarez-Valenzuela, Marco A and Tajuelo, Javier and Fuller, Gerald G},\n  journal={PLOS ONE},\n  volume={12},\n  number={5},\n  pages={e0175753},\n  year={2017},\n  publisher={Public Library of Science},\nURL= {http://journals.plos.org/plosone/article?id=10.1371/journal.pone.0175753}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Temperature controlled tensiometry using droplet microfluidics.\n \n \n \n \n\n\n \n Lee, D.; Fang, C.; Ravan, A. S; Fuller, G. G; and Shen, A. Q\n\n\n \n\n\n\n Lab on a Chip. 2017.\n \n\n\n\n
\n\n\n\n \n \n \"TemperaturePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lee2017temperature,\n  title={Temperature controlled tensiometry using droplet microfluidics},\n  author={Lee, Doojin and Fang, Cifeng and Ravan, Aniket S and Fuller, Gerald G and Shen, Amy Q},\n  journal={Lab on a Chip},\n  doi={10.1039/C6LC01384H},\n  year={2017},\n  publisher={Royal Society of Chemistry},\n  URL= {http://pubs.rsc.org/-/content/articlehtml/2017/lc/c6lc01384h},\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2016\n \n \n (11)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Sphingosine 1-phosphate receptor 1 regulates the directional migration of lymphatic endothelial cells in response to fluid shear stress.\n \n \n \n \n\n\n \n Surya, V. N.; Michalaki, E.; Huang, E. Y.; Fuller, G. G.; and Dunn, A. R.\n\n\n \n\n\n\n Journal of The Royal Society Interface, 13(125). 2016.\n \n\n\n\n
\n\n\n\n \n \n \"SphingosinePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article {Surya20160823,\n\tauthor = {Surya, Vinay N. and Michalaki, Eleftheria and Huang, Eva Y. and Fuller, Gerald G. and Dunn, Alexander R.},\n\ttitle = {Sphingosine 1-phosphate receptor 1 regulates the directional migration of lymphatic endothelial cells in response to fluid shear stress},\n\tvolume = {13},\n\tnumber = {125},\n\tyear = {2016},\n\tdoi = {10.1098/rsif.2016.0823},\n\tpublisher = {The Royal Society},\n\tabstract = {The endothelial cells that line blood and lymphatic vessels undergo complex, collective migration and rearrangement processes during embryonic development, and are known to be exquisitely responsive to fluid flow. At present, the molecular mechanisms by which endothelial cells sense fluid flow remain incompletely understood. Here, we report that both the G-protein-coupled receptor sphingosine 1-phosphate receptor 1 (S1PR1) and its ligand sphingosine 1-phosphate (S1P) are required for collective upstream migration of human lymphatic microvascular endothelial cells in an in vitro setting. These findings are consistent with a model in which signalling via S1P and S1PR1 are integral components in the response of lymphatic endothelial cells to the stimulus provided by fluid flow.},\n\tissn = {1742-5689},\n\tURL = {http://rsif.royalsocietypublishing.org/content/13/125/20160823},\n\teprint = {http://rsif.royalsocietypublishing.org/content/13/125/20160823.full.pdf},\n\tjournal = {Journal of The Royal Society Interface}\n}\n\n
\n
\n\n\n
\n The endothelial cells that line blood and lymphatic vessels undergo complex, collective migration and rearrangement processes during embryonic development, and are known to be exquisitely responsive to fluid flow. At present, the molecular mechanisms by which endothelial cells sense fluid flow remain incompletely understood. Here, we report that both the G-protein-coupled receptor sphingosine 1-phosphate receptor 1 (S1PR1) and its ligand sphingosine 1-phosphate (S1P) are required for collective upstream migration of human lymphatic microvascular endothelial cells in an in vitro setting. These findings are consistent with a model in which signalling via S1P and S1PR1 are integral components in the response of lymphatic endothelial cells to the stimulus provided by fluid flow.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Impact of Compressibility on the Control of Bubble-Pressure Tensiometers.\n \n \n \n \n\n\n \n Chandran Suja, V.; Frostad, J. M.; and Fuller, G. G.\n\n\n \n\n\n\n Langmuir, 32(46): 12031-12038. 2016.\n \n\n\n\n
\n\n\n\n \n \n \"ImpactPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{Vineeth,\nauthor = {Chandran Suja, V. and Frostad, J. M. and Fuller, G. G.},\ntitle = {Impact of Compressibility on the Control of Bubble-Pressure Tensiometers},\njournal = {Langmuir},\nvolume = {32},\nnumber = {46},\npages = {12031-12038},\nyear = {2016},\ndoi = {10.1021/acs.langmuir.6b03258},\nURL = {\n        http://dx.doi.org/10.1021/acs.langmuir.6b03258\n},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Dynamic fluid-film interferometry as a predictor of bulk foam properties.\n \n \n \n \n\n\n \n Frostad, J. M.; Tammaro, D.; Santollani, L.; Bochner de Araujo, S.; and Fuller, G. G.\n\n\n \n\n\n\n Soft Matter, 12: 9266-9279. 2016.\n \n\n\n\n
\n\n\n\n \n \n \"DynamicPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@Article{C6SM01361A,\nauthor ={Frostad, John M. and Tammaro, Daniele and Santollani, Luciano and Bochner de Araujo, Simone and Fuller, Gerald G.},\ntitle  ={Dynamic fluid-film interferometry as a predictor of bulk foam properties},\njournal  ={Soft Matter},\nyear  ={2016},\nvolume  ={12},\nissue  ={46},\npages  ={9266-9279},\npublisher  ={The Royal Society of Chemistry},\ndoi  ={10.1039/C6SM01361A},\nurl  ={http://dx.doi.org/10.1039/C6SM01361A},\n\n}\n\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Multiplexed Fluid Flow Device to Study Cellular Response to Tunable Shear Stress Gradients.\n \n \n \n\n\n \n Ostrowski, M. A; Huang, E. Y; Surya, V. N; Poplawski, C.; Barakat, J. M; Lin, G. L; Fuller, G. G; and Dunn, A. R\n\n\n \n\n\n\n Annals of biomedical engineering,1–12. 2016.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{ostrowski2016multiplexed,\n  title={Multiplexed Fluid Flow Device to Study Cellular Response to Tunable Shear Stress Gradients},\n  author={Ostrowski, Maggie A and Huang, Eva Y and Surya, Vinay N and Poplawski, Charlotte and Barakat, Joseph M and Lin, Gigi L and Fuller, Gerald G and Dunn, Alexander R},\n  journal={Annals of biomedical engineering},\n  pages={1--12},\n  year={2016},\n  publisher={Springer}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Spreading of miscible liquids.\n \n \n \n\n\n \n Walls, D. J; Haward, S. J; Shen, A. Q; and Fuller, G. G\n\n\n \n\n\n\n Physical Review Fluids, 1(1): 013904. 2016.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{walls2016spreading,\n  title={Spreading of miscible liquids},\n  author={Walls, Daniel J and Haward, Simon J and Shen, Amy Q and Fuller, Gerald G},\n  journal={Physical Review Fluids},\n  volume={1},\n  number={1},\n  pages={013904},\n  year={2016},\n  publisher={APS}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Interfacial rheology of hydrogen-bonded polymer multilayers assembled at liquid interfaces. Influence of anchoring energy and hydrophobic interactions.\n \n \n \n\n\n \n Le Tirilly, S.; Tregouët, C.; Reyssat, M.; Bone, S.; Geffroy, C.; Fuller, G. G; Pantoustier, N.; Perrin, P.; and Monteux, C.\n\n\n \n\n\n\n Langmuir. 2016.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{le2016interfacial,\n  title={Interfacial rheology of hydrogen-bonded polymer multilayers assembled at liquid interfaces. Influence of anchoring energy and hydrophobic interactions},\n  author={Le Tirilly, Sandrine and Tregou{\\"e}t, Corentin and Reyssat, Mathilde and Bone, Stephane and Geffroy, C{\\'e}dric and Fuller, Gerald G and Pantoustier, Nadege and Perrin, Patrick and Monteux, Cecile},\n  journal={Langmuir},\n  year={2016},\n  publisher={ACS Publications}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Growth Kinetics and Mechanics of Hydrate Films by Interfacial Rheology.\n \n \n \n\n\n \n Leopércio, B. C; de Souza Mendes, P. R; and Fuller, G. G\n\n\n \n\n\n\n Langmuir, 32(17): 4203–4209. 2016.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{leopercio2016growth,\n  title={Growth Kinetics and Mechanics of Hydrate Films by Interfacial Rheology},\n  author={Leop{\\'e}rcio, Bruna C and de Souza Mendes, Paulo R and Fuller, Gerald G},\n  journal={Langmuir},\n  volume={32},\n  number={17},\n  pages={4203--4209},\n  year={2016},\n  publisher={ACS Publications}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Mechanical Behavior of a Bacillus subtilis Pellicle.\n \n \n \n\n\n \n Hollenbeck, E. C; Douarche, C.; Allain, J.; Roger, P.; Regeard, C.; Cegelski, L.; Fuller, G. G; and Raspaud, E.\n\n\n \n\n\n\n The Journal of Physical Chemistry B. 2016.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hollenbeck2016mechanical,\n  title={Mechanical Behavior of a Bacillus subtilis Pellicle},\n  author={Hollenbeck, Emily C and Douarche, Carine and Allain, Jean-Marc and Roger, Philippe and Regeard, Christophe and Cegelski, Lynette and Fuller, Gerald G and Raspaud, Eric},\n  journal={The Journal of Physical Chemistry B},\n  year={2016},\n  publisher={ACS Publications}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Instability and Breakup of Model Tear FilmsBreakup of Model Tear Films.\n \n \n \n\n\n \n Bhamla, M S.; Chai, C.; Rabiah, N. I; Frostad, J. M; and Fuller, G. G\n\n\n \n\n\n\n Investigative ophthalmology & visual science, 57(3): 949–958. 2016.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{bhamla2016instability,\n  title={Instability and Breakup of Model Tear FilmsBreakup of Model Tear Films},\n  author={Bhamla, M Saad and Chai, Chew and Rabiah, Noelle I and Frostad, John M and Fuller, Gerald G},\n  journal={Investigative ophthalmology \\& visual science},\n  volume={57},\n  number={3},\n  pages={949--958},\n  year={2016},\n  publisher={The Association for Research in Vision and Ophthalmology}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Nonmonotonic Elasticity of the Crude Oil–Brine Interface in Relation to Improved Oil Recovery.\n \n \n \n\n\n \n Chávez-Miyauchi, T. E; Firoozabadi, A.; and Fuller, G. G\n\n\n \n\n\n\n Langmuir, 32(9): 2192–2198. 2016.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chavez2016nonmonotonic,\n  title={Nonmonotonic Elasticity of the Crude Oil--Brine Interface in Relation to Improved Oil Recovery},\n  author={Chávez-Miyauchi, Tomás E and Firoozabadi, Abbas and Fuller, Gerald G},\n  journal={Langmuir},\n  volume={32},\n  number={9},\n  pages={2192--2198},\n  year={2016},\n  publisher={ACS Publications}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Interfacial dilatational deformation accelerates particle formation in monoclonal antibody solutions.\n \n \n \n\n\n \n Lin, G. L; Pathak, J. A; Kim, D. H.; Carlson, M.; Riguero, V.; Kim, Y. J.; Buff, J. S; and Fuller, G. G\n\n\n \n\n\n\n Soft matter, 12(14): 3293–3302. 2016.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lin2016interfacial,\n  title={Interfacial dilatational deformation accelerates particle formation in monoclonal antibody solutions},\n  author={Lin, Gigi L and Pathak, Jai A and Kim, Dong Hyun and Carlson, Marcia and Riguero, Valeria and Kim, Yoen Joo and Buff, Jean S and Fuller, Gerald G},\n  journal={Soft matter},\n  volume={12},\n  number={14},\n  pages={3293--3302},\n  year={2016},\n  publisher={Royal Society of Chemistry}\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2015\n \n \n (8)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Nanoscale patterning of extracellular matrix alters endothelial function under shear stress.\n \n \n \n\n\n \n Nakayama, K. H; Surya, V. N; Gole, M.; Walker, T. W; Yang, W.; Lai, E. S; Ostrowski, M. A; Fuller, G. G; Dunn, A. R; and Huang, N. F\n\n\n \n\n\n\n Nano letters, 16(1): 410–419. 2015.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{nakayama2015nanoscale,\n  title={Nanoscale patterning of extracellular matrix alters endothelial function under shear stress},\n  author={Nakayama, Karina H and Surya, Vinay N and Gole, Monica and Walker, Travis W and Yang, Weiguang and Lai, Edwina S and Ostrowski, Maggie A and Fuller, Gerald G and Dunn, Alexander R and Huang, Ngan F},\n  journal={Nano letters},\n  volume={16},\n  number={1},\n  pages={410--419},\n  year={2015},\n  publisher={ACS Publications}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Integrated microfluidic platform for instantaneous flow and localized temperature control.\n \n \n \n \n\n\n \n Fang, C.; Lee, D.; Stober, B.; Fuller, G.; and Shen, A.\n\n\n \n\n\n\n RSC Advances, 5(104): 85620-85629. 2015.\n cited By 0\n\n\n\n
\n\n\n\n \n \n \"IntegratedPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{Fang201585620,\n\tAuthor = {Fang, C. and Lee, D. and Stober, B. and Fuller, G.G. and Shen, A.Q.},\n\tDate-Added = {2015-11-13 18:21:03 +0000},\n\tDate-Modified = {2015-11-13 18:21:03 +0000},\n\tDocument_Type = {Article},\n\tDoi = {10.1039/c5ra19944a},\n\tJournal = {RSC Advances},\n\tNote = {cited By 0},\n\tNumber = {104},\n\tPages = {85620-85629},\n\tSource = {Scopus},\n\tTitle = {Integrated microfluidic platform for instantaneous flow and localized temperature control},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84944909265&partnerID=40&md5=0ab395177aa83dcc61f9147c46dc4534},\n\tVolume = {5},\n\tYear = {2015},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84944909265&partnerID=40&md5=0ab395177aa83dcc61f9147c46dc4534},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1039/c5ra19944a}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Lung surfactants and different contributions to thin film stability.\n \n \n \n \n\n\n \n Hermans, E.; Saad Bhamla, M.; Kao, P.; Fuller, G. G.; and Vermant, J.\n\n\n \n\n\n\n Soft Matter, 11: 8048-8057. 2015.\n \n\n\n\n
\n\n\n\n \n \n \"LungPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{C5SM01603G,\n\tAbstract = {The surfactant lining the walls of the alveoli in the lungs increases pulmonary compliance and prevents collapse of the lung at the end of expiration. In premature born infants{,} surfactant deficiency causes problems{,} and lung surfactant replacements are instilled to facilitate breathing. These pulmonary surfactants{,} which form complex structured fluid-fluid interfaces{,} need to spread with great efficiency and once in the alveolus they have to form a thin stable film. In the present work{,} we investigate the mechanisms affecting the stability of surfactant-laden thin films during spreading{,} using drainage flows from a hemispherical dome. Three commercial lung surfactant replacements Survanta{,} Curosurf and Infasurf{,} along with the phospholipid dipalmitoylphosphatidylcholine (DPPC){,} are used. The surface of the dome can be covered with human alveolar epithelial cells and experiments are conducted at the physiological temperature. Drainage is slowed down due to the presence of all the different lung surfactant replacements and therefore the thin films show enhanced stability. However{,} a scaling analysis combined with visualization experiments demonstrates that different mechanisms are involved. For Curosurf and Infasurf{,} Marangoni stresses are essential to impart stability and interfacial shear rheology does not play a role{,} in agreement with what is observed for simple surfactants. Survanta{,} which was historically the first natural surfactant used{,} is rheologically active. For DPPC the dilatational properties play a role. Understanding these different modes of stabilization for natural surfactants can benefit the design of effective synthetic surfactant replacements for treating infant and adult respiratory disorders.},\n\tAuthor = {Hermans, Eline and Saad Bhamla, M. and Kao, Peter and Fuller, Gerald G. and Vermant, Jan},\n\tDate-Added = {2015-11-13 18:20:16 +0000},\n\tDate-Modified = {2015-11-13 18:20:16 +0000},\n\tDoi = {10.1039/C5SM01603G},\n\tIssue = {41},\n\tJournal = {Soft Matter},\n\tPages = {8048-8057},\n\tPublisher = {The Royal Society of Chemistry},\n\tTitle = {Lung surfactants and different contributions to thin film stability},\n\tUrl = {http://dx.doi.org/10.1039/C5SM01603G},\n\tVolume = {11},\n\tYear = {2015},\n\tBdsk-Url-1 = {http://dx.doi.org/10.1039/C5SM01603G},\n\tBdsk-File-1 = {YnBsaXN0MDDUAQIDBAUGJCVYJHZlcnNpb25YJG9iamVjdHNZJGFyY2hpdmVyVCR0b3ASAAGGoKgHCBMUFRYaIVUkbnVsbNMJCgsMDxJXTlMua2V5c1pOUy5vYmplY3RzViRjbGFzc6INDoACgAOiEBGABIAFgAdccmVsYXRpdmVQYXRoWWFsaWFzRGF0YV8QJi4uLy4uLy4uLy4uLy4uLy4uL0Rvd25sb2Fkcy9zY29wdXMuYmli0hcLGBlXTlMuZGF0YU8RAWwAAAAAAWwAAgAAC2NoeWVhaGhoaGhoAAAAAAAAAAAAAAAAAAAAAM+e1RlIKwAAAAYsUgpzY29wdXMuYmliAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAAj0XH0mtk0QAAAAAAAAAAAAYAAgAACSAAAAAAAAAAAAAAAAAAAAAJRG93bmxvYWRzAAAQAAgAAM+fN4kAAAARAAgAANJr1VEAAAABAAwABixSAAYsTwACZvsAAgAwY2h5ZWFoaGhoaGg6VXNlcnM6AGR3YWxsczoARG93bmxvYWRzOgBzY29wdXMuYmliAA4AFgAKAHMAYwBvAHAAdQBzAC4AYgBpAGIADwAYAAsAYwBoAHkAZQBhAGgAaABoAGgAaABoABIAIVVzZXJzL2R3YWxscy9Eb3dubG9hZHMvc2NvcHVzLmJpYgAAEwABLwAAFQACAA3//wAAgAbSGxwdHlokY2xhc3NuYW1lWCRjbGFzc2VzXU5TTXV0YWJsZURhdGGjHR8gVk5TRGF0YVhOU09iamVjdNIbHCIjXE5TRGljdGlvbmFyeaIiIF8QD05TS2V5ZWRBcmNoaXZlctEmJ1Ryb290gAEACAARABoAIwAtADIANwBAAEYATQBVAGAAZwBqAGwAbgBxAHMAdQB3AIQAjgC3ALwAxAI0AjYCOwJGAk8CXQJhAmgCcQJ2AoMChgKYApsCoAAAAAAAAAIBAAAAAAAAACgAAAAAAAAAAAAAAAAAAAKi}}\n\n
\n
\n\n\n
\n The surfactant lining the walls of the alveoli in the lungs increases pulmonary compliance and prevents collapse of the lung at the end of expiration. In premature born infants, surfactant deficiency causes problems, and lung surfactant replacements are instilled to facilitate breathing. These pulmonary surfactants, which form complex structured fluid-fluid interfaces, need to spread with great efficiency and once in the alveolus they have to form a thin stable film. In the present work, we investigate the mechanisms affecting the stability of surfactant-laden thin films during spreading, using drainage flows from a hemispherical dome. Three commercial lung surfactant replacements Survanta, Curosurf and Infasurf, along with the phospholipid dipalmitoylphosphatidylcholine (DPPC), are used. The surface of the dome can be covered with human alveolar epithelial cells and experiments are conducted at the physiological temperature. Drainage is slowed down due to the presence of all the different lung surfactant replacements and therefore the thin films show enhanced stability. However, a scaling analysis combined with visualization experiments demonstrates that different mechanisms are involved. For Curosurf and Infasurf, Marangoni stresses are essential to impart stability and interfacial shear rheology does not play a role, in agreement with what is observed for simple surfactants. Survanta, which was historically the first natural surfactant used, is rheologically active. For DPPC the dilatational properties play a role. Understanding these different modes of stabilization for natural surfactants can benefit the design of effective synthetic surfactant replacements for treating infant and adult respiratory disorders.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Multiphase flow of miscible liquids: jets and drops.\n \n \n \n \n\n\n \n Walker, T. W; Logia, A. N; and Fuller, G. G\n\n\n \n\n\n\n Experiments in Fluids, 56(5): 1–14. May 2015.\n \n\n\n\n
\n\n\n\n \n \n \"MultiphasePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{walker_multiphase_2015,\n\tAuthor = {Walker, Travis W and Logia, Alison N and Fuller, Gerald G},\n\tDoi = {10.1007/s00348-015-1974-y},\n\tJournal = {Experiments in Fluids},\n\tLanguage = {English},\n\tMonth = may,\n\tNumber = {5},\n\tPages = {1--14},\n\tTitle = {Multiphase flow of miscible liquids: jets and drops},\n\tUrl = {http://link.springer.com/article/10.1007/s00348-015-1974-y/fulltext.html},\n\tVolume = {56},\n\tYear = {2015},\n\tBdsk-Url-1 = {http://link.springer.com/article/10.1007/s00348-015-1974-y/fulltext.html},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1007/s00348-015-1974-y}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Dewetting and deposition of thin films with insoluble surfactants from curved silicone hydrogel substrates.\n \n \n \n \n\n\n \n Bhamla, M S.; Balemans, C.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Colloid and Interface Science,–. March 2015.\n \n\n\n\n
\n\n\n\n \n \n \"DewettingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{bhamla_dewetting_2015,\n\tAbstract = {[Display omitted]},\n\tAuthor = {Bhamla, M Saad and Balemans, Caroline and Fuller, Gerald G},\n\tDoi = {10.1016/j.jcis.2015.01.002},\n\tJournal = {Journal of Colloid and Interface Science},\n\tLanguage = {English},\n\tMonth = mar,\n\tPages = {--},\n\tPmid = {25628055},\n\tTitle = {Dewetting and deposition of thin films with insoluble surfactants from curved silicone hydrogel substrates},\n\tUrl = {http://linkinghub.elsevier.com/retrieve/pii/S0021979715000181},\n\tYear = {2015},\n\tBdsk-Url-1 = {http://linkinghub.elsevier.com/retrieve/pii/S0021979715000181},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1016/j.jcis.2015.01.002}}\n\n
\n
\n\n\n
\n [Display omitted]\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Quantification of stromal vascular cell mechanics with a linear cell monolayer rheometer.\n \n \n \n \n\n\n \n Elkins, C. M; Shen, W.; Khor, V. K; Kraemer, F. B; and Fuller, G. G\n\n\n \n\n\n\n Journal of Rheology (1978-present), 59(1): 33–50. January 2015.\n \n\n\n\n
\n\n\n\n \n \n \"QuantificationPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{elkins_quantification_2015,\n\tAbstract = {Over the past few decades researchers have developed a variety of methods for measuring the mechanical properties of whole cells, including traction force microscopy, atomic force microscopy (AFM), and single-cell tensile testing. Though each of these techniques provides insight into cell mechanics, most also involve some nonideal conditions for acquiring live cell data, such as probing only one portion of a cell at a time, or placing the cell in a nonrepresentative geometry during testing. In the present work, we describe the development of a linear cell monolayer rheometer (LCMR) and its application to measure the mechanics of a live, confluent monolayer of stromal vascular cells. In the LCMR, a monolayer of cells is contacted on both top and bottom by two collagen-coated plates and allowed to adhere. The top plate then shears the monolayer by stepping forward to induce a predetermined step strain, while a force transducer attached to the top plate collects stress information. The stress and strain data are then used to determine the maximum relaxation modulus recorded after step-strain, G r 0 , referred to as the zero-time relaxation modulus of the cell monolayer. The present study validates the ability of the LCMR to quantify cell mechanics by measuring the change in G r 0 of a confluent cell monolayer upon the selective inhibition of three major cytoskeletal components (actin microfilaments, vimentin intermediate filaments, and microtubules). The LCMR results indicate that both actin- and vimentin-deficient cells had ∼50\\% lower G r 0 values than wild-type, whereas tubulin deficiency resulted in ∼100\\% higher G r 0 values. These findings constitute the first use of a cell monolayer rheometer to quantitatively distinguish the roles of different cytoskeletal elements in maintaining cell stiffness and structure. Significantly, they are consistent with results obtained using single-cell mechanical testing methods, suggesting that the rheology-based LCMR technique may be a useful tool for rapid analysis of cell mechanics by shearing an entire cell monolayer.},\n\tAuthor = {Elkins, Claire M and Shen, Wen-Jun and Khor, Victor K and Kraemer, Fredric B and Fuller, Gerald G},\n\tDoi = {10.1122/1.4902437},\n\tJournal = {Journal of Rheology (1978-present)},\n\tLanguage = {English},\n\tMonth = jan,\n\tNumber = {1},\n\tPages = {33--50},\n\tTitle = {Quantification of stromal vascular cell mechanics with a linear cell monolayer rheometer},\n\tUrl = {http://scitation.aip.org/content/sor/journal/jor2/59/1/10.1122/1.4902437},\n\tVolume = {59},\n\tYear = {2015},\n\tBdsk-Url-1 = {http://scitation.aip.org/content/sor/journal/jor2/59/1/10.1122/1.4902437},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1122/1.4902437}}\n\n
\n
\n\n\n
\n Over the past few decades researchers have developed a variety of methods for measuring the mechanical properties of whole cells, including traction force microscopy, atomic force microscopy (AFM), and single-cell tensile testing. Though each of these techniques provides insight into cell mechanics, most also involve some nonideal conditions for acquiring live cell data, such as probing only one portion of a cell at a time, or placing the cell in a nonrepresentative geometry during testing. In the present work, we describe the development of a linear cell monolayer rheometer (LCMR) and its application to measure the mechanics of a live, confluent monolayer of stromal vascular cells. In the LCMR, a monolayer of cells is contacted on both top and bottom by two collagen-coated plates and allowed to adhere. The top plate then shears the monolayer by stepping forward to induce a predetermined step strain, while a force transducer attached to the top plate collects stress information. The stress and strain data are then used to determine the maximum relaxation modulus recorded after step-strain, G r 0 , referred to as the zero-time relaxation modulus of the cell monolayer. The present study validates the ability of the LCMR to quantify cell mechanics by measuring the change in G r 0 of a confluent cell monolayer upon the selective inhibition of three major cytoskeletal components (actin microfilaments, vimentin intermediate filaments, and microtubules). The LCMR results indicate that both actin- and vimentin-deficient cells had ∼50% lower G r 0 values than wild-type, whereas tubulin deficiency resulted in ∼100% higher G r 0 values. These findings constitute the first use of a cell monolayer rheometer to quantitatively distinguish the roles of different cytoskeletal elements in maintaining cell stiffness and structure. Significantly, they are consistent with results obtained using single-cell mechanical testing methods, suggesting that the rheology-based LCMR technique may be a useful tool for rapid analysis of cell mechanics by shearing an entire cell monolayer.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Influence of lipid coatings on surface wettability characteristics of silicone hydrogels.\n \n \n \n \n\n\n \n Bhamla, M S.; Nash, W. L; Elliott, S.; and Fuller, G. G\n\n\n \n\n\n\n Langmuir, 31(13): 3820–3828. April 2015.\n \n\n\n\n
\n\n\n\n \n \n \"InfluencePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{bhamla_influence_2015,\n\tAbstract = {Insoluble lipids serve vital functions in our bodies and interact with biomedical devices, e.g., the tear film on a contact lens. Over a period of time, these naturally occurring lipids form interfacial coatings that modify the wettability characteristics of these foreign synthetic surfaces. In this study, we examine the deposition and consequences of tear film lipids on silicone hydrogel (SiHy) contact lenses. We use bovine meibum, which is a complex mixture of waxy esters, cholesterol esters, and lipids that is secreted from the meibomian glands located on the upper and lower eyelids of mammals. For comparison, we study two commercially available model materials: dipalmitoylphosphatidylcholine (DPPC) and cholesterol. Upon deposition, we find that DPPC and meibum remain closer to the SiHy surface than cholesterol, which diffuses further into the porous SiHy matrix. In addition, we also monitor the fate of unstable thin liquid films that consequently rupture and dewet on these lipid-decorated surfaces. This dewetting provides valuable qualitative and quantitative information about the wetting characteristics of these SiHy substrates. We observe that decorating the SiHy surface with simple model lipids such as DPPC and cholesterol increases the hydrophilicity, which consequently inhibits dewetting, whereas meibum behaves conversely.},\n\tAuthor = {Bhamla, M Saad and Nash, Walter L and Elliott, Stacey and Fuller, Gerald G},\n\tDoi = {10.1021/la503437a},\n\tJournal = {Langmuir},\n\tLanguage = {English},\n\tMonth = apr,\n\tNumber = {13},\n\tPages = {3820--3828},\n\tPmid = {25280206},\n\tTitle = {Influence of lipid coatings on surface wettability characteristics of silicone hydrogels.},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/la503437a},\n\tVolume = {31},\n\tYear = {2015},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/la503437a},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/la503437a}}\n\n
\n
\n\n\n
\n Insoluble lipids serve vital functions in our bodies and interact with biomedical devices, e.g., the tear film on a contact lens. Over a period of time, these naturally occurring lipids form interfacial coatings that modify the wettability characteristics of these foreign synthetic surfaces. In this study, we examine the deposition and consequences of tear film lipids on silicone hydrogel (SiHy) contact lenses. We use bovine meibum, which is a complex mixture of waxy esters, cholesterol esters, and lipids that is secreted from the meibomian glands located on the upper and lower eyelids of mammals. For comparison, we study two commercially available model materials: dipalmitoylphosphatidylcholine (DPPC) and cholesterol. Upon deposition, we find that DPPC and meibum remain closer to the SiHy surface than cholesterol, which diffuses further into the porous SiHy matrix. In addition, we also monitor the fate of unstable thin liquid films that consequently rupture and dewet on these lipid-decorated surfaces. This dewetting provides valuable qualitative and quantitative information about the wetting characteristics of these SiHy substrates. We observe that decorating the SiHy surface with simple model lipids such as DPPC and cholesterol increases the hydrophilicity, which consequently inhibits dewetting, whereas meibum behaves conversely.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Interplay of Hydrogen Bonding and Hydrophobic Interactions to Control the Mechanical Properties of Polymer Multilayers at the Oil–Water Interface.\n \n \n \n \n\n\n \n Le Tirilly, S.; Tregouët, C.; Bône, S.; Geffroy, C.; Fuller, G.; Pantoustier, N.; Perrin, P.; and Monteux, C.\n\n\n \n\n\n\n ACS Macro Letters, 4(1): 25–29. January 2015.\n \n\n\n\n
\n\n\n\n \n \n \"InterplayPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{le_tirilly_interplay_2015,\n\tAbstract = {We probe the mechanical shear and compression properties of hydrogen-bonded polymer multilayers directly assembled at the oil?water interface using interfacial rheology techniques. We show that the polymer multilayers behave mechanically like a transient network, with elastic moduli that can be varied over 2 orders of magnitude by controlling the type and strength of physical interactions involved in the multilayers, which are controlled by the pH and the hydrophobicity of the polymer. Indeed, the interplay of hydrogen and hydrophobic interactions enables one to obtain a tighter and stronger network at the interface. Moreover, we show how a simple LBL process applied directly on emulsion droplets leads to encapsulation of a model oil, dodecane, as well as perfume molecules.},\n\tAuthor = {Le Tirilly, Sandrine and Tregou{\\"e}t, Corentin and B{\\^o}ne, St{\\'e}phane and Geffroy, C{\\'e}dric and Fuller, Gerald and Pantoustier, Nad{\\`e}ge and Perrin, Patrick and Monteux, C{\\'e}cile},\n\tDoi = {10.1021/mz5005772},\n\tFile = {ACS Full Text PDF w/ Links:/Users/dwalls/Library/Application Support/Zotero/Profiles/z5rwszv8.default/zotero/storage/UNP5GSTW/Le Tirilly et al. - 2015 - Interplay of Hydrogen Bonding and Hydrophobic Inte.pdf:application/pdf;ACS Full Text Snapshot:/Users/dwalls/Library/Application Support/Zotero/Profiles/z5rwszv8.default/zotero/storage/4KDZVB4H/mz5005772.html:text/html},\n\tJournal = {ACS Macro Letters},\n\tMonth = jan,\n\tNumber = {1},\n\tPages = {25--29},\n\tTitle = {Interplay of {Hydrogen} {Bonding} and {Hydrophobic} {Interactions} to {Control} the {Mechanical} {Properties} of {Polymer} {Multilayers} at the {Oil}--{Water} {Interface}},\n\tUrl = {http://dx.doi.org/10.1021/mz5005772},\n\tUrldate = {2015-08-19},\n\tVolume = {4},\n\tYear = {2015},\n\tBdsk-Url-1 = {http://dx.doi.org/10.1021/mz5005772}}\n
\n
\n\n\n
\n We probe the mechanical shear and compression properties of hydrogen-bonded polymer multilayers directly assembled at the oil?water interface using interfacial rheology techniques. We show that the polymer multilayers behave mechanically like a transient network, with elastic moduli that can be varied over 2 orders of magnitude by controlling the type and strength of physical interactions involved in the multilayers, which are controlled by the pH and the hydrophobicity of the polymer. Indeed, the interplay of hydrogen and hydrophobic interactions enables one to obtain a tighter and stronger network at the interface. Moreover, we show how a simple LBL process applied directly on emulsion droplets leads to encapsulation of a model oil, dodecane, as well as perfume molecules.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2014\n \n \n (8)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Influence of interfacial rheology on drainage from curved surfaces.\n \n \n \n \n\n\n \n Bhamla, M S.; Giacomin, C. E; Balemans, C.; and Fuller, G. G\n\n\n \n\n\n\n Soft Matter, 10(36): 6917–6925. September 2014.\n \n\n\n\n
\n\n\n\n \n \n \"InfluencePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{bhamla_influence_2014,\n\tAbstract = {Thin lubrication flows accompanying drainage from curved surfaces surround us (e.g., the drainage of the tear film on our eyes). These draining aqueous layers are normally covered with surface-active molecules that render the free surface viscoelastic. The non-Newtonian character of these surfaces fundamentally alters the dynamics of drainage. We show that increased film stability during drainage can occur as a consequence of enhanced surface rheology. Increasing the surfactant layer viscosity decreases the rate of drainage; however, this retarding influence is most pronounced when the insoluble surfactant layer has significant elasticity. We also present a simple theoretical model that offers qualitative support to our experimental findings.},\n\tAuthor = {Bhamla, M Saad and Giacomin, Caroline E and Balemans, Caroline and Fuller, Gerald G},\n\tDoi = {10.1039/c3sm52934g},\n\tJournal = {Soft Matter},\n\tLanguage = {English},\n\tMonth = sep,\n\tNumber = {36},\n\tPages = {6917--6925},\n\tPmid = {25140576},\n\tTitle = {Influence of interfacial rheology on drainage from curved surfaces.},\n\tUrl = {http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=2014SMat...10.6917B&link_type=EJOURNAL},\n\tVolume = {10},\n\tYear = {2014},\n\tBdsk-Url-1 = {http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=2014SMat...10.6917B&link_type=EJOURNAL},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1039/c3sm52934g}}\n\n
\n
\n\n\n
\n Thin lubrication flows accompanying drainage from curved surfaces surround us (e.g., the drainage of the tear film on our eyes). These draining aqueous layers are normally covered with surface-active molecules that render the free surface viscoelastic. The non-Newtonian character of these surfaces fundamentally alters the dynamics of drainage. We show that increased film stability during drainage can occur as a consequence of enhanced surface rheology. Increasing the surfactant layer viscosity decreases the rate of drainage; however, this retarding influence is most pronounced when the insoluble surfactant layer has significant elasticity. We also present a simple theoretical model that offers qualitative support to our experimental findings.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Instabilities and elastic recoil of the two-fluid circular hydraulic jump.\n \n \n \n \n\n\n \n Hsu, T. T; Walker, T. W; Frank, C. W; and Fuller, G. G\n\n\n \n\n\n\n Experiments in Fluids, 55(1): 1645. 2014.\n \n\n\n\n
\n\n\n\n \n \n \"InstabilitiesPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hsu_instabilities_2014,\n\tAbstract = {The two-fluid circular hydraulic jump, also called "rinsing flow," is a common process where a jet of one liquid impinges upon a layer of a second liquid. We present an experimental analysis of rinsing flows using a high-speed camera and model fluids to decouple the effect of shear-thinning and elasticity. Varying the rheology of the coating fluid produced several types of instabilities at both the air-liquid interface and liquid-liquid interface. Layered "stepped jumps" and "crowning" on the rim of the jumps were both suppressed by fluid elasticity, while Saffman-Taylor fingering patterns showed strong dependence on both shear-thinning and normal stresses. In addition, the hydraulic jump evolution was quantitatively determined using a laser triangulation technique, and "recoil" of the jump front resulting from fluid elasticity was observed. Our work shows that the non-Newtonian two-fluid circular hydraulic jump is very complex, and the instabilities that arise also introduce additional complications when developing theoretical models. textcopyright Springer-Verlag Berlin Heidelberg 2013.},\n\tAuthor = {Hsu, Tienyi T and Walker, Travis W and Frank, Curtis W and Fuller, Gerald G},\n\tDoi = {10.1007/s00348-013-1645-9},\n\tJournal = {Experiments in Fluids},\n\tNumber = {1},\n\tPages = {1645},\n\tTitle = {Instabilities and elastic recoil of the two-fluid circular hydraulic jump},\n\tUrl = {http://link.springer.com/10.1007/s00348-013-1645-9},\n\tVolume = {55},\n\tYear = {2014},\n\tBdsk-Url-1 = {http://link.springer.com/10.1007/s00348-013-1645-9},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1007/s00348-013-1645-9}}\n\n
\n
\n\n\n
\n The two-fluid circular hydraulic jump, also called \"rinsing flow,\" is a common process where a jet of one liquid impinges upon a layer of a second liquid. We present an experimental analysis of rinsing flows using a high-speed camera and model fluids to decouple the effect of shear-thinning and elasticity. Varying the rheology of the coating fluid produced several types of instabilities at both the air-liquid interface and liquid-liquid interface. Layered \"stepped jumps\" and \"crowning\" on the rim of the jumps were both suppressed by fluid elasticity, while Saffman-Taylor fingering patterns showed strong dependence on both shear-thinning and normal stresses. In addition, the hydraulic jump evolution was quantitatively determined using a laser triangulation technique, and \"recoil\" of the jump front resulting from fluid elasticity was observed. Our work shows that the non-Newtonian two-fluid circular hydraulic jump is very complex, and the instabilities that arise also introduce additional complications when developing theoretical models. textcopyright Springer-Verlag Berlin Heidelberg 2013.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Using in-Situ Polymerization of Conductive Polymers to Enhance the Electrical Properties of Solution-Processed Carbon Nanotube Films and Fibers.\n \n \n \n \n\n\n \n Allen, R.; Pan, L.; Fuller, G. G; and Bao, Z.\n\n\n \n\n\n\n ACS Applied Materials and Interfaces, 6(13): 9966–9974. July 2014.\n \n\n\n\n
\n\n\n\n \n \n \"UsingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{allen_using_2014,\n\tAuthor = {Allen, Ranulfo and Pan, Lijia and Fuller, Gerald G and Bao, Zhenan},\n\tDoi = {10.1021/am5019995},\n\tJournal = {ACS Applied Materials and Interfaces},\n\tLanguage = {English},\n\tMonth = jul,\n\tNumber = {13},\n\tPages = {9966--9974},\n\tTitle = {Using in-{Situ} {Polymerization} of {Conductive} {Polymers} to {Enhance} the {Electrical} {Properties} of {Solution}-{Processed} {Carbon} {Nanotube} {Films} and {Fibers}},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/am5019995},\n\tVolume = {6},\n\tYear = {2014},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/am5019995},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/am5019995}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Microvascular Endothelial Cells Migrate Upstream and Align Against the Shear Stress Field Created by Impinging Flow.\n \n \n \n \n\n\n \n Ostrowski, M. A; Huang, N. F; Walker, T. W; Verwijlen, T.; Poplawski, C.; Khoo, A. S; Cooke, J. P; Fuller, G. G; and Dunn, A. R\n\n\n \n\n\n\n Biophysical Journal, 106(2): 366–374. 2014.\n \n\n\n\n
\n\n\n\n \n \n \"MicrovascularPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{ostrowski_microvascular_2014,\n\tAbstract = {At present, little is known about how endothelial cells respond to spatial variations in fluid shear stress such as those that occur locally during embryonic development, at heart valve leaflets, and at sites of aneurysm formation. We built an impinging flow device that exposes endothelial cells to gradients of shear stress. Using this device, we investigated the response of microvascular endothelial cells to shear-stress gradients that ranged from 0 to a peak shear stress of 9-210 dyn/cm2. We observe that at high confluency, these cells migrate against the direction of fluid flow and concentrate in the region of maximum wall shear stress, whereas low-density microvascular endothelial cells that lack cell-cell contacts migrate in the flow direction. In addition, the cells align parallel to the flow at low wall shear stresses but orient perpendicularly to the flow direction above a critical threshold in local wall shear stress. Our observations suggest that endothelial cells are exquisitely sensitive to both magnitude and spatial gradients in wall shear stress. The impinging flow device provides a, to our knowledge, novel means to study endothelial cell migration and polarization in response to gradients in physical forces such as wall shear stress. textcopyright 2014 Biophysical Society.},\n\tAuthor = {Ostrowski, Maggie A and Huang, Ngan F and Walker, Travis W and Verwijlen, Tom and Poplawski, Charlotte and Khoo, Amanda S and Cooke, John P and Fuller, Gerald G and Dunn, Alexander R},\n\tDoi = {10.1016/j.bpj.2013.11.4502},\n\tJournal = {Biophysical Journal},\n\tNumber = {2},\n\tPages = {366--374},\n\tTitle = {Microvascular {Endothelial} {Cells} {Migrate} {Upstream} and {Align} {Against} the {Shear} {Stress} {Field} {Created} by {Impinging} {Flow}},\n\tUrl = {http://linkinghub.elsevier.com/retrieve/pii/S0006349513058074},\n\tVolume = {106},\n\tYear = {2014},\n\tBdsk-Url-1 = {http://linkinghub.elsevier.com/retrieve/pii/S0006349513058074},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1016/j.bpj.2013.11.4502}}\n\n
\n
\n\n\n
\n At present, little is known about how endothelial cells respond to spatial variations in fluid shear stress such as those that occur locally during embryonic development, at heart valve leaflets, and at sites of aneurysm formation. We built an impinging flow device that exposes endothelial cells to gradients of shear stress. Using this device, we investigated the response of microvascular endothelial cells to shear-stress gradients that ranged from 0 to a peak shear stress of 9-210 dyn/cm2. We observe that at high confluency, these cells migrate against the direction of fluid flow and concentrate in the region of maximum wall shear stress, whereas low-density microvascular endothelial cells that lack cell-cell contacts migrate in the flow direction. In addition, the cells align parallel to the flow at low wall shear stresses but orient perpendicularly to the flow direction above a critical threshold in local wall shear stress. Our observations suggest that endothelial cells are exquisitely sensitive to both magnitude and spatial gradients in wall shear stress. The impinging flow device provides a, to our knowledge, novel means to study endothelial cell migration and polarization in response to gradients in physical forces such as wall shear stress. textcopyright 2014 Biophysical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Scaling analysis and mathematical theory of the interfacial stress rheometer.\n \n \n \n \n\n\n \n Fitzgibbon, S; Shaqfeh, E S G; Fuller, G.; and Walker, T W\n\n\n \n\n\n\n Journal of Rheology, 58(4): 999–1038. 2014.\n \n\n\n\n
\n\n\n\n \n \n \"ScalingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fitzgibbon_scaling_2014,\n\tAbstract = {Not Available},\n\tAuthor = {Fitzgibbon, S and Shaqfeh, E S G and Fuller, G.G. and Walker, T W},\n\tDoi = {10.1122/1.4876955},\n\tJournal = {Journal of Rheology},\n\tLanguage = {English},\n\tNumber = {4},\n\tPages = {999--1038},\n\tTitle = {Scaling analysis and mathematical theory of the interfacial stress rheometer},\n\tUrl = {http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=2014JRheo..58..999F&link_type=EJOURNAL},\n\tVolume = {58},\n\tYear = {2014},\n\tBdsk-Url-1 = {http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=2014JRheo..58..999F&link_type=EJOURNAL},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1122/1.4876955}}\n\n
\n
\n\n\n
\n Not Available\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Corneal Cell Adhesion to Contact Lens Hydrogel Materials Enhanced via Tear Film Protein Deposition.\n \n \n \n \n\n\n \n Elkins, C. M; Qi, Q. M; and Fuller, G. G\n\n\n \n\n\n\n PLoS ONE, 9(8): e105512. 2014.\n \n\n\n\n
\n\n\n\n \n \n \"CornealPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{elkins_corneal_2014,\n\tAbstract = {Tear film protein deposition on contact lens hydrogels has been well characterized from the perspective of bacterial adhesion and viability. However, the effect of protein deposition on lens interactions with the corneal epithelium remains largely unexplored. The current study employs a live cell rheometer to quantify human corneal epithelial cell adhesion to soft contact lenses fouled with the tear film protein lysozyme. PureVision balafilcon A and AirOptix lotrafilcon B lenses were soaked for five days in either phosphate buffered saline (PBS), borate buffered saline (BBS), or Sensitive Eyes Plus Saline Solution (Sensitive Eyes), either pure or in the presence of lysozyme. Treated contact lenses were then contacted to a live monolayer of corneal epithelial cells for two hours, after which the contact lens was sheared laterally. The apparent cell monolayer relaxation modulus was then used to quantify the extent of cell adhesion to the contact lens surface. For both lens types, lysozyme increased corneal cell adhesion to the contact lens, with the apparent cell monolayer relaxation modulus increasing up to an order of magnitude in the presence of protein. The magnitude of this increase depended on the identity of the soaking solution: lenses soaked in borate-buffered solutions (BBS, Sensitive Eyes) exhibited a much greater increase in cell attachment upon protein addition than those soaked in PBS. Significantly, all measurements were conducted while subjecting the cells to moderate surface pressures and shear rates, similar to those experienced by corneal cells in vivo.},\n\tAuthor = {Elkins, Claire M and Qi, Qin M and Fuller, Gerald G},\n\tDoi = {10.1371/journal.pone.0105512},\n\tJournal = {PLoS ONE},\n\tLanguage = {English},\n\tNumber = {8},\n\tPages = {e105512},\n\tPmid = {25144576},\n\tTitle = {Corneal {Cell} {Adhesion} to {Contact} {Lens} {Hydrogel} {Materials} {Enhanced} via {Tear} {Film} {Protein} {Deposition}.},\n\tUrl = {http://eutils.ncbi.nlm.nih.gov/entrez/eutils/elink.fcgi?dbfrom=pubmed&id=25144576&retmode=ref&cmd=prlinks},\n\tVolume = {9},\n\tYear = {2014},\n\tBdsk-Url-1 = {http://eutils.ncbi.nlm.nih.gov/entrez/eutils/elink.fcgi?dbfrom=pubmed&id=25144576&retmode=ref&cmd=prlinks},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1371/journal.pone.0105512}}\n\n
\n
\n\n\n
\n Tear film protein deposition on contact lens hydrogels has been well characterized from the perspective of bacterial adhesion and viability. However, the effect of protein deposition on lens interactions with the corneal epithelium remains largely unexplored. The current study employs a live cell rheometer to quantify human corneal epithelial cell adhesion to soft contact lenses fouled with the tear film protein lysozyme. PureVision balafilcon A and AirOptix lotrafilcon B lenses were soaked for five days in either phosphate buffered saline (PBS), borate buffered saline (BBS), or Sensitive Eyes Plus Saline Solution (Sensitive Eyes), either pure or in the presence of lysozyme. Treated contact lenses were then contacted to a live monolayer of corneal epithelial cells for two hours, after which the contact lens was sheared laterally. The apparent cell monolayer relaxation modulus was then used to quantify the extent of cell adhesion to the contact lens surface. For both lens types, lysozyme increased corneal cell adhesion to the contact lens, with the apparent cell monolayer relaxation modulus increasing up to an order of magnitude in the presence of protein. The magnitude of this increase depended on the identity of the soaking solution: lenses soaked in borate-buffered solutions (BBS, Sensitive Eyes) exhibited a much greater increase in cell attachment upon protein addition than those soaked in PBS. Significantly, all measurements were conducted while subjecting the cells to moderate surface pressures and shear rates, similar to those experienced by corneal cells in vivo.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Molecular Determinants of Mechanical Properties of V. cholerae Biofilms at the Air-Liquid Interface.\n \n \n \n \n\n\n \n Hollenbeck, E. C; Fong, J. C.; Lim, J. Y.; Yildiz, F. H; Fuller, G. G; and Cegelski, L.\n\n\n \n\n\n\n Biophysical Journal, 107(10): 2245–2252. November 2014.\n \n\n\n\n
\n\n\n\n \n \n \"MolecularPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hollenbeck_molecular_2014,\n\tAbstract = {Biofilm formation increases both the survival and infectivity of Vibrio cholerae, the causative agent of cholera. V. cholerae is capable of forming biofilms on solid surfaces and at the air-liquid interface, termed pellicles. Known components of the extracellular matrix include the matrix proteins Bap1, RbmA, and RbmC, an exopolysaccharide termed Vibrio polysaccharide, and DNA. In this work, we examined a rugose strain of V. cholerae and its mutants unable to produce matrix proteins by interfacial rheology to compare the evolution of pellicle elasticity in real time to understand the molecular basis of matrix protein contributions to pellicle integrity and elasticity. Together with electron micrographs, visual inspection, and contact angle measurements of the pellicles, we defined distinct contributions of the matrix proteins to pellicle morphology, microscale architecture, and mechanical properties. Furthermore, we discovered that Bap1 is uniquely required for the maintenance of the mechanical strength of the pellicle over time and contributes to the hydrophobicity of the pellicle. Thus, Bap1 presents an important matrix component to target in the prevention and dispersal of V. cholerae biofilms.},\n\tAuthor = {Hollenbeck, Emily C and Fong, Jiunn CN and Lim, Ji Youn and Yildiz, Fitnat H and Fuller, Gerald G and Cegelski, Lynette},\n\tDoi = {10.1016/j.bpj.2014.10.015},\n\tJournal = {Biophysical Journal},\n\tLanguage = {English},\n\tMonth = nov,\n\tNumber = {10},\n\tPages = {2245--2252},\n\tTitle = {Molecular {Determinants} of {Mechanical} {Properties} of {V}. cholerae {Biofilms} at the {Air}-{Liquid} {Interface}},\n\tUrl = {http://linkinghub.elsevier.com/retrieve/pii/S0006349514010662},\n\tVolume = {107},\n\tYear = {2014},\n\tBdsk-Url-1 = {http://linkinghub.elsevier.com/retrieve/pii/S0006349514010662},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1016/j.bpj.2014.10.015}}\n\n
\n
\n\n\n
\n Biofilm formation increases both the survival and infectivity of Vibrio cholerae, the causative agent of cholera. V. cholerae is capable of forming biofilms on solid surfaces and at the air-liquid interface, termed pellicles. Known components of the extracellular matrix include the matrix proteins Bap1, RbmA, and RbmC, an exopolysaccharide termed Vibrio polysaccharide, and DNA. In this work, we examined a rugose strain of V. cholerae and its mutants unable to produce matrix proteins by interfacial rheology to compare the evolution of pellicle elasticity in real time to understand the molecular basis of matrix protein contributions to pellicle integrity and elasticity. Together with electron micrographs, visual inspection, and contact angle measurements of the pellicles, we defined distinct contributions of the matrix proteins to pellicle morphology, microscale architecture, and mechanical properties. Furthermore, we discovered that Bap1 is uniquely required for the maintenance of the mechanical strength of the pellicle over time and contributes to the hydrophobicity of the pellicle. Thus, Bap1 presents an important matrix component to target in the prevention and dispersal of V. cholerae biofilms.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Enhanced particle removal using viscoelastic fluids.\n \n \n \n\n\n \n Walker, T W; Hsu, T T; Fitzgibbon, S; Frank, C.; Mui, D S L; Zhu, J; Mendiratta, A; and Fuller, G.\n\n\n \n\n\n\n Journal of Rheology, 58(1): 63–88. 2014.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{walker_enhanced_2014,\n\tAuthor = {Walker, T W and Hsu, T T and Fitzgibbon, S and Frank, C.W. and Mui, D S L and Zhu, J and Mendiratta, A and Fuller, G.G.},\n\tJournal = {Journal of Rheology},\n\tNumber = {1},\n\tPages = {63--88},\n\tTitle = {Enhanced particle removal using viscoelastic fluids},\n\tVolume = {58},\n\tYear = {2014}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2013\n \n \n (10)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Thermoresponsiveness of PDMAEMA. Electrostatic and Stereochemical Effects.\n \n \n \n \n\n\n \n Niskanen, J.; Wu, C.; Ostrowski, M.; Fuller, G. G; Hietala, S.; and Tenhu, H.\n\n\n \n\n\n\n Macromolecules, 46(6): 2331–2340. 2013.\n \n\n\n\n
\n\n\n\n \n \n \"ThermoresponsivenessPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{niskanen_thermoresponsiveness_2013,\n\tAbstract = {Isotactic triads are introduced into poly(dimethylaminoethyl methacrylate) (PDMAEMA) when a Lewis acid yttrium(III)trifluoromethanesulfonate, Y(OTf)3, is present during the ATRP polymerization. The changes in the tacticities of the polymers are modest. However, the tacticity affects the phase separation process but in a different way in two studied cases, at pH 8 and 9. The pH, and thus the charge of the polymer, affects the balance between electrostatic and stereochemical effects. Upon the chain collapse, the zeta potential of the polymer decreases discontinuously at pH 9, whereas at pH 8 the potential keeps almost constant. However, even in the latter case the influence of the isotactic segments on the thermal transition may be observed. Increasing isotacticity is suggested to decrease the flexibility of the polymer chain. It also causes the polymers to adsorb in a more organized manner to the air/water interface than the atactic ones do. The change in the thermoresponsive behavior due to the changing t...},\n\tAuthor = {Niskanen, Jukka and Wu, Cynthia and Ostrowski, Maggie and Fuller, Gerald G and Hietala, Sami and Tenhu, Heikki},\n\tDoi = {10.1021/ma302648w},\n\tJournal = {Macromolecules},\n\tNumber = {6},\n\tPages = {2331--2340},\n\tTitle = {Thermoresponsiveness of {PDMAEMA}. {Electrostatic} and {Stereochemical} {Effects}},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/ma302648w},\n\tVolume = {46},\n\tYear = {2013},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/ma302648w},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/ma302648w}}\n\n
\n
\n\n\n
\n Isotactic triads are introduced into poly(dimethylaminoethyl methacrylate) (PDMAEMA) when a Lewis acid yttrium(III)trifluoromethanesulfonate, Y(OTf)3, is present during the ATRP polymerization. The changes in the tacticities of the polymers are modest. However, the tacticity affects the phase separation process but in a different way in two studied cases, at pH 8 and 9. The pH, and thus the charge of the polymer, affects the balance between electrostatic and stereochemical effects. Upon the chain collapse, the zeta potential of the polymer decreases discontinuously at pH 9, whereas at pH 8 the potential keeps almost constant. However, even in the latter case the influence of the isotactic segments on the thermal transition may be observed. Increasing isotacticity is suggested to decrease the flexibility of the polymer chain. It also causes the polymers to adsorb in a more organized manner to the air/water interface than the atactic ones do. The change in the thermoresponsive behavior due to the changing t...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Aligned SWNT films from low-yield stress gels and their transparent electrode performance.\n \n \n \n \n\n\n \n Allen, R; Fuller, G.; and Bao, Z\n\n\n \n\n\n\n ACS Applied Materials and Interfaces, 5(15): 7244–7252. 2013.\n \n\n\n\n
\n\n\n\n \n \n \"AlignedPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{allen_aligned_2013,\n\tAbstract = {Carbon nanotube films are promising for transparent electrodes for solar cells and displays. Large-area alignment of the nanotubes in these films is needed to minimize the sheet resistance. We present a novel coating method to coat high-density, aligned nanotubes over large areas. Carbon nanotube gel dispersions used in this study have aligned domains and a low yield stress. A simple shearing force allows these domains to uniformly align. We use this to correlate the transparent electrode performance of single-walled carbon nanotube films with the level of partial alignment. We have found that the transparent electrode performance improves with increasing levels of alignment and in a manner slightly better than what has been previously predicted. textcopyright 2013 American Chemical Society.},\n\tAuthor = {Allen, R and Fuller, G.G. and Bao, Z},\n\tJournal = {ACS Applied Materials and Interfaces},\n\tNumber = {15},\n\tPages = {7244--7252},\n\tTitle = {Aligned {SWNT} films from low-yield stress gels and their transparent electrode performance},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84882777892&partnerID=40&md5=9abbba93d13469527a497088d951477c},\n\tVolume = {5},\n\tYear = {2013},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84882777892&partnerID=40&md5=9abbba93d13469527a497088d951477c}}\n\n
\n
\n\n\n
\n Carbon nanotube films are promising for transparent electrodes for solar cells and displays. Large-area alignment of the nanotubes in these films is needed to minimize the sheet resistance. We present a novel coating method to coat high-density, aligned nanotubes over large areas. Carbon nanotube gel dispersions used in this study have aligned domains and a low yield stress. A simple shearing force allows these domains to uniformly align. We use this to correlate the transparent electrode performance of single-walled carbon nanotube films with the level of partial alignment. We have found that the transparent electrode performance improves with increasing levels of alignment and in a manner slightly better than what has been previously predicted. textcopyright 2013 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n In-situ quantification of the interfacial rheological response of bacterial biofilms to environmental stimuli.\n \n \n \n\n\n \n Rühs, P A; Böni, L; Fuller, G.; Inglis, R F; and Fischer, P\n\n\n \n\n\n\n PLoS ONE, 8(11): e78524–e78524. 2013.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{ruhs_-situ_2013,\n\tAuthor = {R{\\"u}hs, P A and B{\\"o}ni, L and Fuller, G.G. and Inglis, R F and Fischer, P},\n\tJournal = {PLoS ONE},\n\tNumber = {11},\n\tPages = {e78524--e78524},\n\tTitle = {In-situ quantification of the interfacial rheological response of bacterial biofilms to environmental stimuli},\n\tVolume = {8},\n\tYear = {2013}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Disruption of escherichia coli amyloid-integrated biofilm formation at the air-liquid interface by a polysorbate surfactant.\n \n \n \n \n\n\n \n Wu, C; Lim, J Y; Fuller, G.; and Cegelski, L\n\n\n \n\n\n\n Langmuir, 29(3): 920–926. 2013.\n \n\n\n\n
\n\n\n\n \n \n \"DisruptionPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wu_disruption_2013,\n\tAbstract = {Functional amyloid fibers termed curli contribute to bacterial adhesion and biofilm formation in Escherichia coli. We discovered that the nonionic surfactant Tween 20 inhibits biofilm formation by uropathogenic E. coli at the air-liquid interface, referred to as pellicle formation, and at the solid-liquid interface. At Tween 20 concentrations near and above the critical micelle concentration, the interfacial viscoelastic modulus is reduced to zero as cellular aggregates at the air-liquid interface are locally disconnected and eventually eliminated. Tween 20 does not inhibit the production of curli but prevents curli-integrated film formation. Our results support a model in which the hydrophobic curli fibers associated with bacteria near the air-liquid interface require access to the gas phase to formed strong physical entanglements and to form a network that can support shear stress. textcopyright 2012 American Chemical Society.},\n\tAuthor = {Wu, C and Lim, J Y and Fuller, G.G. and Cegelski, L},\n\tJournal = {Langmuir},\n\tNumber = {3},\n\tPages = {920--926},\n\tTitle = {Disruption of escherichia coli amyloid-integrated biofilm formation at the air-liquid interface by a polysorbate surfactant},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84872744999&partnerID=40&md5=6f7cdf11d62d0ecf2222407de00ecda9},\n\tVolume = {29},\n\tYear = {2013},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84872744999&partnerID=40&md5=6f7cdf11d62d0ecf2222407de00ecda9}}\n\n
\n
\n\n\n
\n Functional amyloid fibers termed curli contribute to bacterial adhesion and biofilm formation in Escherichia coli. We discovered that the nonionic surfactant Tween 20 inhibits biofilm formation by uropathogenic E. coli at the air-liquid interface, referred to as pellicle formation, and at the solid-liquid interface. At Tween 20 concentrations near and above the critical micelle concentration, the interfacial viscoelastic modulus is reduced to zero as cellular aggregates at the air-liquid interface are locally disconnected and eventually eliminated. Tween 20 does not inhibit the production of curli but prevents curli-integrated film formation. Our results support a model in which the hydrophobic curli fibers associated with bacteria near the air-liquid interface require access to the gas phase to formed strong physical entanglements and to form a network that can support shear stress. textcopyright 2012 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Structural and rheological properties of meibomian lipid.\n \n \n \n \n\n\n \n Rosenfeld, L; Cerretani, C; Leiske, D.; Toney, M F; Radke, C J; and Fuller, G.\n\n\n \n\n\n\n Investigative ophthalmology & visual science, 54(4): 2720–2732. 2013.\n \n\n\n\n
\n\n\n\n \n \n \"StructuralPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{rosenfeld_structural_2013,\n\tAuthor = {Rosenfeld, L and Cerretani, C and Leiske, D.L. and Toney, M F and Radke, C J and Fuller, G.G.},\n\tDoi = {10.1167/iovs.12-10987},\n\tJournal = {Investigative ophthalmology \\& visual science},\n\tNumber = {4},\n\tPages = {2720--2732},\n\tTitle = {Structural and rheological properties of meibomian lipid},\n\tUrl = {http://www.scopus.com/scopus/record/display.url?fedsrfIntegrator=MEKPAPERS-SCOCIT&origin=fedsrf&view=basic&eid=2-s2.0-84876479100},\n\tVolume = {54},\n\tYear = {2013},\n\tBdsk-Url-1 = {http://www.scopus.com/scopus/record/display.url?fedsrfIntegrator=MEKPAPERS-SCOCIT&origin=fedsrf&view=basic&eid=2-s2.0-84876479100},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1167/iovs.12-10987}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Spatial patterning of endothelium modulates cell morphology, adhesiveness and transcriptional signature.\n \n \n \n \n\n\n \n Huang, N F; Lai, E S; Ribeiro, A J S; Pan, S; Pruitt, B L; Fuller, G.; and Cooke, J P\n\n\n \n\n\n\n Biomaterials, 34(12): 2928–2937. 2013.\n \n\n\n\n
\n\n\n\n \n \n \"SpatialPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{huang_spatial_2013,\n\tAbstract = {Microscale and nanoscale structures can spatially pattern endothelial cells (ECs) into parallel-aligned organization, mimicking their cellular alignment in blood vessels exposed to laminar shear stress. However, the effects of spatial patterning on the function and global transcriptome of ECs are incompletely characterized. We used both parallel-aligned micropatterned and nanopatterned biomaterials to evaluate the effects of spatial patterning on the phenotype of ECs, based on gene expression profiling, functional characterization of monocyte adhesion, and quantification of cellular morphology. We demonstrate that both micropatterned and aligned nanofibrillar biomaterials could effectively guide EC organization along the direction of the micropatterned channels or nanofibrils, respectively. The ability of ECs to sense spatial patterning cues were abrogated in the presence of cytoskeletal disruption agents. Moreover, both micropatterned and aligned nanofibrillar substrates promoted an athero-resistant EC phenotype by reducing endothelial adhesiveness for monocytes and platelets, as well as by downregulating the expression of adhesion proteins and chemokines. We further found that micropatterned ECs have a transcriptional signature that is unique from non-patterned ECs, as well as from ECs aligned by shear stress. These findings highlight the importance of spatial patterning cues in guiding EC organization and function, which may have clinical relevance in the development of vascular grafts that promote patency. textcopyright 2013 Elsevier Ltd.},\n\tAuthor = {Huang, N F and Lai, E S and Ribeiro, A J S and Pan, S and Pruitt, B L and Fuller, G.G. and Cooke, J P},\n\tJournal = {Biomaterials},\n\tNumber = {12},\n\tPages = {2928--2937},\n\tTitle = {Spatial patterning of endothelium modulates cell morphology, adhesiveness and transcriptional signature},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84873473274&partnerID=40&md5=e880594361501a8198ff447671e6b254},\n\tVolume = {34},\n\tYear = {2013},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84873473274&partnerID=40&md5=e880594361501a8198ff447671e6b254}}\n\n
\n
\n\n\n
\n Microscale and nanoscale structures can spatially pattern endothelial cells (ECs) into parallel-aligned organization, mimicking their cellular alignment in blood vessels exposed to laminar shear stress. However, the effects of spatial patterning on the function and global transcriptome of ECs are incompletely characterized. We used both parallel-aligned micropatterned and nanopatterned biomaterials to evaluate the effects of spatial patterning on the phenotype of ECs, based on gene expression profiling, functional characterization of monocyte adhesion, and quantification of cellular morphology. We demonstrate that both micropatterned and aligned nanofibrillar biomaterials could effectively guide EC organization along the direction of the micropatterned channels or nanofibrils, respectively. The ability of ECs to sense spatial patterning cues were abrogated in the presence of cytoskeletal disruption agents. Moreover, both micropatterned and aligned nanofibrillar substrates promoted an athero-resistant EC phenotype by reducing endothelial adhesiveness for monocytes and platelets, as well as by downregulating the expression of adhesion proteins and chemokines. We further found that micropatterned ECs have a transcriptional signature that is unique from non-patterned ECs, as well as from ECs aligned by shear stress. These findings highlight the importance of spatial patterning cues in guiding EC organization and function, which may have clinical relevance in the development of vascular grafts that promote patency. textcopyright 2013 Elsevier Ltd.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n 3-Hydroxybutyric Acid Interacts with Lipid Monolayers at Concentrations That Impair Consciousness.\n \n \n \n \n\n\n \n Hsu, T. T; Leiske, D. L; Rosenfeld, L.; Sonner, J. M; and Fuller, G. G\n\n\n \n\n\n\n Langmuir, 29(6): 1948–1955. February 2013.\n \n\n\n\n
\n\n\n\n \n \n \"3-HydroxybutyricPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hsu_3-hydroxybutyric_2013,\n\tAbstract = {3-Hydroxybutyric acid (also referred to as β-hydroxybutyric acid or BHB), a small molecule metabolite whose concentration is elevated in type I diabetes and diabetic coma, was found to modulate the properties of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) monolayers when added to the subphase at clinical concentrations. This is a key piece of evidence supporting the hypothesis that the anesthetic actions of BHB are due to the metabolite's abilities to alter physical properties of cell membranes, leading to indirect effects on membrane protein function. Pressure--area isotherms show that BHB changes the compressibility of the monolayer and decrease the size of the two-phase coexistence region. Epi-fluorescent microscopy further reveals that the reduction of the coexistence region is due to the significant reduction in morphology of the liquid condensed domains in the two-phase coexistence region. These changes in monolayer morphology are associated with the diminished interfacial viscosity of the mon...},\n\tAuthor = {Hsu, Tienyi T and Leiske, Danielle L and Rosenfeld, Liat and Sonner, James M and Fuller, Gerald G},\n\tDoi = {10.1021/la304712f},\n\tJournal = {Langmuir},\n\tLanguage = {English},\n\tMonth = feb,\n\tNumber = {6},\n\tPages = {1948--1955},\n\tTitle = {3-{Hydroxybutyric} {Acid} {Interacts} with {Lipid} {Monolayers} at {Concentrations} {That} {Impair} {Consciousness}},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/la304712f},\n\tVolume = {29},\n\tYear = {2013},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/la304712f},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/la304712f}}\n\n
\n
\n\n\n
\n 3-Hydroxybutyric acid (also referred to as β-hydroxybutyric acid or BHB), a small molecule metabolite whose concentration is elevated in type I diabetes and diabetic coma, was found to modulate the properties of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) monolayers when added to the subphase at clinical concentrations. This is a key piece of evidence supporting the hypothesis that the anesthetic actions of BHB are due to the metabolite's abilities to alter physical properties of cell membranes, leading to indirect effects on membrane protein function. Pressure–area isotherms show that BHB changes the compressibility of the monolayer and decrease the size of the two-phase coexistence region. Epi-fluorescent microscopy further reveals that the reduction of the coexistence region is due to the significant reduction in morphology of the liquid condensed domains in the two-phase coexistence region. These changes in monolayer morphology are associated with the diminished interfacial viscosity of the mon...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Oriented, polymer-stabilized carbon nanotube films: Influence of dispersion rheology.\n \n \n \n \n\n\n \n Allen, R; Bao, Z; and Fuller, G.\n\n\n \n\n\n\n Nanotechnology, 24(1). 2013.\n \n\n\n\n
\n\n\n\n \n \n \"Oriented,Paper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{allen_oriented_2013,\n\tAbstract = {Thin carbon nanotube films have great potential for transparent electrodes for solar cells and displays. One advantage for using carbon nanotubes is the potential for solution processing. However, research has not been done to connect solution rheological properties with the corresponding film characteristics. Here we study the rheological properties of single-walled carbon nanotube/polythiophene composite dispersions to better understand the alignment that can be achieved during deposition. Several parameters are varied to explore the cause of the alignment and the requirements of achieving a uniform, aligned carbon nanotube/polythiophene film. By understanding the dispersions thoroughly, the film quality can be predicted. textcopyright 2013 IOP Publishing Ltd.},\n\tAuthor = {Allen, R and Bao, Z and Fuller, G.G.},\n\tJournal = {Nanotechnology},\n\tNumber = {1},\n\tTitle = {Oriented, polymer-stabilized carbon nanotube films: {Influence} of dispersion rheology},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84871019910&partnerID=40&md5=de93355b84239fb0acf1b2ece973f8bc},\n\tVolume = {24},\n\tYear = {2013},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84871019910&partnerID=40&md5=de93355b84239fb0acf1b2ece973f8bc}}\n\n
\n
\n\n\n
\n Thin carbon nanotube films have great potential for transparent electrodes for solar cells and displays. One advantage for using carbon nanotubes is the potential for solution processing. However, research has not been done to connect solution rheological properties with the corresponding film characteristics. Here we study the rheological properties of single-walled carbon nanotube/polythiophene composite dispersions to better understand the alignment that can be achieved during deposition. Several parameters are varied to explore the cause of the alignment and the requirements of achieving a uniform, aligned carbon nanotube/polythiophene film. By understanding the dispersions thoroughly, the film quality can be predicted. textcopyright 2013 IOP Publishing Ltd.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The modulation of endothelial cell morphology, function, and survival using anisotropic nanofibrillar collagen scaffolds.\n \n \n \n \n\n\n \n Huang, N. F; Okogbaa, J.; Lee, J. C; Jha, A.; Zaitseva, T. S; Paukshto, M. V; Sun, J. S; Punjya, N.; Fuller, G. G; and Cooke, J. P\n\n\n \n\n\n\n Biomaterials, 34(16): 4038–4047. May 2013.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{huang_modulation_2013,\n\tAbstract = {Endothelial cells (ECs) are aligned longitudinally under laminar flow, whereas they are polygonal and poorly aligned in regions of disturbed flow. The unaligned ECs in disturbed flow fields manifest altered function and reduced survival that promote lesion formation. We demonstrate that the alignment of the ECs may directly influence their biology, independent of fluid flow. We developed aligned nanofibrillar collagen scaffolds that mimic the structure of collagen bundles in blood vessels, and examined the effects of these materials on EC alignment, function, and in vivo survival. ECs cultured on 30-nm diameter aligned fibrils re-organized their F-actin along the nanofibril direction, and were 50\\% less adhesive for monocytes than the ECs grown on randomly oriented fibrils. After EC transplantation into both subcutaneous tissue and the ischemic hindlimb, EC viability was enhanced when ECs were cultured and implanted on aligned nanofibrillar scaffolds, in contrast to non-patterned scaffolds. ECs derived from human induced pluripotent stem cells and cultured on aligned scaffolds also persisted for over 28 days, as assessed by bioluminescence imaging, when implanted in ischemic tissue. By contrast, ECs implanted on scaffolds without nanopatterning generated no detectable bioluminescent signal by day 4 in either normal or ischemic tissues. We demonstrate that 30-nm aligned nanofibrillar collagen scaffolds guide cellular organization, modulate endothelial inflammatory response, and enhance cell survival after implantation in normal and ischemic tissues. textcopyright 2013.},\n\tAuthor = {Huang, Ngan F and Okogbaa, Janet and Lee, Jerry C and Jha, Arshi and Zaitseva, Tatiana S and Paukshto, Michael V and Sun, John S and Punjya, Niraj and Fuller, Gerald G and Cooke, John P},\n\tDoi = {10.1016/j.biomaterials.2013.02.036},\n\tJournal = {Biomaterials},\n\tMonth = may,\n\tNumber = {16},\n\tPages = {4038--4047},\n\tPmid = {23480958},\n\tTitle = {The modulation of endothelial cell morphology, function, and survival using anisotropic nanofibrillar collagen scaffolds},\n\tUrl = {http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3695739&tool=pmcentrez&rendertype=abstract http://linkinghub.elsevier.com/retrieve/pii/S0142961213002196},\n\tVolume = {34},\n\tYear = {2013},\n\tBdsk-Url-1 = {http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3695739&tool=pmcentrez&rendertype=abstract%20http://linkinghub.elsevier.com/retrieve/pii/S0142961213002196},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1016/j.biomaterials.2013.02.036}}\n\n
\n
\n\n\n
\n Endothelial cells (ECs) are aligned longitudinally under laminar flow, whereas they are polygonal and poorly aligned in regions of disturbed flow. The unaligned ECs in disturbed flow fields manifest altered function and reduced survival that promote lesion formation. We demonstrate that the alignment of the ECs may directly influence their biology, independent of fluid flow. We developed aligned nanofibrillar collagen scaffolds that mimic the structure of collagen bundles in blood vessels, and examined the effects of these materials on EC alignment, function, and in vivo survival. ECs cultured on 30-nm diameter aligned fibrils re-organized their F-actin along the nanofibril direction, and were 50% less adhesive for monocytes than the ECs grown on randomly oriented fibrils. After EC transplantation into both subcutaneous tissue and the ischemic hindlimb, EC viability was enhanced when ECs were cultured and implanted on aligned nanofibrillar scaffolds, in contrast to non-patterned scaffolds. ECs derived from human induced pluripotent stem cells and cultured on aligned scaffolds also persisted for over 28 days, as assessed by bioluminescence imaging, when implanted in ischemic tissue. By contrast, ECs implanted on scaffolds without nanopatterning generated no detectable bioluminescent signal by day 4 in either normal or ischemic tissues. We demonstrate that 30-nm aligned nanofibrillar collagen scaffolds guide cellular organization, modulate endothelial inflammatory response, and enhance cell survival after implantation in normal and ischemic tissues. textcopyright 2013.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Tracking the interfacial dynamics of PNiPAM soft microgels particles adsorbed at the airtextendashwater interface and in thin liquid films.\n \n \n \n \n\n\n \n Cohin, Y.; Fisson, M.; Jourde, K e v.; Fuller, G. G; Sanson, N.; Talini, L.; and Monteux, C e c.\n\n\n \n\n\n\n Rheologica Acta, 52(5): 445–454. 2013.\n \n\n\n\n
\n\n\n\n \n \n \"TrackingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{cohin_tracking_2013,\n\tAuthor = {Cohin, Yann and Fisson, Maelle and Jourde, K e vin and Fuller, Gerald G and Sanson, Nicolas and Talini, Laurence and Monteux, C e cile},\n\tDoi = {10.1007/s00397-013-0697-3},\n\tJournal = {Rheologica Acta},\n\tNumber = {5},\n\tPages = {445--454},\n\tTitle = {Tracking the interfacial dynamics of {PNiPAM} soft microgels particles adsorbed at the airtextendashwater interface and in thin liquid films},\n\tUrl = {http://link.springer.com/article/10.1007/s00397-013-0697-3/fulltext.html},\n\tVolume = {52},\n\tYear = {2013},\n\tBdsk-Url-1 = {http://link.springer.com/article/10.1007/s00397-013-0697-3/fulltext.html},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1007/s00397-013-0697-3}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2012\n \n \n (11)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Temperature-Induced Transitions in the Structure and Interfacial Rheology of Human Meibum.\n \n \n \n \n\n\n \n Leiske, D. L; Leiske, C. I; Leiske, D. R; Toney, M. F; Senchyna, M.; Ketelson, H. A; Meadows, D. L; and Fuller, G. G\n\n\n \n\n\n\n Biophysical Journal, 102(2): 369–376. 2012.\n \n\n\n\n
\n\n\n\n \n \n \"Temperature-InducedPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{leiske_temperature-induced_2012,\n\tAbstract = {Meibomian lipids are the primary component of the lipid layer of the tear film. Composed primarily of a mixture of lipids, meibum exhibits a range of melt temperatures. Compositional changes that occur with disease may alter the temperature at which meibum melts. Here we explore how the mechanical properties and structure of meibum from healthy subjects depend on temperature. Interfacial films of meibum were highly viscoelastic at 17textdegreeC, but as the films were heated to 30textdegreeC the surface moduli decreased by more than two orders of magnitude. Brewster angle microscopy revealed the presence of micron-scale inhomogeneities in meibum films at higher temperatures. Crystalline structure was probed by small angle x-ray scattering of bulk meibum, which showed evidence of a majority crystalline structure in all samples with lamellar spacing of 49 that melted at 34textdegreeC. A minority structure was observed in some samples with d-spacing at 110 that persisted up to 40textdegreeC. The melting of crystalline phases accompanied by a reduction in interfacial viscosity and elasticity has implications in meibum behavior in the tear film. If the melt temperature of meibum was altered significantly from disease-induced compositional changes, the resultant change in viscosity could alter secretion of lipids from meibomian glands, or tear-film stabilization properties of the lipid layer. textcopyright 2012 Biophysical Society.},\n\tAuthor = {Leiske, Danielle L and Leiske, Christopher I and Leiske, Daniel R and Toney, Michael F and Senchyna, Michelle and Ketelson, Howard A and Meadows, David L and Fuller, Gerald G},\n\tDoi = {10.1016/j.bpj.2011.12.017},\n\tJournal = {Biophysical Journal},\n\tNumber = {2},\n\tPages = {369--376},\n\tTitle = {Temperature-{Induced} {Transitions} in the {Structure} and {Interfacial} {Rheology} of {Human} {Meibum}},\n\tUrl = {http://linkinghub.elsevier.com/retrieve/pii/S0006349511054191},\n\tVolume = {102},\n\tYear = {2012},\n\tBdsk-Url-1 = {http://linkinghub.elsevier.com/retrieve/pii/S0006349511054191},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1016/j.bpj.2011.12.017}}\n\n
\n
\n\n\n
\n Meibomian lipids are the primary component of the lipid layer of the tear film. Composed primarily of a mixture of lipids, meibum exhibits a range of melt temperatures. Compositional changes that occur with disease may alter the temperature at which meibum melts. Here we explore how the mechanical properties and structure of meibum from healthy subjects depend on temperature. Interfacial films of meibum were highly viscoelastic at 17textdegreeC, but as the films were heated to 30textdegreeC the surface moduli decreased by more than two orders of magnitude. Brewster angle microscopy revealed the presence of micron-scale inhomogeneities in meibum films at higher temperatures. Crystalline structure was probed by small angle x-ray scattering of bulk meibum, which showed evidence of a majority crystalline structure in all samples with lamellar spacing of 49 that melted at 34textdegreeC. A minority structure was observed in some samples with d-spacing at 110 that persisted up to 40textdegreeC. The melting of crystalline phases accompanied by a reduction in interfacial viscosity and elasticity has implications in meibum behavior in the tear film. If the melt temperature of meibum was altered significantly from disease-induced compositional changes, the resultant change in viscosity could alter secretion of lipids from meibomian glands, or tear-film stabilization properties of the lipid layer. textcopyright 2012 Biophysical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Role of shear-thinning on the dynamics of rinsing flow by an impinging jet.\n \n \n \n \n\n\n \n Walker, T. W; Hsu, T. T; Frank, C. W; and Fuller, G. G\n\n\n \n\n\n\n Physics of Fluids, 24(9): 93102–93116. September 2012.\n \n\n\n\n
\n\n\n\n \n \n \"RolePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{walker_role_2012,\n\tAuthor = {Walker, Travis W and Hsu, Tienyi T and Frank, Curtis W and Fuller, Gerald G},\n\tDoi = {doi:10.1063/1.4752765},\n\tJournal = {Physics of Fluids},\n\tMonth = sep,\n\tNumber = {9},\n\tPages = {93102--93116},\n\tTitle = {Role of shear-thinning on the dynamics of rinsing flow by an impinging jet},\n\tUrl = {http://link.aip.org/link/?PHFLE6/24/093102/1},\n\tVolume = {24},\n\tYear = {2012},\n\tBdsk-Url-1 = {http://link.aip.org/link/?PHFLE6/24/093102/1},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1063/1.4752765}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Quantitative analysis of amyloid-integrated biofilms formed by uropathogenic escherichia coli at the air-liquid interface.\n \n \n \n \n\n\n \n Wu, C; Lim, J Y; Fuller, G.; and Cegelski, L\n\n\n \n\n\n\n Biophysical Journal, 103(3): 464–471. 2012.\n \n\n\n\n
\n\n\n\n \n \n \"QuantitativePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wu_quantitative_2012,\n\tAbstract = {Bacterial biofilms are complex multicellular assemblies, characterized by a heterogeneous extracellular polymeric matrix, that have emerged as hallmarks of persistent infectious diseases. New approaches and quantitative data are needed to elucidate the composition and architecture of biofilms, and such data need to be correlated with mechanical and physicochemical properties that relate to function. We performed a panel of interfacial rheological measurements during biofilm formation at the air-liquid interface by the Escherichia coli strain UTI89, which is noted for its importance in studies of urinary tract infection and for its assembly of functional amyloid fibers termed curli. Brewster-angle microscopy and measurements of the surface elasticity (Gstextasciiacute) and stress-strain response provided sensitive and quantitative parameters that revealed distinct stages during bacterial colonization, aggregation, and eventual formation of a pellicle at the air-liquid interface. Pellicles that formed under conditions that upregulate curli production exhibited an increase in strength and viscoelastic properties as well as a greater ability to recover from stress-strain perturbation. The results suggest that curli, as hydrophobic extracellular amyloid fibers, enhance the strength, viscoelasticity, and resistance to strain of E. coli biofilms formed at the air-liquid interface. textcopyright 2012 by the Biophysical Society.},\n\tAuthor = {Wu, C and Lim, J Y and Fuller, G.G. and Cegelski, L},\n\tJournal = {Biophysical Journal},\n\tNumber = {3},\n\tPages = {464--471},\n\tTitle = {Quantitative analysis of amyloid-integrated biofilms formed by uropathogenic escherichia coli at the air-liquid interface},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84864717953&partnerID=40&md5=1e4286733a4716dd24fcd0f1820030d4},\n\tVolume = {103},\n\tYear = {2012},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84864717953&partnerID=40&md5=1e4286733a4716dd24fcd0f1820030d4}}\n\n
\n
\n\n\n
\n Bacterial biofilms are complex multicellular assemblies, characterized by a heterogeneous extracellular polymeric matrix, that have emerged as hallmarks of persistent infectious diseases. New approaches and quantitative data are needed to elucidate the composition and architecture of biofilms, and such data need to be correlated with mechanical and physicochemical properties that relate to function. We performed a panel of interfacial rheological measurements during biofilm formation at the air-liquid interface by the Escherichia coli strain UTI89, which is noted for its importance in studies of urinary tract infection and for its assembly of functional amyloid fibers termed curli. Brewster-angle microscopy and measurements of the surface elasticity (Gstextasciiacute) and stress-strain response provided sensitive and quantitative parameters that revealed distinct stages during bacterial colonization, aggregation, and eventual formation of a pellicle at the air-liquid interface. Pellicles that formed under conditions that upregulate curli production exhibited an increase in strength and viscoelastic properties as well as a greater ability to recover from stress-strain perturbation. The results suggest that curli, as hydrophobic extracellular amyloid fibers, enhance the strength, viscoelasticity, and resistance to strain of E. coli biofilms formed at the air-liquid interface. textcopyright 2012 by the Biophysical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Complex Fluid-Fluid Interfaces: Rheology and Structure.\n \n \n \n \n\n\n \n Fuller, G. G; and Vermant, J.\n\n\n \n\n\n\n Annual Review of Chemical and Biomolecular Engineering, 3(1): 519–543. 2012.\n \n\n\n\n
\n\n\n\n \n \n \"ComplexPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_complex_2012,\n\tAuthor = {Fuller, Gerald G and Vermant, Jan},\n\tDoi = {10.1146/annurev-chembioeng-061010-114202},\n\tJournal = {Annual Review of Chemical and Biomolecular Engineering},\n\tNumber = {1},\n\tPages = {519--543},\n\tTitle = {Complex {Fluid}-{Fluid} {Interfaces}: {Rheology} and {Structure}},\n\tUrl = {http://www.annualreviews.org/doi/abs/10.1146/annurev-chembioeng-061010-114202},\n\tVolume = {3},\n\tYear = {2012},\n\tBdsk-Url-1 = {http://www.annualreviews.org/doi/abs/10.1146/annurev-chembioeng-061010-114202},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1146/annurev-chembioeng-061010-114202}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Aligned nanofibrillar collagen regulates endothelial organization and migration.\n \n \n \n\n\n \n Lai, E S; Huang, N F; Cooke, J P; and Fuller, G.\n\n\n \n\n\n\n Regenerative Medicine, 7(5): 649–661. 2012.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lai_aligned_2012,\n\tAbstract = {Aim: Modulating endothelial cell (EC) morphology and motility, with the aim to influence their biology, might be beneficial for the treatment of vascular disease. We examined the effect of nanoscale matrix anisotropy on EC organization and migration for vascular tissue engineering applications. Materials \\{\\vphantom{\\}}\\&},\n\tAuthor = {Lai, E S and Huang, N F and Cooke, J P and Fuller, G.G.},\n\tJournal = {Regenerative Medicine},\n\tNumber = {5},\n\tPages = {649--661},\n\tTitle = {Aligned nanofibrillar collagen regulates endothelial organization and migration},\n\tVolume = {7},\n\tYear = {2012}}\n\n
\n
\n\n\n
\n Aim: Modulating endothelial cell (EC) morphology and motility, with the aim to influence their biology, might be beneficial for the treatment of vascular disease. We examined the effect of nanoscale matrix anisotropy on EC organization and migration for vascular tissue engineering applications. Materials \\p̌hantom\\&\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Interfacial rheology of natural silk fibroin at air/water and oil/water interfaces.\n \n \n \n \n\n\n \n Wang, L; Xie, H; Qiao, X; Goffin, A; Hodgkinson, T; Yuan, X; Sun, K; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 28(1): 459–467. 2012.\n \n\n\n\n
\n\n\n\n \n \n \"InterfacialPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wang_interfacial_2012,\n\tAbstract = {The interfacial viscoelastic behavior of natural silk fibroin at both the air/water and oil/water interfaces is reported. This natural multiblock copolymer is found to be strongly amphiphilic and forms stable films at these interfaces. The result is an interfacial layer that is rheologically complex with strong surface elastic moduli that are only slightly frequency-dependent. The kinetics of surface viscoelastic evolution are reported as functions of time for various concentrations of the spread films. Films deposited by Langmuir-Blodgett deposition were studied by scanning electron microscopy (SEM) to reveal a fibrous structure at the interface. The production of stable O/W emulsions by silk fibroin further confirms the generation of the elastic films at the oil/water interfaces. textcopyright 2011 American Chemical Society.},\n\tAuthor = {Wang, L and Xie, H and Qiao, X and Goffin, A and Hodgkinson, T and Yuan, X and Sun, K and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {1},\n\tPages = {459--467},\n\tTitle = {Interfacial rheology of natural silk fibroin at air/water and oil/water interfaces},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84862908737&partnerID=40&md5=b1f3142d359c49fe71498fbf359d09ca},\n\tVolume = {28},\n\tYear = {2012},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84862908737&partnerID=40&md5=b1f3142d359c49fe71498fbf359d09ca}}\n\n
\n
\n\n\n
\n The interfacial viscoelastic behavior of natural silk fibroin at both the air/water and oil/water interfaces is reported. This natural multiblock copolymer is found to be strongly amphiphilic and forms stable films at these interfaces. The result is an interfacial layer that is rheologically complex with strong surface elastic moduli that are only slightly frequency-dependent. The kinetics of surface viscoelastic evolution are reported as functions of time for various concentrations of the spread films. Films deposited by Langmuir-Blodgett deposition were studied by scanning electron microscopy (SEM) to reveal a fibrous structure at the interface. The production of stable O/W emulsions by silk fibroin further confirms the generation of the elastic films at the oil/water interfaces. textcopyright 2011 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Interfacial and fluorescence studies on stereoblock poly(N -isopropylacryl amide)s.\n \n \n \n \n\n\n \n Niskanen, J; Wu, C; Ostrowski, M; Fuller, G.; Tenhu, H; and Hietala, S\n\n\n \n\n\n\n Langmuir, 28(41): 14792–14798. 2012.\n \n\n\n\n
\n\n\n\n \n \n \"InterfacialPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{niskanen_interfacial_2012,\n\tAbstract = {Aqueous solution and water-air interfacial properties of associative thermally responsive A-B-A stereoblock poly(N-isopropylacryl amide), PNIPAM, polymers were studied and compared to atactic PNIPAM. The A-B-A polymers consist of atactic PNIPAM as a hydrophilic block (either A or B) and a water-insoluble block of isotactic PNIPAM. The surface tensions of aqueous PNIPAM solutions were measured as a function of both temperature and concentration. The isotactic blocks did not have an effect on the surface activity of the solutions. Rheological measurements on the water-air interface showed that the aggregated PNIPAMs containing isotactic blocks increased the elasticity of the surface significantly as compared to the atactic reference upon heating. Two fluorescence probes, pyrene and (4-(dicyanomethylene)-2-methyl-6-(4- dimethylaminostyryl)-4H-pyran (4HP), added to the aqueous polymer solutions were concluded to reside in surroundings with lower polarity and increased microviscosity in cases when the polymers contained isotactic blocks, as compared to ordinary atactic polymers. textcopyright 2012 American Chemical Society.},\n\tAuthor = {Niskanen, J and Wu, C and Ostrowski, M and Fuller, G.G. and Tenhu, H and Hietala, S},\n\tJournal = {Langmuir},\n\tNumber = {41},\n\tPages = {14792--14798},\n\tTitle = {Interfacial and fluorescence studies on stereoblock poly({N} -isopropylacryl amide)s},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84867570609&partnerID=40&md5=5baf3c43c422209a816d0aa83c526882},\n\tVolume = {28},\n\tYear = {2012},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-84867570609&partnerID=40&md5=5baf3c43c422209a816d0aa83c526882}}\n\n
\n
\n\n\n
\n Aqueous solution and water-air interfacial properties of associative thermally responsive A-B-A stereoblock poly(N-isopropylacryl amide), PNIPAM, polymers were studied and compared to atactic PNIPAM. The A-B-A polymers consist of atactic PNIPAM as a hydrophilic block (either A or B) and a water-insoluble block of isotactic PNIPAM. The surface tensions of aqueous PNIPAM solutions were measured as a function of both temperature and concentration. The isotactic blocks did not have an effect on the surface activity of the solutions. Rheological measurements on the water-air interface showed that the aggregated PNIPAMs containing isotactic blocks increased the elasticity of the surface significantly as compared to the atactic reference upon heating. Two fluorescence probes, pyrene and (4-(dicyanomethylene)-2-methyl-6-(4- dimethylaminostyryl)-4H-pyran (4HP), added to the aqueous polymer solutions were concluded to reside in surroundings with lower polarity and increased microviscosity in cases when the polymers contained isotactic blocks, as compared to ordinary atactic polymers. textcopyright 2012 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Oriented collagen as a potential cochlear implant electrode surface coating to achieve directed neurite outgrowth.\n \n \n \n \n\n\n \n Volkenstein, S.; Kirkwood, J. E; Lai, E.; Dazert, S.; Fuller, G. G; and Heller, S.\n\n\n \n\n\n\n European archives of oto-rhino-laryngology : official journal of the European Federation of Oto-Rhino-Laryngological Societies (EUFOS) : affiliated with the German Society for Oto-Rhino-Laryngology - Head and Neck Surgery, 269(4): 1111–1116. April 2012.\n \n\n\n\n
\n\n\n\n \n \n \"OrientedPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{volkenstein_oriented_2012,\n\tAbstract = {In patients with severe to profound hearing loss, cochlear implants (CIs) are currently the only therapeutic option when the amplification with conventional hearing aids does no longer lead to a useful hearing experience. Despite its great success, there are patients in which benefit from these devices is rather limited. One reason may be a poor neuron-device interaction, where the electric fields generated by the electrode array excite a wide range of tonotopically organized spiral ganglion neurons at the cost of spatial resolution. Coating of CI electrodes to provide a welcoming environment combined with suitable surface chemistry (e.g. with neurotrophic factors) has been suggested to create a closer bioelectrical interface between the electrode array and the target tissue, which might lead to better spatial resolution, better frequency discrimination, and ultimately may improve speech perception in patients. Here we investigate the use of a collagen surface with a cholesteric banding structure, whose orientation can be systemically controlled as a guiding structure for neurite outgrowth. We demonstrate that spiral ganglion neurons survive on collagen-coated surfaces and display a directed neurite growth influenced by the direction of collagen fibril deposition. The majority of neurites grow parallel to the orientation direction of the collagen. We suggest collagen coating as a possible future option in CI technology to direct neurite outgrowth and improve hearing results for affected patients.},\n\tAuthor = {Volkenstein, Stefan and Kirkwood, John E and Lai, Edwina and Dazert, Stefan and Fuller, Gerald G and Heller, Stefan},\n\tDoi = {10.1007/s00405-011-1775-8},\n\tJournal = {European archives of oto-rhino-laryngology : official journal of the European Federation of Oto-Rhino-Laryngological Societies (EUFOS) : affiliated with the German Society for Oto-Rhino-Laryngology - Head and Neck Surgery},\n\tLanguage = {English},\n\tMonth = apr,\n\tNumber = {4},\n\tPages = {1111--1116},\n\tPmid = {21952794},\n\tTitle = {Oriented collagen as a potential cochlear implant electrode surface coating to achieve directed neurite outgrowth.},\n\tUrl = {http://link.springer.com/10.1007/s00405-011-1775-8},\n\tVolume = {269},\n\tYear = {2012},\n\tBdsk-Url-1 = {http://link.springer.com/10.1007/s00405-011-1775-8},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1007/s00405-011-1775-8}}\n\n
\n
\n\n\n
\n In patients with severe to profound hearing loss, cochlear implants (CIs) are currently the only therapeutic option when the amplification with conventional hearing aids does no longer lead to a useful hearing experience. Despite its great success, there are patients in which benefit from these devices is rather limited. One reason may be a poor neuron-device interaction, where the electric fields generated by the electrode array excite a wide range of tonotopically organized spiral ganglion neurons at the cost of spatial resolution. Coating of CI electrodes to provide a welcoming environment combined with suitable surface chemistry (e.g. with neurotrophic factors) has been suggested to create a closer bioelectrical interface between the electrode array and the target tissue, which might lead to better spatial resolution, better frequency discrimination, and ultimately may improve speech perception in patients. Here we investigate the use of a collagen surface with a cholesteric banding structure, whose orientation can be systemically controlled as a guiding structure for neurite outgrowth. We demonstrate that spiral ganglion neurons survive on collagen-coated surfaces and display a directed neurite growth influenced by the direction of collagen fibril deposition. The majority of neurites grow parallel to the orientation direction of the collagen. We suggest collagen coating as a possible future option in CI technology to direct neurite outgrowth and improve hearing results for affected patients.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Molecular structure of interfacial human meibum films.\n \n \n \n \n\n\n \n Leiske, D. L; Miller, C. E; Rosenfeld, L.; Cerretani, C.; Ayzner, A.; Lin, B.; Meron, M.; Senchyna, M.; Ketelson, H. A; Meadows, D.; Srinivasan, S.; Jones, L.; Radke, C. J; Toney, M. F; and Fuller, G. G\n\n\n \n\n\n\n Langmuir, 28(32): 11858–11865. August 2012.\n \n\n\n\n
\n\n\n\n \n \n \"MolecularPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{leiske_molecular_2012,\n\tAbstract = {Meibum is the primary component of the tear film lipid layer. Thought to play a role in tear film stabilization, understanding the physical properties of meibum and how they change with disease will be valuable in identifying dry eye treatment targets. Grazing incidence X-ray diffraction and X-ray reflectivity were applied to meibum films at an air-water interface to identify molecular organization. At room temperature, interfacial meibum films formed two coexisting scattering phases with rectangular lattices and next-nearest neighbor tilts, similar to the Ov phase previously identified in fatty acids. The intensity of the diffraction peaks increased with compression, although the lattice spacing and molecular tilt angle remained constant. Reflectivity measurements at surface pressures of 18 mN/m and above revealed multilayers with d-spacings of 50 {\\textbackslash}AA, suggesting that vertical organization rather than lateral was predominantly affected by meibum-film compression.},\n\tAuthor = {Leiske, Danielle L and Miller, Chad E and Rosenfeld, Liat and Cerretani, Colin and Ayzner, Alexander and Lin, Binhua and Meron, Mati and Senchyna, Michelle and Ketelson, Howard A and Meadows, David and Srinivasan, Sruthi and Jones, Lyndon and Radke, Clayton J and Toney, Michael F and Fuller, Gerald G},\n\tDoi = {10.1021/la301321r},\n\tJournal = {Langmuir},\n\tLanguage = {English},\n\tMonth = aug,\n\tNumber = {32},\n\tPages = {11858--11865},\n\tPmid = {22783994},\n\tTitle = {Molecular structure of interfacial human meibum films.},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/la301321r},\n\tVolume = {28},\n\tYear = {2012},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/la301321r},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/la301321r}}\n\n
\n
\n\n\n
\n Meibum is the primary component of the tear film lipid layer. Thought to play a role in tear film stabilization, understanding the physical properties of meibum and how they change with disease will be valuable in identifying dry eye treatment targets. Grazing incidence X-ray diffraction and X-ray reflectivity were applied to meibum films at an air-water interface to identify molecular organization. At room temperature, interfacial meibum films formed two coexisting scattering phases with rectangular lattices and next-nearest neighbor tilts, similar to the Ov phase previously identified in fatty acids. The intensity of the diffraction peaks increased with compression, although the lattice spacing and molecular tilt angle remained constant. Reflectivity measurements at surface pressures of 18 mN/m and above revealed multilayers with d-spacings of 50 \\AA, suggesting that vertical organization rather than lateral was predominantly affected by meibum-film compression.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Consequences of interfacial viscoelasticity on thin film stability.\n \n \n \n \n\n\n \n Rosenfeld, L.; and Fuller, G. G\n\n\n \n\n\n\n Langmuir, 28(40): 14238–14244. October 2012.\n \n\n\n\n
\n\n\n\n \n \n \"ConsequencesPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{rosenfeld_consequences_2012,\n\tAbstract = {The phenomenon of dewetting is frequently observed in our everyday life. It is of central importance in many technological applications as well as in a variety of physical and biological systems. The presence of nonsoluble surfactants at an air/liquid interface may affect the dewetting properties of the aqueous layer. An important example is the tear film, which comprises an aqueous layer covered with a ∼100-nm-thick blanket of lipids, known as the meibomian lipids. Interfacial rheological measurements of meibomian lipids reveal that these films are remarkably viscoelastic. Tear film dewetting is of central importance to understanding tear film stability. To better understand the role of surface viscoelasticity in tear film stability, we have developed a methodology to systematically control interfacial rheology of thin aqueous layers at the onset of dewetting events. The apparatus allows control over the surface pressure of the monolayer, which is a key feature since this variable controls the surface viscoelasticity. Three insoluble monolayer materials were used: newtonian arachidyl alcohol (AA), DPPC, a phospholipid that is slightly viscoelastic, and meibum, which produces a strongly viscoelastic monolayer. It is reported that monolayers of viscoelastic surfactants are able to stabilize thin films against spontaneous dewetting. As the surface pressure of these layers is increased, their effectiveness is enhanced. Moreover, these surfactants are able to reduce the critical film thickness for dewetting. Meibum is particularly effective in stabilizing thin films. Our results suggest that the meibomian lipids play a vital role in maintaining tear film stability in addition to suppressing evaporation.},\n\tAuthor = {Rosenfeld, Liat and Fuller, Gerald G},\n\tDoi = {10.1021/la302731z},\n\tJournal = {Langmuir},\n\tLanguage = {English},\n\tMonth = oct,\n\tNumber = {40},\n\tPages = {14238--14244},\n\tPmid = {22989061},\n\tTitle = {Consequences of interfacial viscoelasticity on thin film stability.},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/la302731z},\n\tVolume = {28},\n\tYear = {2012},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/la302731z},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/la302731z}}\n\n
\n
\n\n\n
\n The phenomenon of dewetting is frequently observed in our everyday life. It is of central importance in many technological applications as well as in a variety of physical and biological systems. The presence of nonsoluble surfactants at an air/liquid interface may affect the dewetting properties of the aqueous layer. An important example is the tear film, which comprises an aqueous layer covered with a ∼100-nm-thick blanket of lipids, known as the meibomian lipids. Interfacial rheological measurements of meibomian lipids reveal that these films are remarkably viscoelastic. Tear film dewetting is of central importance to understanding tear film stability. To better understand the role of surface viscoelasticity in tear film stability, we have developed a methodology to systematically control interfacial rheology of thin aqueous layers at the onset of dewetting events. The apparatus allows control over the surface pressure of the monolayer, which is a key feature since this variable controls the surface viscoelasticity. Three insoluble monolayer materials were used: newtonian arachidyl alcohol (AA), DPPC, a phospholipid that is slightly viscoelastic, and meibum, which produces a strongly viscoelastic monolayer. It is reported that monolayers of viscoelastic surfactants are able to stabilize thin films against spontaneous dewetting. As the surface pressure of these layers is increased, their effectiveness is enhanced. Moreover, these surfactants are able to reduce the critical film thickness for dewetting. Meibum is particularly effective in stabilizing thin films. Our results suggest that the meibomian lipids play a vital role in maintaining tear film stability in addition to suppressing evaporation.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Extensional rheometry at interfaces: Analysis of the Cambridge Interfacial Tensiometer.\n \n \n \n \n\n\n \n Verwijlen, T; Leiske, D.; Moldenaers, P; Vermant, J; and Fuller, G.\n\n\n \n\n\n\n Journal of Rheology, 56(5): 1225. 2012.\n \n\n\n\n
\n\n\n\n \n \n \"ExtensionalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{verwijlen_extensional_2012,\n\tAbstract = {Not Available},\n\tAuthor = {Verwijlen, T and Leiske, D.L. and Moldenaers, P and Vermant, J and Fuller, G.G.},\n\tDoi = {10.1122/1.4733717},\n\tJournal = {Journal of Rheology},\n\tLanguage = {English},\n\tNumber = {5},\n\tPages = {1225},\n\tTitle = {Extensional rheometry at interfaces: {Analysis} of the {Cambridge} {Interfacial} {Tensiometer}},\n\tUrl = {http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=2012JRheo..56.1225V&link_type=EJOURNAL},\n\tVolume = {56},\n\tYear = {2012},\n\tBdsk-Url-1 = {http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=2012JRheo..56.1225V&link_type=EJOURNAL},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1122/1.4733717}}\n\n
\n
\n\n\n
\n Not Available\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2011\n \n \n (7)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Interfacial shear rheology of highly confined glassy polymers.\n \n \n \n \n\n\n \n Srivastava, S; Leiske, D; Basu, J K; and Fuller, G.\n\n\n \n\n\n\n Soft Matter, 7(5): 1994–2000. 2011.\n \n\n\n\n
\n\n\n\n \n \n \"InterfacialPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{srivastava_interfacial_2011,\n\tAbstract = {We present results on interfacial shear rheology measurements on Langmuir monolayers of two different polymers, poly(vinyl acetate) and poly(methyl methacrylate) as a function of surface concentration and temperature. While for the high glass transition poly(methyl methacrylate) polymer we find a systematic transition from a viscous dominated regime to an elastic dominated regime as surface concentration is increased, monolayers of the low glass transition polymer, poly(vinyl acetate), remain viscous even at very high surface concentrations. We further interpret the results in terms of the soft glassy rheology model of Sollich et al. [P. Sollich, F. C. Lequeux, P. Hebraud and M. E. Cates, Phys. Rev. Lett., 1997, 78, 2020-2023] and provide evidence of possible reduction in glass transition temperatures in both poly(methyl methacrylate) and poly(vinyl acetate) monolayers due to finite size effects. textcopyright 2011 The Royal Society of Chemistry.},\n\tAuthor = {Srivastava, S and Leiske, D and Basu, J K and Fuller, G.G.},\n\tJournal = {Soft Matter},\n\tNumber = {5},\n\tPages = {1994--2000},\n\tTitle = {Interfacial shear rheology of highly confined glassy polymers},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-79951940549&partnerID=40&md5=ab9998fa27c8a608a5f9f4b331294cd3},\n\tVolume = {7},\n\tYear = {2011},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-79951940549&partnerID=40&md5=ab9998fa27c8a608a5f9f4b331294cd3}}\n\n
\n
\n\n\n
\n We present results on interfacial shear rheology measurements on Langmuir monolayers of two different polymers, poly(vinyl acetate) and poly(methyl methacrylate) as a function of surface concentration and temperature. While for the high glass transition poly(methyl methacrylate) polymer we find a systematic transition from a viscous dominated regime to an elastic dominated regime as surface concentration is increased, monolayers of the low glass transition polymer, poly(vinyl acetate), remain viscous even at very high surface concentrations. We further interpret the results in terms of the soft glassy rheology model of Sollich et al. [P. Sollich, F. C. Lequeux, P. Hebraud and M. E. Cates, Phys. Rev. Lett., 1997, 78, 2020-2023] and provide evidence of possible reduction in glass transition temperatures in both poly(methyl methacrylate) and poly(vinyl acetate) monolayers due to finite size effects. textcopyright 2011 The Royal Society of Chemistry.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Preparation of mineralized nanofibers: Collagen fibrils containing calcium phosphate.\n \n \n \n \n\n\n \n Maas, M; Guo, P; Keeney, M; Yang, F; Hsu, T M; Fuller, G.; Martin, C R; and Zare, R N\n\n\n \n\n\n\n Nano Letters, 11(3): 1383–1388. 2011.\n \n\n\n\n
\n\n\n\n \n \n \"PreparationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{maas_preparation_2011,\n\tAbstract = {We report a straightforward, bottom-up, scalable process for preparing mineralized nanofibers. Our procedure is based on flowing feed solution, containing both inorganic cations and polymeric molecules, through a nanoporous membrane into a receiver solution with anions, which leads to the formation of mineralized nanofibers at the exit of the pores. With this strategy, we were able to achieve size control of the nanofiber diameters. We illustrate this approach by producing collagen fibrils with calcium phosphate incorporated inside the fibrils. This structure, which resembles the basic constituent of bones, assembles itself without the addition of noncollagenous proteins or their polymeric substitutes. Rheological experiments demonstrated that the stiffness of gels derived from these fibrils is enhanced by mineralization. Growth experiments of human adipose derived stem cells on these gels showed the compatibility of the fibrils in a tissue-regeneration context. textcopyright 2011 American Chemical Society.},\n\tAuthor = {Maas, M and Guo, P and Keeney, M and Yang, F and Hsu, T M and Fuller, G.G. and Martin, C R and Zare, R N},\n\tJournal = {Nano Letters},\n\tNumber = {3},\n\tPages = {1383--1388},\n\tTitle = {Preparation of mineralized nanofibers: {Collagen} fibrils containing calcium phosphate},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-79952600465&partnerID=40&md5=707c391e32e7a8f2f48ea7a9ce6a6a9b},\n\tVolume = {11},\n\tYear = {2011},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-79952600465&partnerID=40&md5=707c391e32e7a8f2f48ea7a9ce6a6a9b}}\n\n
\n
\n\n\n
\n We report a straightforward, bottom-up, scalable process for preparing mineralized nanofibers. Our procedure is based on flowing feed solution, containing both inorganic cations and polymeric molecules, through a nanoporous membrane into a receiver solution with anions, which leads to the formation of mineralized nanofibers at the exit of the pores. With this strategy, we were able to achieve size control of the nanofiber diameters. We illustrate this approach by producing collagen fibrils with calcium phosphate incorporated inside the fibrils. This structure, which resembles the basic constituent of bones, assembles itself without the addition of noncollagenous proteins or their polymeric substitutes. Rheological experiments demonstrated that the stiffness of gels derived from these fibrils is enhanced by mineralization. Growth experiments of human adipose derived stem cells on these gels showed the compatibility of the fibrils in a tissue-regeneration context. textcopyright 2011 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Vascular anastomosis using controlled phase transitions in poloxamer gels.\n \n \n \n \n\n\n \n Chang, E I; Galvez, M G; Glotzbach, J P; Hamou, C D; El-Ftesi, S; Rappleye, C T; Sommer, K M; Rajadas, J; Abilez, O J; Fuller, G.; Longaker, M T; and Gurtner, G C\n\n\n \n\n\n\n Nature Medicine, 17(9): 1147–1152. 2011.\n \n\n\n\n
\n\n\n\n \n \n \"VascularPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chang_vascular_2011,\n\tAbstract = {Vascular anastomosis is the cornerstone of vascular, cardiovascular and transplant surgery. Most anastomoses are performed with sutures, which are technically challenging and can lead to failure from intimal hyperplasia and foreign body reaction. Numerous alternatives to sutures have been proposed, but none has proven superior, particularly in small or atherosclerotic vessels. We have developed a new method of sutureless and atraumatic vascular anastomosis that uses US Food and Drug Administration (FDA)-approved thermoreversible tri-block polymers to temporarily maintain an open lumen for precise approximation with commercially available glues. We performed end-to-end anastomoses five times more rapidly than we performed hand-sewn controls, and vessels that were too small ({\\textless}1.0 mm) to sew were successfully reconstructed with this sutureless approach. Imaging of reconstructed rat aorta confirmed equivalent patency, flow and burst strength, and histological analysis demonstrated decreased inflammation and fibrosis at up to 2 years after the procedure. This new technology has potential for improving efficiency and outcomes in the surgical treatment of cardiovascular disease. textcopyright 2011 Nature America, Inc. All rights reserved.},\n\tAuthor = {Chang, E I and Galvez, M G and Glotzbach, J P and Hamou, C D and El-Ftesi, S and Rappleye, C T and Sommer, K M and Rajadas, J and Abilez, O J and Fuller, G.G. and Longaker, M T and Gurtner, G C},\n\tJournal = {Nature Medicine},\n\tNumber = {9},\n\tPages = {1147--1152},\n\tTitle = {Vascular anastomosis using controlled phase transitions in poloxamer gels},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-80052442975&partnerID=40&md5=060d0569ca899858f658dc3b056fa876},\n\tVolume = {17},\n\tYear = {2011},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-80052442975&partnerID=40&md5=060d0569ca899858f658dc3b056fa876}}\n\n
\n
\n\n\n
\n Vascular anastomosis is the cornerstone of vascular, cardiovascular and transplant surgery. Most anastomoses are performed with sutures, which are technically challenging and can lead to failure from intimal hyperplasia and foreign body reaction. Numerous alternatives to sutures have been proposed, but none has proven superior, particularly in small or atherosclerotic vessels. We have developed a new method of sutureless and atraumatic vascular anastomosis that uses US Food and Drug Administration (FDA)-approved thermoreversible tri-block polymers to temporarily maintain an open lumen for precise approximation with commercially available glues. We performed end-to-end anastomoses five times more rapidly than we performed hand-sewn controls, and vessels that were too small (\\textless1.0 mm) to sew were successfully reconstructed with this sutureless approach. Imaging of reconstructed rat aorta confirmed equivalent patency, flow and burst strength, and histological analysis demonstrated decreased inflammation and fibrosis at up to 2 years after the procedure. This new technology has potential for improving efficiency and outcomes in the surgical treatment of cardiovascular disease. textcopyright 2011 Nature America, Inc. All rights reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Editorial: Dynamics and rheology of complex fluid-fluid interfaces.\n \n \n \n \n\n\n \n Fuller, G.; and Vermant, J\n\n\n \n\n\n\n Soft Matter, 7(17): 7583–7585. 2011.\n \n\n\n\n
\n\n\n\n \n \n \"Editorial:Paper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_editorial:_2011,\n\tAuthor = {Fuller, G.G. and Vermant, J},\n\tJournal = {Soft Matter},\n\tNumber = {17},\n\tPages = {7583--7585},\n\tTitle = {Editorial: {Dynamics} and rheology of complex fluid-fluid interfaces},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-82955168870&partnerID=40&md5=9a0b647a89475e590a8300668d1a023d},\n\tVolume = {7},\n\tYear = {2011},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-82955168870&partnerID=40&md5=9a0b647a89475e590a8300668d1a023d}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Designing a tubular matrix of oriented collagen fibrils for tissue engineering.\n \n \n \n \n\n\n \n Lai, E S; Anderson, C M; and Fuller, G.\n\n\n \n\n\n\n Acta Biomaterialia, 7(6): 2448–2456. 2011.\n \n\n\n\n
\n\n\n\n \n \n \"DesigningPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lai_designing_2011,\n\tAbstract = {A scaffold composed entirely of an extracellular matrix component, such as collagen, with cellular level control would be highly desirable for applications in tissue engineering. In this article we introduce a novel, straightforward flow processing technique that fabricates a small diameter tubular matrix constructed of anisotropic collagen fibrils. Scanning electron microscopy confirmed the uniform alignment of the collagen fibrils and subsequent matrix-induced alignment of human fibroblasts. The uniform alignment of the fibroblasts along the collagen fibrils demonstrated the ability of the aligned fibrils to successfully dictate the directional growth of human fibroblasts through contact guidance. Various non-cytotoxic cross-linking techniques were also applied to the collagen conduit to enhance the mechanical properties. Tensile testing and burst pressure were the two measurements performed to characterize the mechanical integrity of the conduit. Mechanical characterization of the cross-linked collagen conduits identified 1-ethyl-3-(3-dimethyl aminopropyl) carbodiimide hydrochloride cross-linking as the most promising technique to reinforce the mechanical properties of native collagen. An oriented conduit of biocompatible material has been fabricated with decent mechanical strength and at a small diameter scale, which is especially applicable in engineering cardiovascular tissues and nerve grafts. textcopyright 2011 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.},\n\tAuthor = {Lai, E S and Anderson, C M and Fuller, G.G.},\n\tJournal = {Acta Biomaterialia},\n\tNumber = {6},\n\tPages = {2448--2456},\n\tTitle = {Designing a tubular matrix of oriented collagen fibrils for tissue engineering},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-79955589110&partnerID=40&md5=2a4d82c7320d6a2812fbdaf49b687679},\n\tVolume = {7},\n\tYear = {2011},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-79955589110&partnerID=40&md5=2a4d82c7320d6a2812fbdaf49b687679}}\n\n
\n
\n\n\n
\n A scaffold composed entirely of an extracellular matrix component, such as collagen, with cellular level control would be highly desirable for applications in tissue engineering. In this article we introduce a novel, straightforward flow processing technique that fabricates a small diameter tubular matrix constructed of anisotropic collagen fibrils. Scanning electron microscopy confirmed the uniform alignment of the collagen fibrils and subsequent matrix-induced alignment of human fibroblasts. The uniform alignment of the fibroblasts along the collagen fibrils demonstrated the ability of the aligned fibrils to successfully dictate the directional growth of human fibroblasts through contact guidance. Various non-cytotoxic cross-linking techniques were also applied to the collagen conduit to enhance the mechanical properties. Tensile testing and burst pressure were the two measurements performed to characterize the mechanical integrity of the conduit. Mechanical characterization of the cross-linked collagen conduits identified 1-ethyl-3-(3-dimethyl aminopropyl) carbodiimide hydrochloride cross-linking as the most promising technique to reinforce the mechanical properties of native collagen. An oriented conduit of biocompatible material has been fabricated with decent mechanical strength and at a small diameter scale, which is especially applicable in engineering cardiovascular tissues and nerve grafts. textcopyright 2011 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Role of fluid elasticity on the dynamics of rinsing flow by an impinging jet.\n \n \n \n \n\n\n \n Hsu, T T; Walker, T W; Frank, C.; and Fuller, G.\n\n\n \n\n\n\n Physics of Fluids, 23(3). 2011.\n \n\n\n\n
\n\n\n\n \n \n \"RolePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hsu_role_2011,\n\tAbstract = {Rinsing flows are common processes where a jet of one liquid impinges upon a layer of a second liquid for the purpose of removing the second liquid. An imaging setup has been developed to obtain both qualitative and quantitative data on the rinsing flow of a jet of water impinging on either layers of Newtonian or elastic fluids. Three classes of test fluids have been investigated: a Newtonian glycerol-water solution, a semidilute aqueous solution of high molecular weight polyacrylamide solution displaying both elasticity and shear thinning, and an elastic but non-shear thinning Boger fluid. The fluids were designed to have approximately equal zero-shear viscosities. For all cases, a circular hydraulic jump occurs and Saffman-Taylor instabilities were observed at the interface between the low viscosity jet and the higher viscosity coating liquids. Results show that the elasticity (extensional viscosity) of the samples influences the pattern of the instabilities and contributes to dampening surface disturbances in the vicinity of the hydraulic jump. Quantitative measurements of liquid layer thicknesses were obtained using a laser triangulation technique. We observed that shear thinning contributes to increasing the velocity of the hydraulic jump circle growth, and the growth profile appears to be linear instead of logarithmic-like as in the Newtonian fluids. Shear thinning characteristics of the samples also contribute to a larger vertical height of the hydraulic jump and an undercutting phenomenon. The elasticity of the fluids contributes to a "recoil" of the hydraulic jump circle, causing the circle, after initial expansion, to shrink in size before expanding again. textcopyright 2011 American Institute of Physics.},\n\tAuthor = {Hsu, T T and Walker, T W and Frank, C.W. and Fuller, G.G.},\n\tJournal = {Physics of Fluids},\n\tNumber = {3},\n\tTitle = {Role of fluid elasticity on the dynamics of rinsing flow by an impinging jet},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-79953272487&partnerID=40&md5=59a6dd56a3416e2e57017633b5019b59},\n\tVolume = {23},\n\tYear = {2011},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-79953272487&partnerID=40&md5=59a6dd56a3416e2e57017633b5019b59}}\n\n
\n
\n\n\n
\n Rinsing flows are common processes where a jet of one liquid impinges upon a layer of a second liquid for the purpose of removing the second liquid. An imaging setup has been developed to obtain both qualitative and quantitative data on the rinsing flow of a jet of water impinging on either layers of Newtonian or elastic fluids. Three classes of test fluids have been investigated: a Newtonian glycerol-water solution, a semidilute aqueous solution of high molecular weight polyacrylamide solution displaying both elasticity and shear thinning, and an elastic but non-shear thinning Boger fluid. The fluids were designed to have approximately equal zero-shear viscosities. For all cases, a circular hydraulic jump occurs and Saffman-Taylor instabilities were observed at the interface between the low viscosity jet and the higher viscosity coating liquids. Results show that the elasticity (extensional viscosity) of the samples influences the pattern of the instabilities and contributes to dampening surface disturbances in the vicinity of the hydraulic jump. Quantitative measurements of liquid layer thicknesses were obtained using a laser triangulation technique. We observed that shear thinning contributes to increasing the velocity of the hydraulic jump circle growth, and the growth profile appears to be linear instead of logarithmic-like as in the Newtonian fluids. Shear thinning characteristics of the samples also contribute to a larger vertical height of the hydraulic jump and an undercutting phenomenon. The elasticity of the fluids contributes to a \"recoil\" of the hydraulic jump circle, causing the circle, after initial expansion, to shrink in size before expanding again. textcopyright 2011 American Institute of Physics.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Insertion mechanism of a poly(ethylene oxide)-poly(butylene oxide) block copolymer into a DPPC monolayer.\n \n \n \n \n\n\n \n Leiske, D. L; Meckes, B.; Miller, C. E; Wu, C.; Walker, T. W; Lin, B.; Meron, M.; Ketelson, H. A; Toney, M. F; and Fuller, G. G\n\n\n \n\n\n\n Langmuir, 27(18): 11444–11450. September 2011.\n \n\n\n\n
\n\n\n\n \n \n \"InsertionPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{leiske_insertion_2011,\n\tAbstract = {Interactions between amphiphilic block copolymers and lipids are of medical interest for applications such as drug delivery and the restoration of damaged cell membranes. A series of monodisperse poly(ethylene oxide)-poly(butylene oxide) (EOBO) block copolymers were obtained with two ratios of hydrophilic/hydrophobic block lengths. We have explored the surface activity of EOBO at a clean interface and under 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) monolayers as a simple cell membrane model. At the same subphase concentration, EOBO achieved higher equilibrium surface pressures under DPPC compared to a bare interface, and the surface activity was improved with longer poly(butylene oxide) blocks. Further investigation of the DPPC/EOBO monolayers showed that combined films exhibited similar surface rheology compared to pure DPPC at the same surface pressures. DPPC/EOBO phase separation was observed in fluorescently doped monolayers, and within the liquid-expanded liquid-condensed coexistence region for DPPC, EOBO did not drastically alter the liquid-condensed domain shapes. Grazing incidence X-ray diffraction (GIXD) and X-ray reflectivity (XRR) quantitatively confirmed that the lattice spacings and tilt of DPPC in lipid-rich regions of the monolayer were nearly equivalent to those of a pure DPPC monolayer at the same surface pressures.},\n\tAuthor = {Leiske, Danielle L and Meckes, Brian and Miller, Chad E and Wu, Cynthia and Walker, Travis W and Lin, Binhua and Meron, Mati and Ketelson, Howard A and Toney, Michael F and Fuller, Gerald G},\n\tDoi = {10.1021/la2016879},\n\tJournal = {Langmuir},\n\tLanguage = {English},\n\tMonth = sep,\n\tNumber = {18},\n\tPages = {11444--11450},\n\tPmid = {21834565},\n\tTitle = {Insertion mechanism of a poly(ethylene oxide)-poly(butylene oxide) block copolymer into a {DPPC} monolayer.},\n\tUrl = {http://eutils.ncbi.nlm.nih.gov/entrez/eutils/elink.fcgi?dbfrom=pubmed&id=21834565&retmode=ref&cmd=prlinks},\n\tVolume = {27},\n\tYear = {2011},\n\tBdsk-Url-1 = {http://eutils.ncbi.nlm.nih.gov/entrez/eutils/elink.fcgi?dbfrom=pubmed&id=21834565&retmode=ref&cmd=prlinks},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/la2016879}}\n\n
\n
\n\n\n
\n Interactions between amphiphilic block copolymers and lipids are of medical interest for applications such as drug delivery and the restoration of damaged cell membranes. A series of monodisperse poly(ethylene oxide)-poly(butylene oxide) (EOBO) block copolymers were obtained with two ratios of hydrophilic/hydrophobic block lengths. We have explored the surface activity of EOBO at a clean interface and under 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) monolayers as a simple cell membrane model. At the same subphase concentration, EOBO achieved higher equilibrium surface pressures under DPPC compared to a bare interface, and the surface activity was improved with longer poly(butylene oxide) blocks. Further investigation of the DPPC/EOBO monolayers showed that combined films exhibited similar surface rheology compared to pure DPPC at the same surface pressures. DPPC/EOBO phase separation was observed in fluorescently doped monolayers, and within the liquid-expanded liquid-condensed coexistence region for DPPC, EOBO did not drastically alter the liquid-condensed domain shapes. Grazing incidence X-ray diffraction (GIXD) and X-ray reflectivity (XRR) quantitatively confirmed that the lattice spacings and tilt of DPPC in lipid-rich regions of the monolayer were nearly equivalent to those of a pure DPPC monolayer at the same surface pressures.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2010\n \n \n (5)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n A double wall-ring geometry for interfacial shear rheometry.\n \n \n \n\n\n \n Vandebril, S.; Franck, A.; Fuller, G.; Moldenaers, P.; and Vermant, J.\n\n\n \n\n\n\n Rheologica Acta, 49(2): 131–144. February 2010.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{vandebril_double_2010,\n\tAuthor = {Vandebril, Steven and Franck, Aly and Fuller, Gerald and Moldenaers, Paula and Vermant, Jan},\n\tJournal = {Rheologica Acta},\n\tMonth = feb,\n\tNumber = {2},\n\tPages = {131--144},\n\tTitle = {A double wall-ring geometry for interfacial shear rheometry},\n\tVolume = {49},\n\tYear = {2010}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The interfacial viscoelastic properties and structures of human and animal Meibomian lipids.\n \n \n \n \n\n\n \n Leiske, D. L; Raju, S. R; Ketelson, H. A; Millar, T. J; and Fuller, G. G\n\n\n \n\n\n\n Experimental Eye Research, 90(5): 598–604. May 2010.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{leiske_interfacial_2010,\n\tAuthor = {Leiske, Danielle L and Raju, Shiwani R and Ketelson, Howard A and Millar, Thomas J and Fuller, Gerald G},\n\tDoi = {10.1016/j.exer.2010.02.004},\n\tJournal = {Experimental Eye Research},\n\tLanguage = {English},\n\tMonth = may,\n\tNumber = {5},\n\tPages = {598--604},\n\tTitle = {The interfacial viscoelastic properties and structures of human and animal {Meibomian} lipids},\n\tUrl = {http://linkinghub.elsevier.com/retrieve/pii/S0014483510000497},\n\tVolume = {90},\n\tYear = {2010},\n\tBdsk-Url-1 = {http://linkinghub.elsevier.com/retrieve/pii/S0014483510000497},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1016/j.exer.2010.02.004}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Charge interaction between particle-laden fluid interfaces.\n \n \n \n\n\n \n Xu, H; Kirkwood, J; Lask, M; and Fuller, G\n\n\n \n\n\n\n Langmuir, 26(5): 3160–3164. 2010.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{xu_charge_2010,\n\tAuthor = {Xu, H and Kirkwood, J and Lask, M and Fuller, G},\n\tJournal = {Langmuir},\n\tNumber = {5},\n\tPages = {3160--3164},\n\tTitle = {Charge interaction between particle-laden fluid interfaces},\n\tVolume = {26},\n\tYear = {2010}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Interfacial flow processing of collagen.\n \n \n \n \n\n\n \n Goffin, A J; Rajadas, J; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 26(5): 3514–3521. 2010.\n \n\n\n\n
\n\n\n\n \n \n \"InterfacialPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{goffin_interfacial_2010,\n\tAbstract = {A new method for creating substrates made out of ordered collagen fibers, on which cells in culture can align, is proposed. The substrates can be used for research in cell culture, and this research presents a significant advance in the technology to coat implants in order to improve cell adhesion. In the procedure presented here, a molecular solution of collagen is spread at the interface of a saline solution and air to induce fiber formation, compressed at a high speed to induce orientation and deposited on solid substrates via Langmuir-Blodgett transfer. Several interfacial techniques are employed to investigate the behavior of collagen, which is shown to be dependent on the salt concentration of the subphase as well as the temperature. After Langmuir-Blodgett transfer, primary human fibroblasts and adipose-derived stem cells are cultured on the collagen substrates. Both types of cells respond favorably to the collagen orientation and align with the deposited fibers. The technique presented here provides a simple method to produce well-controlled, oriented collagen substrates that can be used in tissue culture research or scaffolding applications without the use of additives and/or bioincompatible materials.},\n\tAuthor = {Goffin, A J and Rajadas, J and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {5},\n\tPages = {3514--3521},\n\tTitle = {Interfacial flow processing of collagen.},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-77952788615&partnerID=40&md5=1e78e2382dfa2f16c6f410adc98462a6},\n\tVolume = {26},\n\tYear = {2010},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-77952788615&partnerID=40&md5=1e78e2382dfa2f16c6f410adc98462a6}}\n\n
\n
\n\n\n
\n A new method for creating substrates made out of ordered collagen fibers, on which cells in culture can align, is proposed. The substrates can be used for research in cell culture, and this research presents a significant advance in the technology to coat implants in order to improve cell adhesion. In the procedure presented here, a molecular solution of collagen is spread at the interface of a saline solution and air to induce fiber formation, compressed at a high speed to induce orientation and deposited on solid substrates via Langmuir-Blodgett transfer. Several interfacial techniques are employed to investigate the behavior of collagen, which is shown to be dependent on the salt concentration of the subphase as well as the temperature. After Langmuir-Blodgett transfer, primary human fibroblasts and adipose-derived stem cells are cultured on the collagen substrates. Both types of cells respond favorably to the collagen orientation and align with the deposited fibers. The technique presented here provides a simple method to produce well-controlled, oriented collagen substrates that can be used in tissue culture research or scaffolding applications without the use of additives and/or bioincompatible materials.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Thin film formation of silica nanoparticle/lipid composite films at the fluid-fluid interface.\n \n \n \n \n\n\n \n Maas, M.; Ooi, C. C; and Fuller, G. G\n\n\n \n\n\n\n Langmuir, 26(23): 17867–17873. 2010.\n \n\n\n\n
\n\n\n\n \n \n \"ThinPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{maas_thin_2010,\n\tAbstract = {We report a new and simple method for the formation of thin films at the interface between aqueous silica Ludox dispersions and lipid solutions in decane. The lipids used are stearic acid, stearyl amine, and stearyl alcohol alongside silica Ludox nanoparticle dispersions of varying pH. At basic pH thin films consisting of a mixture of stearic acid and silica nanoparticles precipitate at the interface. At acidic and neutral pH we were able to produce thin films consisting of stearyl amine and silica particles. The film growth was studied in situ with interfacial shear rheology. In addition to that, surface pressure isotherm and dynamic light scattering experiments were performed. The films all exhibit strong dynamic rheological moduli, rendering them an interesting material for applications such as capsule formation, surface coating, or as functional membranes.},\n\tAuthor = {Maas, Michael and Ooi, Chin C and Fuller, Gerald G},\n\tDoi = {10.1021/la103492a},\n\tJournal = {Langmuir},\n\tNumber = {23},\n\tPages = {17867--17873},\n\tPmid = {21067193},\n\tTitle = {Thin film formation of silica nanoparticle/lipid composite films at the fluid-fluid interface.},\n\tUrl = {http://eutils.ncbi.nlm.nih.gov/entrez/eutils/elink.fcgi?dbfrom=pubmed&id=21067193&retmode=ref&cmd=prlinks},\n\tVolume = {26},\n\tYear = {2010},\n\tBdsk-Url-1 = {http://eutils.ncbi.nlm.nih.gov/entrez/eutils/elink.fcgi?dbfrom=pubmed&id=21067193&retmode=ref&cmd=prlinks},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/la103492a}}\n\n
\n
\n\n\n
\n We report a new and simple method for the formation of thin films at the interface between aqueous silica Ludox dispersions and lipid solutions in decane. The lipids used are stearic acid, stearyl amine, and stearyl alcohol alongside silica Ludox nanoparticle dispersions of varying pH. At basic pH thin films consisting of a mixture of stearic acid and silica nanoparticles precipitate at the interface. At acidic and neutral pH we were able to produce thin films consisting of stearyl amine and silica particles. The film growth was studied in situ with interfacial shear rheology. In addition to that, surface pressure isotherm and dynamic light scattering experiments were performed. The films all exhibit strong dynamic rheological moduli, rendering them an interesting material for applications such as capsule formation, surface coating, or as functional membranes.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2009\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Isovaleric, methylmalonic, and propionic acid decrease anesthetic ec50 in tadpoles, modulate glycine receptor function, and interact with the lipid 1,2-dipalmitoyl-Sn-glycero-3-phosphocholine.\n \n \n \n \n\n\n \n Weng, Y; Hsu, T T; Zhao, J; Nishimura, S; Fuller, G.; and Sonner, J M\n\n\n \n\n\n\n Anesthesia and Analgesia, 108(5): 1538–1545. 2009.\n \n\n\n\n
\n\n\n\n \n \n \"Isovaleric,Paper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{weng_isovaleric_2009,\n\tAbstract = {Introduction: Elevated concentrations of isovaleric (IVA), methylmalonic (MMA), and propionic acid are associated with impaired consciousness in genetic diseases (organic acidemias). We conjectured that part of the central nervous system depression observed in these disorders was due to anesthetic effects of these metabolites. We tested three hypotheses. First, that these metabolites would have anesthetic-sparing effects, possibly being anesthetics by themselves. Second, that these compounds would modulate glycine and γ-aminobutyric acid (GABAA) receptor function, increasing chloride currents through these channels as potent clinical inhaled anesthetics do. Third, that these compounds would affect physical properties of lipids. Methods: Anesthetic EC50s were measured in Xenopus laevis tadpoles. Glycine and GABAA receptors were expressed in Xenopus laevis oocytes and studied using two-electrode voltage clamping. Pressure-area isotherms of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) monolayers were measured with and without added organic acids. Results: IVA acid was an anesthetic in tadpoles, whereas MMA and propionic acid decreased isoflurane's EC50 by half. All three organic acids concentration-dependently increased current through α1 glycine receptors. There were minimal effects on α1β2γ2s GABAA receptors. The organic acids increased total lateral pressure (surface pressure) of DPPC monolayers, including at mean molecular areas typical of bilayers. Conclusion: IVA, MMA, and propionic acid have anesthetic effects in tadpoles, positively modulate glycine receptor function and affect physical properties of DPPC monolayers. textcopyright 2009 International Anesthesia Research Society.},\n\tAuthor = {Weng, Y and Hsu, T T and Zhao, J and Nishimura, S and Fuller, G.G. and Sonner, J M},\n\tJournal = {Anesthesia and Analgesia},\n\tNumber = {5},\n\tPages = {1538--1545},\n\tTitle = {Isovaleric, methylmalonic, and propionic acid decrease anesthetic ec50 in tadpoles, modulate glycine receptor function, and interact with the lipid 1,2-dipalmitoyl-{Sn}-glycero-3-phosphocholine},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-65349147704&partnerID=40&md5=2acd89da62269460347894a72123469c},\n\tVolume = {108},\n\tYear = {2009},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-65349147704&partnerID=40&md5=2acd89da62269460347894a72123469c}}\n\n
\n
\n\n\n
\n Introduction: Elevated concentrations of isovaleric (IVA), methylmalonic (MMA), and propionic acid are associated with impaired consciousness in genetic diseases (organic acidemias). We conjectured that part of the central nervous system depression observed in these disorders was due to anesthetic effects of these metabolites. We tested three hypotheses. First, that these metabolites would have anesthetic-sparing effects, possibly being anesthetics by themselves. Second, that these compounds would modulate glycine and γ-aminobutyric acid (GABAA) receptor function, increasing chloride currents through these channels as potent clinical inhaled anesthetics do. Third, that these compounds would affect physical properties of lipids. Methods: Anesthetic EC50s were measured in Xenopus laevis tadpoles. Glycine and GABAA receptors were expressed in Xenopus laevis oocytes and studied using two-electrode voltage clamping. Pressure-area isotherms of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) monolayers were measured with and without added organic acids. Results: IVA acid was an anesthetic in tadpoles, whereas MMA and propionic acid decreased isoflurane's EC50 by half. All three organic acids concentration-dependently increased current through α1 glycine receptors. There were minimal effects on α1β2γ2s GABAA receptors. The organic acids increased total lateral pressure (surface pressure) of DPPC monolayers, including at mean molecular areas typical of bilayers. Conclusion: IVA, MMA, and propionic acid have anesthetic effects in tadpoles, positively modulate glycine receptor function and affect physical properties of DPPC monolayers. textcopyright 2009 International Anesthesia Research Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Surface rheology of a polymer monolayer: Effects of polymer chain length and compression rate.\n \n \n \n \n\n\n \n Spigone, E.; Cho, G. Y; Fuller, G. G; and Cicuta, P.\n\n\n \n\n\n\n Langmuir, 25(13): 7457–7464. July 2009.\n \n\n\n\n
\n\n\n\n \n \n \"SurfacePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{spigone_surface_2009,\n\tAbstract = {We study surface layers of a simple homopolymer poly(vinyl acetate) on the air-water interface as a function of the concentration and the polymer molecular weight. Our results suggest that there is an effect of the compression rate on both the structure of the layers and their rheological behavior, while the length of the chain influences only the rheology. At very low compression speeds, the surface layer of short chains does not exhibit the classical semi-dilute regime behavior, forming instead a solid phase. For fluid layers, we report on the dependence of surface viscosity upon the concentration, showing a first crossover, which happens close to the semi-dilute-concentrated regime boundary, from a scaling behavior with the concentration to an Eyring-like liquid. A second rheological transition happens at very high concentrations, near close packing, where the Newtonian liquid phase gives way to a soft solid phase. textcopyright 2009 American Chemical Society.},\n\tAuthor = {Spigone, Elisabetta and Cho, Gil-Young Y and Fuller, Gerald G and Cicuta, Pietro},\n\tDoi = {10.1021/la900385y},\n\tJournal = {Langmuir},\n\tMonth = jul,\n\tNumber = {13},\n\tPages = {7457--7464},\n\tPmid = {19374337},\n\tTitle = {Surface rheology of a polymer monolayer: {Effects} of polymer chain length and compression rate},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-67650090012&partnerID=40&md5=572302c42ee030e9828e80cea8fdda70 http://dx.doi.org/10.1021/la900385y},\n\tVolume = {25},\n\tYear = {2009},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-67650090012&partnerID=40&md5=572302c42ee030e9828e80cea8fdda70%20http://dx.doi.org/10.1021/la900385y},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/la900385y}}\n\n
\n
\n\n\n
\n We study surface layers of a simple homopolymer poly(vinyl acetate) on the air-water interface as a function of the concentration and the polymer molecular weight. Our results suggest that there is an effect of the compression rate on both the structure of the layers and their rheological behavior, while the length of the chain influences only the rheology. At very low compression speeds, the surface layer of short chains does not exhibit the classical semi-dilute regime behavior, forming instead a solid phase. For fluid layers, we report on the dependence of surface viscosity upon the concentration, showing a first crossover, which happens close to the semi-dilute-concentrated regime boundary, from a scaling behavior with the concentration to an Eyring-like liquid. A second rheological transition happens at very high concentrations, near close packing, where the Newtonian liquid phase gives way to a soft solid phase. textcopyright 2009 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Liquid crystalline collagen: A self-assembled morphology for the orientation of mammalian cells.\n \n \n \n \n\n\n \n Kirkwood, J E; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 25(5): 3200–3206. 2009.\n \n\n\n\n
\n\n\n\n \n \n \"LiquidPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kirkwood_liquid_2009,\n\tAbstract = {We report the creation of collagen films having a cholesteric banding structure with an orientation that can be systematically controlled. The action of hydrodynamic flow and rapid desiccation was used to influence the orientation of collagen fibrils, producing a film with a twisted plywood architecture. Adult human fibroblasts cultured on these substrates orient in the direction of the flow deposition, and filopodia are extended onto individual bands. Atomic force microscopy reveals the assembly of 30 nm collagen fibrils into the uniform cholesteric collagen films with a periodic surface relief. The generation of collagen with a reticular, "basket-weave" morphology when using lower concentrations is also discussed. textcopyright Copyright 2009 American Chemical Society.},\n\tAuthor = {Kirkwood, J E and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {5},\n\tPages = {3200--3206},\n\tTitle = {Liquid crystalline collagen: {A} self-assembled morphology for the orientation of mammalian cells},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-65249133911&partnerID=40&md5=6ffc8877d728237be96d35aaf9bd0b39},\n\tVolume = {25},\n\tYear = {2009},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-65249133911&partnerID=40&md5=6ffc8877d728237be96d35aaf9bd0b39}}\n\n
\n
\n\n\n
\n We report the creation of collagen films having a cholesteric banding structure with an orientation that can be systematically controlled. The action of hydrodynamic flow and rapid desiccation was used to influence the orientation of collagen fibrils, producing a film with a twisted plywood architecture. Adult human fibroblasts cultured on these substrates orient in the direction of the flow deposition, and filopodia are extended onto individual bands. Atomic force microscopy reveals the assembly of 30 nm collagen fibrils into the uniform cholesteric collagen films with a periodic surface relief. The generation of collagen with a reticular, \"basket-weave\" morphology when using lower concentrations is also discussed. textcopyright Copyright 2009 American Chemical Society.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2008\n \n \n (5)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Langmuir monolayers of straight-chain and branched hexadecanol and eicosanol mixtures.\n \n \n \n\n\n \n Kurtz, R E; Toney, M F; Pople, J A; Lin, B; Meron, M; Majewski, J; Lange, A; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 24(24): 14005–14014. 2008.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kurtz_langmuir_2008,\n\tAuthor = {Kurtz, R E and Toney, M F and Pople, J A and Lin, B and Meron, M and Majewski, J and Lange, A and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {24},\n\tPages = {14005--14014},\n\tTitle = {Langmuir monolayers of straight-chain and branched hexadecanol and eicosanol mixtures},\n\tVolume = {24},\n\tYear = {2008}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Surface rheology of hydrophobically-modified peg polymers associating with a phospholipid monolayer at the air-water interface.\n \n \n \n \n\n\n \n Auguste, D T; Kirkwood, J E; Kohn, J; Fuller, G; and Prud'Homme, R K\n\n\n \n\n\n\n In AIChE Annu. Meet. Conf. Proc., Philadelphia, PA, 2008. Chemical Engineering, Princeton University, A301 E-Quad, Princeton, NJ 08544, United States\n \n\n\n\n
\n\n\n\n \n \n \"SurfacePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@inproceedings{auguste_surface_2008,\n\tAddress = {Philadelphia, PA},\n\tAuthor = {Auguste, D T and Kirkwood, J E and Kohn, J and Fuller, G and Prud'Homme, R K},\n\tBooktitle = {{AIChE} {Annu}. {Meet}. {Conf}. {Proc}.},\n\tIsbn = {9780816910502},\n\tPublisher = {Chemical Engineering, Princeton University, A301 E-Quad, Princeton, NJ 08544, United States},\n\tTitle = {Surface rheology of hydrophobically-modified peg polymers associating with a phospholipid monolayer at the air-water interface},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-79952304895&partnerID=40&md5=13f9bd8db778bf02f7cbd1a82dee8bdf},\n\tYear = {2008},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-79952304895&partnerID=40&md5=13f9bd8db778bf02f7cbd1a82dee8bdf}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Interaction of human whole saliva and astringent dietary compounds investigated by interfacial shear rheology.\n \n \n \n\n\n \n Rossetti, D; Yakubov, G E; Stokes, J R; Williamson, A M; and Fuller, G.\n\n\n \n\n\n\n Food Hydrocolloids, 22(6): 1068–1078. 2008.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{rossetti_interaction_2008,\n\tAuthor = {Rossetti, D and Yakubov, G E and Stokes, J R and Williamson, A M and Fuller, G.G.},\n\tJournal = {Food Hydrocolloids},\n\tNumber = {6},\n\tPages = {1068--1078},\n\tTitle = {Interaction of human whole saliva and astringent dietary compounds investigated by interfacial shear rheology},\n\tVolume = {22},\n\tYear = {2008}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Small molecule, non-peptide p75NTR ligands inhibit Aβ-induced neurodegeneration and synaptic impairment.\n \n \n \n \n\n\n \n Yang, T; Knowles, J K; Lu, Q; Zhang, H; Arancio, O; Moore, L A; Chang, T; Wang, Q; Andreasson, K; Rajadas, J; Fuller, G.; Xie, Y; Massa, S M; and Longo, F M\n\n\n \n\n\n\n PLoS ONE, 3(11). 2008.\n \n\n\n\n
\n\n\n\n \n \n \"SmallPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{yang_small_2008,\n\tAbstract = {The p75 neurotrophin receptor (p75NTR) is expressed by neurons particularly vulnerable in Alzheimer's disease (AD). We tested the hypothesis that non-peptide, small molecule p75NTR ligands found to promote survival signaling might prevent Aβ-induced degeneration and synaptic dysfunction. These ligands inhibited Aβ-induced neuritic dystrophy, death of cultured neurons and Aβ-induced death of pyramidal neurons in hippocampal slice cultures. Moreover, ligands inhibited Aβ-induced activation of molecules involved in AD pathology including calpain/cdk5, GSK3β and c-Jun, and tau phosphorylation, and prevented Aβ-induced inactivation of AKT and CREB. Finally, a p75NTR ligand blocked Aβ-induced hippocampal LTP impairment. These studies support an extensive intersection between p75NTR signaling and Aβ pathogenic mechanisms, and introduce a class of specific small molecule ligands with the unique ability to block multiple fundamental AD-related signaling pathways, reverse synaptic impairment and inhibit Aβ-induced neuronal dystrophy and death.},\n\tAuthor = {Yang, T and Knowles, J K and Lu, Q and Zhang, H and Arancio, O and Moore, L A and Chang, T and Wang, Q and Andreasson, K and Rajadas, J and Fuller, G.G. and Xie, Y and Massa, S M and Longo, F M},\n\tJournal = {PLoS ONE},\n\tNumber = {11},\n\tTitle = {Small molecule, non-peptide p75NTR ligands inhibit {Aβ}-induced neurodegeneration and synaptic impairment},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-56149091270&partnerID=40&md5=02b623f6b9db868ddf40a767f45365dd},\n\tVolume = {3},\n\tYear = {2008},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-56149091270&partnerID=40&md5=02b623f6b9db868ddf40a767f45365dd}}\n\n
\n
\n\n\n
\n The p75 neurotrophin receptor (p75NTR) is expressed by neurons particularly vulnerable in Alzheimer's disease (AD). We tested the hypothesis that non-peptide, small molecule p75NTR ligands found to promote survival signaling might prevent Aβ-induced degeneration and synaptic dysfunction. These ligands inhibited Aβ-induced neuritic dystrophy, death of cultured neurons and Aβ-induced death of pyramidal neurons in hippocampal slice cultures. Moreover, ligands inhibited Aβ-induced activation of molecules involved in AD pathology including calpain/cdk5, GSK3β and c-Jun, and tau phosphorylation, and prevented Aβ-induced inactivation of AKT and CREB. Finally, a p75NTR ligand blocked Aβ-induced hippocampal LTP impairment. These studies support an extensive intersection between p75NTR signaling and Aβ pathogenic mechanisms, and introduce a class of specific small molecule ligands with the unique ability to block multiple fundamental AD-related signaling pathways, reverse synaptic impairment and inhibit Aβ-induced neuronal dystrophy and death.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Analysis of the magnetic rod interfacial stress rheometer.\n \n \n \n \n\n\n \n Reynaert, S; Brooks, C.; Moldenaers, P; Vermant, J; and Fuller, G.\n\n\n \n\n\n\n , 52(1): 261–285. 2008.\n \n\n\n\n
\n\n\n\n \n \n \"AnalysisPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{reynaert_analysis_2008,\n\tAuthor = {Reynaert, S and Brooks, C.F. and Moldenaers, P and Vermant, J and Fuller, G.G.},\n\tDoi = {10.1122/1.2798238},\n\tNumber = {1},\n\tPages = {261--285},\n\tTitle = {Analysis of the magnetic rod interfacial stress rheometer},\n\tUrl = {http://link.aip.org/link/?JORHD2/52/261/1},\n\tVolume = {52},\n\tYear = {2008},\n\tBdsk-Url-1 = {http://link.aip.org/link/?JORHD2/52/261/1},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1122/1.2798238}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2007\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Why inhaling salt water changes what we exhale.\n \n \n \n \n\n\n \n Watanabe, W; Thomas, M; Clarke, R; Klibanov, A M; Langer, R; Katstra, J; Fuller, G.; Griel, L C; Fiegel, J; and Edwards, D\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 307(1): 71–78. 2007.\n \n\n\n\n
\n\n\n\n \n \n \"WhyPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{watanabe_why_2007,\n\tAbstract = {We find that inhaling salt water diminishes subsequently exhaled biomaterial in man and animals due to reversible stabilization of the airway lining fluid (ALF)/air interface as a novel potential means for control of the spread of airborne infectious disease. The mechanism of this phenomenon relates to charge shielding of mucin or mucin-like macromolecules that consequently undergo gelation; this gelation alters the physical properties of the ALF surface and reduces its breakup. Cations in the nebulized solution and apparent surface viscoelasticity of the ALF (more than any other ALF intrinsic physical property) appear to be responsible for the reduced tendency of the ALF to disintegrate into very small droplets. We confirm these effects in vivo and show their reversibility through nebulization of saline solutions to anesthetized bull calves. textcopyright 2006 Elsevier Inc. All rights reserved.},\n\tAuthor = {Watanabe, W and Thomas, M and Clarke, R and Klibanov, A M and Langer, R and Katstra, J and Fuller, G.G. and Griel, L C and Fiegel, J and Edwards, D},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {1},\n\tPages = {71--78},\n\tTitle = {Why inhaling salt water changes what we exhale},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-33847039206&partnerID=40&md5=1fed350270c725286102ac7ed6b3cb21},\n\tVolume = {307},\n\tYear = {2007},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-33847039206&partnerID=40&md5=1fed350270c725286102ac7ed6b3cb21}}\n\n
\n
\n\n\n
\n We find that inhaling salt water diminishes subsequently exhaled biomaterial in man and animals due to reversible stabilization of the airway lining fluid (ALF)/air interface as a novel potential means for control of the spread of airborne infectious disease. The mechanism of this phenomenon relates to charge shielding of mucin or mucin-like macromolecules that consequently undergo gelation; this gelation alters the physical properties of the ALF surface and reduces its breakup. Cations in the nebulized solution and apparent surface viscoelasticity of the ALF (more than any other ALF intrinsic physical property) appear to be responsible for the reduced tendency of the ALF to disintegrate into very small droplets. We confirm these effects in vivo and show their reversibility through nebulization of saline solutions to anesthetized bull calves. textcopyright 2006 Elsevier Inc. All rights reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Particle bridging between oil and water interfaces.\n \n \n \n\n\n \n Xu, H; Lask, M; Kirkwood, J; and Fuller, G\n\n\n \n\n\n\n Langmuir, 23(9): 4837–4841. 2007.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{xu_particle_2007,\n\tAuthor = {Xu, H and Lask, M and Kirkwood, J and Fuller, G},\n\tJournal = {Langmuir},\n\tNumber = {9},\n\tPages = {4837--4841},\n\tTitle = {Particle bridging between oil and water interfaces},\n\tVolume = {23},\n\tYear = {2007}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Determining the mechanical response of particle-laden fluid interfaces using surface pressure isotherms and bulk pressure measurements of droplets.\n \n \n \n \n\n\n \n Monteux, C.; Kirkwood, J; Xu, H; Jung, E; and Fuller, G.\n\n\n \n\n\n\n Physical Chemistry Chemical Physics, 9(48): 6344–6350. 2007.\n \n\n\n\n
\n\n\n\n \n \n \"DeterminingPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{monteux_determining_2007,\n\tAbstract = {The mechanical response of particle-laden fluid interfaces is determined by measuring the internal pressures of particle-coated drops as a function of the drop volume. The particle monolayers undergoing compression-expansion cycles exhibit three distinct states: fluid state, jammed state, and buckled state. The P-V curves are compared to the surface pressure isotherms Π-A that are measured using a Langmuir trough and a Wilhelmy plate on a flat water-decane interface covered with the same particles. We find that in the fluid and jammed states, the water drop in decane can be described by the Young-Laplace equation. Therefore in these relatively low compression states, the bulk pressure measurements can be used to deduce the interfacial tension of the droplets and yield similar surface pressure isotherms to the ones measured with the Wilhelmy plate. In the buckled state, the internal pressure of the drop yields a zero value, which is consistent with the zero interfacial tension measured with the Wilhelmy plate. Moreover we find that the compressibility in the jammed state does not depend on the particle size. textcopyright the Owner Societies.},\n\tAuthor = {Monteux, C. and Kirkwood, J and Xu, H and Jung, E and Fuller, G.G.},\n\tJournal = {Physical Chemistry Chemical Physics},\n\tNumber = {48},\n\tPages = {6344--6350},\n\tTitle = {Determining the mechanical response of particle-laden fluid interfaces using surface pressure isotherms and bulk pressure measurements of droplets},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-36949025546&partnerID=40&md5=118e427c87e716c229a84d9cb6a1171d},\n\tVolume = {9},\n\tYear = {2007},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-36949025546&partnerID=40&md5=118e427c87e716c229a84d9cb6a1171d}}\n\n
\n
\n\n\n
\n The mechanical response of particle-laden fluid interfaces is determined by measuring the internal pressures of particle-coated drops as a function of the drop volume. The particle monolayers undergoing compression-expansion cycles exhibit three distinct states: fluid state, jammed state, and buckled state. The P-V curves are compared to the surface pressure isotherms Π-A that are measured using a Langmuir trough and a Wilhelmy plate on a flat water-decane interface covered with the same particles. We find that in the fluid and jammed states, the water drop in decane can be described by the Young-Laplace equation. Therefore in these relatively low compression states, the bulk pressure measurements can be used to deduce the interfacial tension of the droplets and yield similar surface pressure isotherms to the ones measured with the Wilhelmy plate. In the buckled state, the internal pressure of the drop yields a zero value, which is consistent with the zero interfacial tension measured with the Wilhelmy plate. Moreover we find that the compressibility in the jammed state does not depend on the particle size. textcopyright the Owner Societies.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Mechanical properties and structure of particle coated interfaces: Influence of particle size and bidisperse 2D suspensions.\n \n \n \n \n\n\n \n Monteux, C.; Jung, E; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 23(7): 3975–3980. 2007.\n \n\n\n\n
\n\n\n\n \n \n \"MechanicalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{monteux_mechanical_2007,\n\tAbstract = {We report surface pressure-area (prod-A) isotherms of bidisperse mixtures of anionic polystyrene latex particles at a water/n-decane interface as well as optical photographs of the interface for various compressions and mixture ratios. In the case of mixtures of 3 and 5 μm particles, we observe crystalline layers at high or low concentration ratios, where the "impurity" particles concentrate at the grain boundaries of the crystalline structure. At intermediate ratios, the layers become highly disordered. However, in both cases, we show that the shape of the isotherms remains unchanged. In the case of the mixtures of 9 μm particles with either 3 or 5 μm particles, the smaller particles aggregate around the larger particles through capillary interaction resulting in the formation of large fractal aggregates. At high compression, these layers contain holes that seem very compressible. As a result, the surface pressure isotherms show a smaller surface pressure jump than for other mixtures. textcopyright 2007 American Chemical Society.},\n\tAuthor = {Monteux, C. and Jung, E and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {7},\n\tPages = {3975--3980},\n\tTitle = {Mechanical properties and structure of particle coated interfaces: {Influence} of particle size and bidisperse 2D suspensions},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-34147202397&partnerID=40&md5=c90a8634a98156de9fe970b07465cf47},\n\tVolume = {23},\n\tYear = {2007},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-34147202397&partnerID=40&md5=c90a8634a98156de9fe970b07465cf47}}\n\n
\n
\n\n\n
\n We report surface pressure-area (prod-A) isotherms of bidisperse mixtures of anionic polystyrene latex particles at a water/n-decane interface as well as optical photographs of the interface for various compressions and mixture ratios. In the case of mixtures of 3 and 5 μm particles, we observe crystalline layers at high or low concentration ratios, where the \"impurity\" particles concentrate at the grain boundaries of the crystalline structure. At intermediate ratios, the layers become highly disordered. However, in both cases, we show that the shape of the isotherms remains unchanged. In the case of the mixtures of 9 μm particles with either 3 or 5 μm particles, the smaller particles aggregate around the larger particles through capillary interaction resulting in the formation of large fractal aggregates. At high compression, these layers contain holes that seem very compressible. As a result, the surface pressure isotherms show a smaller surface pressure jump than for other mixtures. textcopyright 2007 American Chemical Society.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2006\n \n \n (6)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Well-controlled living polymerization of perylene-labeled polyisoprenes and their use in single-molecule imaging.\n \n \n \n \n\n\n \n Gavranovic, G T; Csihony, S; Bowden, N B; Hawker, C J; Waymouth, R M; Moerner, W E; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 39(23): 8121–8127. 2006.\n \n\n\n\n
\n\n\n\n \n \n \"Well-controlledPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{gavranovic_well-controlled_2006,\n\tAbstract = {Functionalized perylene derivatives were synthesized with one or two alkoxyamine units. These compounds were used as initiators in the well-controlled living radical polymerization of isoprene, resulting in polymers with a perylene dye located at either the middle or end of the polymer. Using this approach, polyisoprenes were prepared with similar molecular weights, low polydispersities, and very good initiator efficiencies. Labeled chains were imaged on the single-molecule level in hosts of unlabeled polymer with similar composition. On the time scale from 0.2 to 2-5 s no difference between center- and end-positioned dyes was observed, suggesting that the dynamics outside of this time range should be studied. However, the most rigid host material led to the longest observed correlation times and the highest fraction of fluorophores that exhibited blinking behavior. textcopyright 2006 American Chemical Society.},\n\tAuthor = {Gavranovic, G T and Csihony, S and Bowden, N B and Hawker, C J and Waymouth, R M and Moerner, W E and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {23},\n\tPages = {8121--8127},\n\tTitle = {Well-controlled living polymerization of perylene-labeled polyisoprenes and their use in single-molecule imaging},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-33845262042&partnerID=40&md5=b1aae16ecb359c454e7978353593e2a7},\n\tVolume = {39},\n\tYear = {2006},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-33845262042&partnerID=40&md5=b1aae16ecb359c454e7978353593e2a7}}\n\n
\n
\n\n\n
\n Functionalized perylene derivatives were synthesized with one or two alkoxyamine units. These compounds were used as initiators in the well-controlled living radical polymerization of isoprene, resulting in polymers with a perylene dye located at either the middle or end of the polymer. Using this approach, polyisoprenes were prepared with similar molecular weights, low polydispersities, and very good initiator efficiencies. Labeled chains were imaged on the single-molecule level in hosts of unlabeled polymer with similar composition. On the time scale from 0.2 to 2-5 s no difference between center- and end-positioned dyes was observed, suggesting that the dynamics outside of this time range should be studied. However, the most rigid host material led to the longest observed correlation times and the highest fraction of fluorophores that exhibited blinking behavior. textcopyright 2006 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Effects of temperature and chemical modification on polymer langmuir films.\n \n \n \n\n\n \n Gavranovic, G T; Smith, M M; Jeong, W; Wong, A Y; Waymouth, R M; and Fuller, G.\n\n\n \n\n\n\n The Journal of Physical Chemistry B, 110(44): 22285–22290. 2006.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{gavranovic_effects_2006,\n\tAuthor = {Gavranovic, G T and Smith, M M and Jeong, W and Wong, A Y and Waymouth, R M and Fuller, G.G.},\n\tJournal = {The Journal of Physical Chemistry B},\n\tNumber = {44},\n\tPages = {22285--22290},\n\tTitle = {Effects of temperature and chemical modification on polymer langmuir films},\n\tVolume = {110},\n\tYear = {2006}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Interfacial rheology and structure of straight-chain and branched fatty alcohol mixtures.\n \n \n \n\n\n \n Kurtz, R. E; Lange, A.; and Fuller, G. G\n\n\n \n\n\n\n Langmuir, 22(12): 5321–5327. June 2006.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kurtz_interfacial_2006,\n\tAuthor = {Kurtz, Rachel E and Lange, Arno and Fuller, Gerald G},\n\tJournal = {Langmuir},\n\tMonth = jun,\n\tNumber = {12},\n\tPages = {5321--5327},\n\tTitle = {Interfacial rheology and structure of straight-chain and branched fatty alcohol mixtures},\n\tVolume = {22},\n\tYear = {2006}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Shear surface rheology of poly(N-isopropylacrylamide) adsorbed layers at the air-water interface.\n \n \n \n \n\n\n \n Monteux, C.; Mangeret, R; Laibe, G; Freyssingeas, E; Bergeron, V; and Fuller, G\n\n\n \n\n\n\n Macromolecules, 39(9): 3408–3414. 2006.\n \n\n\n\n
\n\n\n\n \n \n \"ShearPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{monteux_shear_2006,\n\tAbstract = {We investigate the shear surface rheology of a thermosensitive polymer, PNIPAM (poly(N-isopropylacrylamide)), adsorbed at the air-water interface using an interfacial stress rheometer. This polymer is characterized by a lower critical solution temperature, LCST, of 32textdegreeC; i.e., PNIPAM chains undergo a coil to globule transition as the temperature is increased above the LCST. We measure a spectacular increase in the surface shear elasticity and viscosity of PNIPAM layers as a function of the temperature around the LCST due to an increase in the amount of adsorbed species as the solvent quality decreases for PNIPAM chains. Moreover, the layers undergo a transition from a Newtonian liquidlike state to an elastic state due to an increase in intermolecular entanglements. We also show that the heating protocols have a strong influence on the rheological properties of the layers. In particular, we show that when the solutions are continuously heated from ambient temperature to well above the LCST, PNIPAM chains can be trapped in a nonequilibrium state at the air-water interface above the LCST. Last, we report on the influence that the PNIPAM bulk concentration and molecular mass have on the surface rheological properties at the air-water interface. textcopyright 2006 American Chemical Society.},\n\tAuthor = {Monteux, C. and Mangeret, R and Laibe, G and Freyssingeas, E and Bergeron, V and Fuller, G},\n\tJournal = {Macromolecules},\n\tNumber = {9},\n\tPages = {3408--3414},\n\tTitle = {Shear surface rheology of poly({N}-isopropylacrylamide) adsorbed layers at the air-water interface},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-33744485008&partnerID=40&md5=cd1b7ae9bfaa38d828844b4adb7a0547},\n\tVolume = {39},\n\tYear = {2006},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-33744485008&partnerID=40&md5=cd1b7ae9bfaa38d828844b4adb7a0547}}\n\n
\n
\n\n\n
\n We investigate the shear surface rheology of a thermosensitive polymer, PNIPAM (poly(N-isopropylacrylamide)), adsorbed at the air-water interface using an interfacial stress rheometer. This polymer is characterized by a lower critical solution temperature, LCST, of 32textdegreeC; i.e., PNIPAM chains undergo a coil to globule transition as the temperature is increased above the LCST. We measure a spectacular increase in the surface shear elasticity and viscosity of PNIPAM layers as a function of the temperature around the LCST due to an increase in the amount of adsorbed species as the solvent quality decreases for PNIPAM chains. Moreover, the layers undergo a transition from a Newtonian liquidlike state to an elastic state due to an increase in intermolecular entanglements. We also show that the heating protocols have a strong influence on the rheological properties of the layers. In particular, we show that when the solutions are continuously heated from ambient temperature to well above the LCST, PNIPAM chains can be trapped in a nonequilibrium state at the air-water interface above the LCST. Last, we report on the influence that the PNIPAM bulk concentration and molecular mass have on the surface rheological properties at the air-water interface. textcopyright 2006 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Packing, flipping, and buckling transitions in compressed monolayers of ellipsoidal latex particles.\n \n \n \n \n\n\n \n Basavaraj, M G; Fuller, G.; Fransaer, J; and Vermant, J\n\n\n \n\n\n\n Langmuir, 22(15): 6605–6612. 2006.\n \n\n\n\n
\n\n\n\n \n \n \"Packing,Paper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{basavaraj_packing_2006,\n\tAbstract = {The behavior of monolayers of monodisperse prolate ellipsoidal latex particles with the same surface chemistry but varying aspect ratio has been studied experimentally. Particle monolayers at an air-water interface were subjected to compression in a Langmuir trough. When surface pressure measurements and microscopy observations were combined, possible structural transitions were evaluated. Ellipsoids of a sufficiently large aspect ratio display a less abrupt increase in the compression isotherms than spherical particles. Microscopic observations reveal that a sequence of transitions is responsible for this more gradual increase of the surface pressure. When a percolating aggregate network is used as the starting point, locally ordered regions appear progressively. When it reaches a certain surface pressure, the system "jams", and in-plane rearrangements are no longer possible at this point. A highly localized yielding of the particle network is observed. The compressional stress is relieved by flipping the ellipsoids into an upright position and by expelling particles from the monolayer. The latter does not occur for spherical particles with similar dimensions and surface chemistry. In the final stage of compression, buckling of the monolayer as a whole was observed. The effect of aspect ratio on the pressure area isotherms and on the obtained percolation and packing thresholds was quantified. textcopyright 2006 American Chemical Society.},\n\tAuthor = {Basavaraj, M G and Fuller, G.G. and Fransaer, J and Vermant, J},\n\tJournal = {Langmuir},\n\tNumber = {15},\n\tPages = {6605--6612},\n\tTitle = {Packing, flipping, and buckling transitions in compressed monolayers of ellipsoidal latex particles},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-33746630127&partnerID=40&md5=1dd2983762a933889b411bb89d8c622a},\n\tVolume = {22},\n\tYear = {2006},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-33746630127&partnerID=40&md5=1dd2983762a933889b411bb89d8c622a}}\n\n
\n
\n\n\n
\n The behavior of monolayers of monodisperse prolate ellipsoidal latex particles with the same surface chemistry but varying aspect ratio has been studied experimentally. Particle monolayers at an air-water interface were subjected to compression in a Langmuir trough. When surface pressure measurements and microscopy observations were combined, possible structural transitions were evaluated. Ellipsoids of a sufficiently large aspect ratio display a less abrupt increase in the compression isotherms than spherical particles. Microscopic observations reveal that a sequence of transitions is responsible for this more gradual increase of the surface pressure. When a percolating aggregate network is used as the starting point, locally ordered regions appear progressively. When it reaches a certain surface pressure, the system \"jams\", and in-plane rearrangements are no longer possible at this point. A highly localized yielding of the particle network is observed. The compressional stress is relieved by flipping the ellipsoids into an upright position and by expelling particles from the monolayer. The latter does not occur for spherical particles with similar dimensions and surface chemistry. In the final stage of compression, buckling of the monolayer as a whole was observed. The effect of aspect ratio on the pressure area isotherms and on the obtained percolation and packing thresholds was quantified. textcopyright 2006 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Lipid-induced β-amyloid peptide assemblage fragmentation.\n \n \n \n \n\n\n \n Widenbrant, M J O; Rajadas, J; Sutardja, C; and Fuller, G.\n\n\n \n\n\n\n Biophysical Journal, 91(11): 4071–4080. 2006.\n \n\n\n\n
\n\n\n\n \n \n \"Lipid-inducedPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{widenbrant_lipid-induced_2006,\n\tAbstract = {Alzheimer's disease is the most common cause of dementia and is widely believed to be due to the accumulation of β-amyloid peptides (Aβ) and their interaction with the cell membrane. Aβs are hydrophobic peptides derived from the amyloid precursor proteins by proteolytic cleavage. After cleavage, these peptides are involved in a self-assembly-triggered conformational change. They are transformed into structures that bind to the cell membrane, causing cellular degeneration. However, it is not clear how these peptide assemblages disrupt the structural and functional integrity of the membrane. Membrane fluidity is one of the important parameters involved in pathophysiology of disease-affected cells. Probing the Abaggregate-lipid interactions will help us understand these processes with structural detail. Here we show that a fluid lipid monolayer develop immobile domains upon interaction with Aβ aggregates. Atomic force microscopy and transmission electron microscopy data indicate that peptide fibrils are fragmented into smaller nano-assemblages when interacting with the membrane lipids. Our findings could initiate reappraisal of the interactions between lipid assemblages and Aβ aggregates involved in Alzheimer's disease. textcopyright 2006 by the Biophysical Society.},\n\tAuthor = {Widenbrant, M J O and Rajadas, J and Sutardja, C and Fuller, G.G.},\n\tJournal = {Biophysical Journal},\n\tNumber = {11},\n\tPages = {4071--4080},\n\tTitle = {Lipid-induced β-amyloid peptide assemblage fragmentation},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-33845356957&partnerID=40&md5=54cf53a7f57586289a503eb40f0b8467},\n\tVolume = {91},\n\tYear = {2006},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-33845356957&partnerID=40&md5=54cf53a7f57586289a503eb40f0b8467}}\n\n
\n
\n\n\n
\n Alzheimer's disease is the most common cause of dementia and is widely believed to be due to the accumulation of β-amyloid peptides (Aβ) and their interaction with the cell membrane. Aβs are hydrophobic peptides derived from the amyloid precursor proteins by proteolytic cleavage. After cleavage, these peptides are involved in a self-assembly-triggered conformational change. They are transformed into structures that bind to the cell membrane, causing cellular degeneration. However, it is not clear how these peptide assemblages disrupt the structural and functional integrity of the membrane. Membrane fluidity is one of the important parameters involved in pathophysiology of disease-affected cells. Probing the Abaggregate-lipid interactions will help us understand these processes with structural detail. Here we show that a fluid lipid monolayer develop immobile domains upon interaction with Aβ aggregates. Atomic force microscopy and transmission electron microscopy data indicate that peptide fibrils are fragmented into smaller nano-assemblages when interacting with the membrane lipids. Our findings could initiate reappraisal of the interactions between lipid assemblages and Aβ aggregates involved in Alzheimer's disease. textcopyright 2006 by the Biophysical Society.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2005\n \n \n (7)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Investigation of shear-banding structure in wormlike micellar solution by point-wise flow-induced birefringence measurements.\n \n \n \n \n\n\n \n Lee, J Y; Fuller, G.; Hudson, N E; and Yuan, X F\n\n\n \n\n\n\n Journal of Rheology, 49(2): 537–550. 2005.\n \n\n\n\n
\n\n\n\n \n \n \"InvestigationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lee_investigation_2005,\n\tAbstract = {Shear banding occurs in many wormlike surfactant solutions subject to strong shear flow. We study this interesting phenomenon by pointwise flow-induced birefrigence (FIB) measurements in a well-known aqueous surfactant system, cetylpyridinium chloride/sodium salicylate. The dynamic nature of the local rheo-optical properties, such as birefringence and extinction angle across the gap of a Couette cell, were investigated over a range of shear rates, in particular the stress plateau region where shear stress is nearly independent of shear rate. This is probably the first application of point-wise FIB for the investigation of shear banding. The variation of optical signals along the radial direction of the Couette cell indicates that the fluid becomes inhomogeneous and forms bended structures, with either two bands or three bands across the gap, depending on the applied nominal shear rate. textcopyright 2005 The Society of Rheology.},\n\tAuthor = {Lee, J Y and Fuller, G.G. and Hudson, N E and Yuan, X F},\n\tJournal = {Journal of Rheology},\n\tNumber = {2},\n\tPages = {537--550},\n\tTitle = {Investigation of shear-banding structure in wormlike micellar solution by point-wise flow-induced birefringence measurements},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-31144474037&partnerID=40&md5=3ba2c232b499f28a8648f350110dc786},\n\tVolume = {49},\n\tYear = {2005},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-31144474037&partnerID=40&md5=3ba2c232b499f28a8648f350110dc786}}\n\n
\n
\n\n\n
\n Shear banding occurs in many wormlike surfactant solutions subject to strong shear flow. We study this interesting phenomenon by pointwise flow-induced birefrigence (FIB) measurements in a well-known aqueous surfactant system, cetylpyridinium chloride/sodium salicylate. The dynamic nature of the local rheo-optical properties, such as birefringence and extinction angle across the gap of a Couette cell, were investigated over a range of shear rates, in particular the stress plateau region where shear stress is nearly independent of shear rate. This is probably the first application of point-wise FIB for the investigation of shear banding. The variation of optical signals along the radial direction of the Couette cell indicates that the fluid becomes inhomogeneous and forms bended structures, with either two bands or three bands across the gap, depending on the applied nominal shear rate. textcopyright 2005 The Society of Rheology.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Two-dimensional melts: Polymer chains at the air - Water interface.\n \n \n \n \n\n\n \n Gavranovic, G T; Deutsch, J M; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 38(15): 6672–6679. 2005.\n \n\n\n\n
\n\n\n\n \n \n \"Two-dimensionalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{gavranovic_two-dimensional_2005,\n\tAbstract = {Both physical surface rheology and computer simulation experiments were performed to understand the conformation of polymers in 2D systems. The interfacial stress rheometer was used to measure surface rheological properties of Langmuir monolayers of poly(tert-butyl methacrylate) with molecular weights between 81.5 and 780 kg mol -1. The II - A isotherms for these monolayers show two transitions: a plateau at II approx18 mN m -1 and a bend at II approx 52 mN m -1. The properties of the films are shown to be substantially different above and below the plateau surface pressure, II p. Below II p, the monolayers are primarily viscous, and surface viscosity increases linearly with MW, while above II p, the films are more elastic, and surface viscosity is MW-independent. Computer simulations of these systems produce qualitatively similar II-A isotherms. The observed transition at II p marks the changeover from polymer chains existing in a single layer to forming regions of multilayers. textcopyright 2005 American Chemical Society.},\n\tAuthor = {Gavranovic, G T and Deutsch, J M and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {15},\n\tPages = {6672--6679},\n\tTitle = {Two-dimensional melts: {Polymer} chains at the air - {Water} interface},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-23744480810&partnerID=40&md5=25605c431a87c47c1067d6db54d0c13c},\n\tVolume = {38},\n\tYear = {2005},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-23744480810&partnerID=40&md5=25605c431a87c47c1067d6db54d0c13c}}\n\n
\n
\n\n\n
\n Both physical surface rheology and computer simulation experiments were performed to understand the conformation of polymers in 2D systems. The interfacial stress rheometer was used to measure surface rheological properties of Langmuir monolayers of poly(tert-butyl methacrylate) with molecular weights between 81.5 and 780 kg mol -1. The II - A isotherms for these monolayers show two transitions: a plateau at II approx18 mN m -1 and a bend at II approx 52 mN m -1. The properties of the films are shown to be substantially different above and below the plateau surface pressure, II p. Below II p, the monolayers are primarily viscous, and surface viscosity increases linearly with MW, while above II p, the films are more elastic, and surface viscosity is MW-independent. Computer simulations of these systems produce qualitatively similar II-A isotherms. The observed transition at II p marks the changeover from polymer chains existing in a single layer to forming regions of multilayers. textcopyright 2005 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Pickering emulsions with controllable stability.\n \n \n \n \n\n\n \n Melle, S; Lask, M; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 21(6): 2158–2162. 2005.\n \n\n\n\n
\n\n\n\n \n \n \"PickeringPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{melle_pickering_2005,\n\tAbstract = {We prepare solid-stabilized emulsions using paramagnetic particles at an oil/water interface that can undergo macroscopic phase separation upon application of an external magnetic field. A critical field strength is found for which emulsion droplets begin to translate into the continuous-phase fluid. At higher fields, the emulsions destabilize, leading to a fully phase-separated system. This effect is reversible, and long-term stability can be recovered by remixing the components with mechanical agitation. textcopyright 2005 American Chemical Society.},\n\tAuthor = {Melle, S and Lask, M and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {6},\n\tPages = {2158--2162},\n\tTitle = {Pickering emulsions with controllable stability},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-15844409782&partnerID=40&md5=fd7bf9f5b24e292ad21c475906df9817},\n\tVolume = {21},\n\tYear = {2005},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-15844409782&partnerID=40&md5=fd7bf9f5b24e292ad21c475906df9817}}\n\n
\n
\n\n\n
\n We prepare solid-stabilized emulsions using paramagnetic particles at an oil/water interface that can undergo macroscopic phase separation upon application of an external magnetic field. A critical field strength is found for which emulsion droplets begin to translate into the continuous-phase fluid. At higher fields, the emulsions destabilize, leading to a fully phase-separated system. This effect is reversible, and long-term stability can be recovered by remixing the components with mechanical agitation. textcopyright 2005 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Phase behavior and viscoelastic properties of trisilanolcyclohexyl-POSS at the air/water interface.\n \n \n \n\n\n \n Deng, J; Viers, B D; Esker, A R; Anseth, J W; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 21(6): 2375–2385. 2005.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{deng_phase_2005,\n\tAuthor = {Deng, J and Viers, B D and Esker, A R and Anseth, J W and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {6},\n\tPages = {2375--2385},\n\tTitle = {Phase behavior and viscoelastic properties of trisilanolcyclohexyl-{POSS} at the air/water interface},\n\tVolume = {21},\n\tYear = {2005}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Shape and buckling transitions in solid-stabilized drops.\n \n \n \n \n\n\n \n Xu, H.; Melle, S.; Golemanov, K.; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 21(22): 10016–10020. October 2005.\n \n\n\n\n
\n\n\n\n \n \n \"ShapePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{xu_shape_2005,\n\tAbstract = {We study shape and buckling transitions of particle-laden sessile and pendant droplets that are forced to shrink in size. Monodisperse polystyrene particles were placed at the interface between water and decane at conditions that are known to produce hexagonal, crystalline arrangements on flat interfaces. As the volumes of the drops are reduced, the surface areas are likewise diminished. This effectively compresses the particle monolayer coating and induces a transition from a fluid film to a solid film. Since the particles are firmly attached to the interface by capillary forces, the shape transitions are reversible and shape/volume curves are the same for drainage and inflation. Measurements of the internal pressure of the drops reveal a strong transition in this variable as the buckling transition is approached.},\n\tAuthor = {Xu, Hui and Melle, Sonia and Golemanov, Konstantin and Fuller, Gerald},\n\tDoi = {10.1021/la0507378},\n\tJournal = {Langmuir},\n\tLanguage = {English},\n\tMonth = oct,\n\tNumber = {22},\n\tPages = {10016--10020},\n\tPmid = {16229521},\n\tTitle = {Shape and buckling transitions in solid-stabilized drops.},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/la0507378},\n\tVolume = {21},\n\tYear = {2005},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/la0507378},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/la0507378}}\n\n
\n
\n\n\n
\n We study shape and buckling transitions of particle-laden sessile and pendant droplets that are forced to shrink in size. Monodisperse polystyrene particles were placed at the interface between water and decane at conditions that are known to produce hexagonal, crystalline arrangements on flat interfaces. As the volumes of the drops are reduced, the surface areas are likewise diminished. This effectively compresses the particle monolayer coating and induces a transition from a fluid film to a solid film. Since the particles are firmly attached to the interface by capillary forces, the shape transitions are reversible and shape/volume curves are the same for drainage and inflation. Measurements of the internal pressure of the drops reveal a strong transition in this variable as the buckling transition is approached.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Lung surfactant gelation induced by epithelial cells exposed to air pollution or oxidative stress.\n \n \n \n \n\n\n \n Anseth, J W; Goffin, A J; Fuller, G.; Ghio, A J; Kao, P N; and Upadhyay, D\n\n\n \n\n\n\n American Journal of Respiratory Cell and Molecular Biology, 33(2): 161–168. 2005.\n \n\n\n\n
\n\n\n\n \n \n \"LungPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{anseth_lung_2005,\n\tAbstract = {Lung surfactant lowers surface tension and adjusts interfacial rheology to facilitate breathing. A novel instrument, the interfacial stress rheometer (ISR), uses an oscillating magnetic needle to measure the shear viscosity and elasticity of a surfactant monolayer at the air-water interface. The ISR reveals that calf lung surfactant, Infasurf, exhibits remarkable fluidity, even when exposed to air pollution residual oil fly ash (ROFA), hydrogen peroxide (H 2O2), or conditioned media from resting A549 alveolar epithelial cells (AEC). However, when Infasurf is exposed to a subphase of the soluble fraction of ROFA- or H2O2-treated AEC conditioned media, there is a prominent increase in surfactant elasticity and viscosity, representing two-dimensional gelation. Surfactant gelation is decreased when ROFA-AEC are pretreated with inhibitors of cellular reactive oxygen species (ROS), or with a mitochondrial anion channel inhibitor, as well as when A549-pO cells that lack mitochondrial DNA and functional electron transport are investigated. These results implicate both mitochondrial and nonmitochondrial ROS generation in ROFA-AEC-induced surfactant gelation. A549 cells treated with H2O2 demonstrate a dose-dependent increase in lung surfactant gelation. The ISR is a unique and sensitive instrument to characterize surfactant gelation induced by oxidatively stressed AEC.},\n\tAuthor = {Anseth, J W and Goffin, A J and Fuller, G.G. and Ghio, A J and Kao, P N and Upadhyay, D},\n\tJournal = {American Journal of Respiratory Cell and Molecular Biology},\n\tNumber = {2},\n\tPages = {161--168},\n\tTitle = {Lung surfactant gelation induced by epithelial cells exposed to air pollution or oxidative stress},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-24344437775&partnerID=40&md5=a41466fdb41d88de6d345e253091102f},\n\tVolume = {33},\n\tYear = {2005},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-24344437775&partnerID=40&md5=a41466fdb41d88de6d345e253091102f}}\n\n
\n
\n\n\n
\n Lung surfactant lowers surface tension and adjusts interfacial rheology to facilitate breathing. A novel instrument, the interfacial stress rheometer (ISR), uses an oscillating magnetic needle to measure the shear viscosity and elasticity of a surfactant monolayer at the air-water interface. The ISR reveals that calf lung surfactant, Infasurf, exhibits remarkable fluidity, even when exposed to air pollution residual oil fly ash (ROFA), hydrogen peroxide (H 2O2), or conditioned media from resting A549 alveolar epithelial cells (AEC). However, when Infasurf is exposed to a subphase of the soluble fraction of ROFA- or H2O2-treated AEC conditioned media, there is a prominent increase in surfactant elasticity and viscosity, representing two-dimensional gelation. Surfactant gelation is decreased when ROFA-AEC are pretreated with inhibitors of cellular reactive oxygen species (ROS), or with a mitochondrial anion channel inhibitor, as well as when A549-pO cells that lack mitochondrial DNA and functional electron transport are investigated. These results implicate both mitochondrial and nonmitochondrial ROS generation in ROFA-AEC-induced surfactant gelation. A549 cells treated with H2O2 demonstrate a dose-dependent increase in lung surfactant gelation. The ISR is a unique and sensitive instrument to characterize surfactant gelation induced by oxidatively stressed AEC.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Optics of sheared liquid-crystal polarizer based on aqueous dispersion of dichroic-dye nano-aggregates.\n \n \n \n \n\n\n \n Paukshto, M; Fuller, G; Michailov, A; and Remizov, S\n\n\n \n\n\n\n Journal of the Society for Information Display, 13(9): 765–772. 2005.\n \n\n\n\n
\n\n\n\n \n \n \"OpticsPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{paukshto_optics_2005,\n\tAbstract = {The initial results on the optical properties of lyotropic liquid crystals under shear and following the cessation of flow are presented in the paper. Disk-shaped aromatic molecules that self-assemble into columnar nano-stacks in the water form the supramolecule - the basic structural element of these liquid crystals. The nature of the flow is an important factor determining the material microstructure, providing a global order and initial conditions for the drying process. The dried coating exhibits the properties of an E-type broadband polarizer. Understanding the dynamics of texture evolution and its interaction with the flow is of crucial importance to any development of polarizers coated with the liquid crystals. textcopyright Copyright 2005 Society for Information Display.},\n\tAuthor = {Paukshto, M and Fuller, G and Michailov, A and Remizov, S},\n\tJournal = {Journal of the Society for Information Display},\n\tNumber = {9},\n\tPages = {765--772},\n\tTitle = {Optics of sheared liquid-crystal polarizer based on aqueous dispersion of dichroic-dye nano-aggregates},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-54149118551&partnerID=40&md5=28012b39b4d695a0bfe0fdb9296dea57},\n\tVolume = {13},\n\tYear = {2005},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-54149118551&partnerID=40&md5=28012b39b4d695a0bfe0fdb9296dea57}}\n\n
\n
\n\n\n
\n The initial results on the optical properties of lyotropic liquid crystals under shear and following the cessation of flow are presented in the paper. Disk-shaped aromatic molecules that self-assemble into columnar nano-stacks in the water form the supramolecule - the basic structural element of these liquid crystals. The nature of the flow is an important factor determining the material microstructure, providing a global order and initial conditions for the drying process. The dried coating exhibits the properties of an E-type broadband polarizer. Understanding the dynamics of texture evolution and its interaction with the flow is of crucial importance to any development of polarizers coated with the liquid crystals. textcopyright Copyright 2005 Society for Information Display.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2004\n \n \n (9)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Rheological behavior of precursor PPV monolayers.\n \n \n \n \n\n\n \n Luinge, J W; Nijboer, G W; Hagting, J G; Vorenkamp, E J; Fuller, G.; and Scheuten, A J\n\n\n \n\n\n\n Langmuir, 20(26): 11517–11522. 2004.\n \n\n\n\n
\n\n\n\n \n \n \"RheologicalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{luinge_rheological_2004,\n\tAbstract = {The rheological behavior of different precursor poly(p-phenylene vinylene) (prec-PPV) monolayers at the air-water interface was investigated using an interfacial stress rheometer (ISR). This device nicely reveals a transition of the precursor poly(2,5-dimethoxy-1,4 phenylene vinylene) (prec-DMePPV) monolayer from Newtonian to elastic behavior with increasing surface pressure. The transition is accompanied by an increase in the modulus. This behavior coincides with the coagulation of different 2D condensed domains as revealed by Brewster angle microscopy (BAM). However, partly converted prec-DMePPV monolayers show elastic behavior even at low surface pressures, although a sudden increase of the moduli does occur. This phenomenon is attributed to enhanced hydrophobic interactions between the conjugated moieties in the partly converted polymers. The latter also explains the stretching behavior of the partly converted prec-DMePPV upon transfer in Langmuir-Blodgett-type vertical dipping. The increase of the moduli which is observed is much more gradual in the precursor poly(2,5-dibutoxy- 1,4-phenylene vinylene), prec DBuPPV, a monolayer which is in agreement with the expected expanded state of the latter monolayer.},\n\tAuthor = {Luinge, J W and Nijboer, G W and Hagting, J G and Vorenkamp, E J and Fuller, G.G. and Scheuten, A J},\n\tJournal = {Langmuir},\n\tNumber = {26},\n\tPages = {11517--11522},\n\tTitle = {Rheological behavior of precursor {PPV} monolayers},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-11244283870&partnerID=40&md5=5f11609923a563b771bb2cbac6c1477e},\n\tVolume = {20},\n\tYear = {2004},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-11244283870&partnerID=40&md5=5f11609923a563b771bb2cbac6c1477e}}\n\n
\n
\n\n\n
\n The rheological behavior of different precursor poly(p-phenylene vinylene) (prec-PPV) monolayers at the air-water interface was investigated using an interfacial stress rheometer (ISR). This device nicely reveals a transition of the precursor poly(2,5-dimethoxy-1,4 phenylene vinylene) (prec-DMePPV) monolayer from Newtonian to elastic behavior with increasing surface pressure. The transition is accompanied by an increase in the modulus. This behavior coincides with the coagulation of different 2D condensed domains as revealed by Brewster angle microscopy (BAM). However, partly converted prec-DMePPV monolayers show elastic behavior even at low surface pressures, although a sudden increase of the moduli does occur. This phenomenon is attributed to enhanced hydrophobic interactions between the conjugated moieties in the partly converted polymers. The latter also explains the stretching behavior of the partly converted prec-DMePPV upon transfer in Langmuir-Blodgett-type vertical dipping. The increase of the moduli which is observed is much more gradual in the precursor poly(2,5-dibutoxy- 1,4-phenylene vinylene), prec DBuPPV, a monolayer which is in agreement with the expected expanded state of the latter monolayer.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Interfacial Rheology of Globular and Flexible Proteins at the Hexadecane/Water Interface:\\textasciitilde Comparison of Shear and Dilatation Deformation.\n \n \n \n \n\n\n \n Freer, E. M; Yim, K. S.; Fuller, G. G; and Radke, C. J\n\n\n \n\n\n\n The Journal of Physical Chemistry B, 108(12): 3835–3844. 2004.\n \n\n\n\n
\n\n\n\n \n \n \"InterfacialPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{freer_interfacial_2004,\n\tAuthor = {Freer, Erik M and Yim, Kang Sub and Fuller, Gerald G and Radke, Clayton J},\n\tDoi = {10.1021/jp037236k},\n\tJournal = {The Journal of Physical Chemistry B},\n\tNumber = {12},\n\tPages = {3835--3844},\n\tTitle = {Interfacial {Rheology} of {Globular} and {Flexible} {Proteins} at the {Hexadecane}/{Water} {Interface}:{\\textbackslash}textasciitilde {Comparison} of {Shear} and {Dilatation} {Deformation}},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/jp037236k},\n\tVolume = {108},\n\tYear = {2004},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/jp037236k},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/jp037236k}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Dynamic transitions and oscillatory melting of a two-dimensional crystal subjected to shear flow.\n \n \n \n \n\n\n \n Stancik, E J; Hawkinson, A L; Vermant, J; and Fuller, G.\n\n\n \n\n\n\n Journal of Rheology, 48(1): 159–173. 2004.\n \n\n\n\n
\n\n\n\n \n \n \"DynamicPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{stancik_dynamic_2004,\n\tAbstract = {The effect of a delicate balance of forces on the dynamic behavior of monodisperse spherical polystyrene particles suspended at the interface between decane and water, was studied. The results demonstrated that the application of shear flow to particles arranged in a hexagonal lattice at the interface between decane and water caused the transition of the lattice structure to a new semiordered state. Particle concentration and shear rate were shown to be the key factors in determining the particle dynamics in this new state. A phase diagram compiled from data over a wide range of particle concentrations and shear rates outlined the transition between these two regimes.},\n\tAuthor = {Stancik, E J and Hawkinson, A L and Vermant, J and Fuller, G.G.},\n\tJournal = {Journal of Rheology},\n\tNumber = {1},\n\tPages = {159--173},\n\tTitle = {Dynamic transitions and oscillatory melting of a two-dimensional crystal subjected to shear flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-1642307337&partnerID=40&md5=5a8bd439b35f00254e95f3cfc7bc93ec},\n\tVolume = {48},\n\tYear = {2004},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-1642307337&partnerID=40&md5=5a8bd439b35f00254e95f3cfc7bc93ec}}\n\n
\n
\n\n\n
\n The effect of a delicate balance of forces on the dynamic behavior of monodisperse spherical polystyrene particles suspended at the interface between decane and water, was studied. The results demonstrated that the application of shear flow to particles arranged in a hexagonal lattice at the interface between decane and water caused the transition of the lattice structure to a new semiordered state. Particle concentration and shear rate were shown to be the key factors in determining the particle dynamics in this new state. A phase diagram compiled from data over a wide range of particle concentrations and shear rates outlined the transition between these two regimes.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Influence of Subphase Conditions on Interfacial Viscoelastic Properties of Synthetic Lipids with Gentiobiose Head Groups.\n \n \n \n\n\n \n Tanaka, M; Schiefer, S; Gege, C; Schmidt, R R; and Fuller, G.\n\n\n \n\n\n\n The Journal of Physical Chemistry B, 108(10): 3211–3214. 2004.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{tanaka_influence_2004,\n\tAuthor = {Tanaka, M and Schiefer, S and Gege, C and Schmidt, R R and Fuller, G.G.},\n\tJournal = {The Journal of Physical Chemistry B},\n\tNumber = {10},\n\tPages = {3211--3214},\n\tTitle = {Influence of {Subphase} {Conditions} on {Interfacial} {Viscoelastic} {Properties} of {Synthetic} {Lipids} with {Gentiobiose} {Head} {Groups}},\n\tVolume = {108},\n\tYear = {2004}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Coalescence of Particle-Laden Fluid Interfaces.\n \n \n \n \n\n\n \n Stancik, E J; Kouhkan, M; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 20(1): 90–94. 2004.\n \n\n\n\n
\n\n\n\n \n \n \"CoalescencePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{stancik_coalescence_2004,\n\tAbstract = {Colloidal particles are capable of stabilizing emulsions and, thus, slowing or preventing their complete breakdown into phase-separated systems. Direct observations of the dynamics of such particles on both water and oil droplets are reported as two colloid-laden interfaces are brought into contact with each other. As coalescence proceeds, the complementary systems, representing oil-in-water and water-in-oil emulsions, exhibit contrasting mechanisms for the formation of ring and disk structures by the particles as they serve to temporarily stabilize the approaching surfaces. An explanation of such behavior leads to a better understanding of the stabilization and breaking mechanisms of so-called Pickering emulsions.},\n\tAuthor = {Stancik, E J and Kouhkan, M and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {1},\n\tPages = {90--94},\n\tTitle = {Coalescence of {Particle}-{Laden} {Fluid} {Interfaces}},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0347758324&partnerID=40&md5=3c81e37db7c0be3b2ad558da453c2853},\n\tVolume = {20},\n\tYear = {2004},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0347758324&partnerID=40&md5=3c81e37db7c0be3b2ad558da453c2853}}\n\n
\n
\n\n\n
\n Colloidal particles are capable of stabilizing emulsions and, thus, slowing or preventing their complete breakdown into phase-separated systems. Direct observations of the dynamics of such particles on both water and oil droplets are reported as two colloid-laden interfaces are brought into contact with each other. As coalescence proceeds, the complementary systems, representing oil-in-water and water-in-oil emulsions, exhibit contrasting mechanisms for the formation of ring and disk structures by the particles as they serve to temporarily stabilize the approaching surfaces. An explanation of such behavior leads to a better understanding of the stabilization and breaking mechanisms of so-called Pickering emulsions.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Connect the drops: Using solids as adhesives for liquids.\n \n \n \n \n\n\n \n Stancik, E J; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 20(12): 4805–4808. 2004.\n \n\n\n\n
\n\n\n\n \n \n \"ConnectPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{stancik_connect_2004,\n\tAbstract = {Colloidal particles are shown to be capable of developing adhesion between liquid phases through a bridging mechanism by which intervening, micrometer-scaled, fluid films are stabilized. Particle dynamics leading to the assembly of the stabilizing structure are discussed. Models for the resulting adhesive force are developed from considerations of both interface shape perturbation and the force applied by surface tension on an individual particle. Finally, predictions from these models are compared to direct measurements of the forces that arise during the separation of adhering interfaces. Such comparisons lead to a novel method for determining the three-phase contact angle inherent to particles residing at fluid interfaces.},\n\tAuthor = {Stancik, E J and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {12},\n\tPages = {4805--4808},\n\tTitle = {Connect the drops: {Using} solids as adhesives for liquids},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-2942719124&partnerID=40&md5=edf6940bedad98a87eefd37a62922696},\n\tVolume = {20},\n\tYear = {2004},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-2942719124&partnerID=40&md5=edf6940bedad98a87eefd37a62922696}}\n\n
\n
\n\n\n
\n Colloidal particles are shown to be capable of developing adhesion between liquid phases through a bridging mechanism by which intervening, micrometer-scaled, fluid films are stabilized. Particle dynamics leading to the assembly of the stabilizing structure are discussed. Models for the resulting adhesive force are developed from considerations of both interface shape perturbation and the force applied by surface tension on an individual particle. Finally, predictions from these models are compared to direct measurements of the forces that arise during the separation of adhering interfaces. Such comparisons lead to a novel method for determining the three-phase contact angle inherent to particles residing at fluid interfaces.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Development characteristics of drag-reducing surfactant solution flow in a duct.\n \n \n \n \n\n\n \n Suzuki, H; Fuller, G.; Nakayama, T; and Usui, H\n\n\n \n\n\n\n Rheologica Acta, 43(3): 232–239. 2004.\n \n\n\n\n
\n\n\n\n \n \n \"DevelopmentPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{suzuki_development_2004,\n\tAuthor = {Suzuki, H and Fuller, G.G. and Nakayama, T and Usui, H},\n\tDoi = {10.1007/s00397-003-0335-6},\n\tJournal = {Rheologica Acta},\n\tNumber = {3},\n\tPages = {232--239},\n\tTitle = {Development characteristics of drag-reducing surfactant solution flow in a duct},\n\tUrl = {http://link.springer.com/10.1007/s00397-003-0335-6},\n\tVolume = {43},\n\tYear = {2004},\n\tBdsk-Url-1 = {http://link.springer.com/10.1007/s00397-003-0335-6},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1007/s00397-003-0335-6}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Shear and dilational surface rheology of oppositely charged polyelectrolyte/surfactant microgels adsorbed at the air-water interface. influence on foam stability.\n \n \n \n \n\n\n \n Monteux, C.; Fuller, G.; and Bergeron, V\n\n\n \n\n\n\n Journal of Physical Chemistry B, 108(42): 16473–16482. 2004.\n \n\n\n\n
\n\n\n\n \n \n \"ShearPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{monteux_shear_2004,\n\tAuthor = {Monteux, C. and Fuller, G.G. and Bergeron, V},\n\tDoi = {10.1021/jp047462},\n\tJournal = {Journal of Physical Chemistry B},\n\tLanguage = {English},\n\tNumber = {42},\n\tPages = {16473--16482},\n\tTitle = {Shear and dilational surface rheology of oppositely charged polyelectrolyte/surfactant microgels adsorbed at the air-water interface. influence on foam stability},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/jp047462%2B},\n\tVolume = {108},\n\tYear = {2004},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/jp047462%2B},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/jp047462}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Shear and dilatational relaxation mechanisms of globular and flexible proteins at the hexadecane/water interface.\n \n \n \n \n\n\n \n Freer, E. M; Yim, K. S.; Fuller, G. G; and Radke, C. J\n\n\n \n\n\n\n Langmuir, 20(23): 10159–10167. November 2004.\n \n\n\n\n
\n\n\n\n \n \n \"ShearPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{freer_shear_2004,\n\tAbstract = {Proteins adsorbed at fluid/fluid interfaces influence many phenomena: food emulsion and foam stability (Murray et al. Langmuir 2002, 18, 9476 and Borbas et al. Colloids Surf., A 2003, 213, 93), two-phase enzyme catalysis (Cascao-Pereira et al. Biotechnol. Bioeng. 2003, 83, 498; 2002, 78, 595), human lung function (Lunkenheimer et al. Colloids Surf., A 1996, 114, 199; Wustneck et al.; and Banerjee et al. 2000, 15, 14), and cell membrane mechanical properties (Mohandas et al. 1994, 23, 787). Time scales important to these phenomena are broad, necessitating an understanding of the dynamics of biological macromolecules at interfaces. We utilize interfacial shear and dilatational deformations to study the rheology of a globular protein, lysozyme, and a disordered protein, beta-casein, at the hexadecane/water interface. Linear viscoelastic properties are measured using small amplitude oscillatory flow, stress relaxation after a sudden dilatational displacement, and shear creep response to probe the rheological response over broad experimental time scales. Our studies of lysozyme and beta-casein reveal that the interfacial dissipation mechanisms are strongly coupled to changes in the protein structure upon and after adsorption. For beta-casein, the interfacial response is fluidlike in shear deformation and is dominated by interfacial viscous dissipation, particularly at low frequencies. Conversely, the dilatational response of beta-casein is dominated by diffusion dissipation at low frequencies and viscous dissipation at higher frequencies (i.e., when the experimental time scale is faster than the characteristic time for diffusion). For lysozyme in shear deformation, the adsorbed protein layer is primarily elastic with only a weak frequency dependence. Similarly, the interfacial dilatational moduli change very little with frequency. In comparison to beta-casein, the frequency response of lysozyme does not change substantially after washing the protein from the bulk solution. Apparently, it is the irreversibly adsorbed fraction that dominates the dynamic rheological response for lysozyme. Using stress relaxation after a sudden dilatational displacement and shear creep response, the characteristic time of relaxation was found to be 1000 s in both modes of deformation. The very long relaxation time for lysozyme likely results from the formation of a glassy interfacial network. This network develops at high interfacial concentrations where the molecules are highly constrained because of conformation changes that prevent desorption.},\n\tAuthor = {Freer, Erik M and Yim, Kang Sub and Fuller, Gerald G and Radke, Clayton J},\n\tDoi = {10.1021/la0485226},\n\tJournal = {Langmuir},\n\tLanguage = {English},\n\tMonth = nov,\n\tNumber = {23},\n\tPages = {10159--10167},\n\tPmid = {15518508},\n\tTitle = {Shear and dilatational relaxation mechanisms of globular and flexible proteins at the hexadecane/water interface.},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/la0485226},\n\tVolume = {20},\n\tYear = {2004},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/la0485226},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1021/la0485226}}\n\n
\n
\n\n\n
\n Proteins adsorbed at fluid/fluid interfaces influence many phenomena: food emulsion and foam stability (Murray et al. Langmuir 2002, 18, 9476 and Borbas et al. Colloids Surf., A 2003, 213, 93), two-phase enzyme catalysis (Cascao-Pereira et al. Biotechnol. Bioeng. 2003, 83, 498; 2002, 78, 595), human lung function (Lunkenheimer et al. Colloids Surf., A 1996, 114, 199; Wustneck et al.; and Banerjee et al. 2000, 15, 14), and cell membrane mechanical properties (Mohandas et al. 1994, 23, 787). Time scales important to these phenomena are broad, necessitating an understanding of the dynamics of biological macromolecules at interfaces. We utilize interfacial shear and dilatational deformations to study the rheology of a globular protein, lysozyme, and a disordered protein, beta-casein, at the hexadecane/water interface. Linear viscoelastic properties are measured using small amplitude oscillatory flow, stress relaxation after a sudden dilatational displacement, and shear creep response to probe the rheological response over broad experimental time scales. Our studies of lysozyme and beta-casein reveal that the interfacial dissipation mechanisms are strongly coupled to changes in the protein structure upon and after adsorption. For beta-casein, the interfacial response is fluidlike in shear deformation and is dominated by interfacial viscous dissipation, particularly at low frequencies. Conversely, the dilatational response of beta-casein is dominated by diffusion dissipation at low frequencies and viscous dissipation at higher frequencies (i.e., when the experimental time scale is faster than the characteristic time for diffusion). For lysozyme in shear deformation, the adsorbed protein layer is primarily elastic with only a weak frequency dependence. Similarly, the interfacial dilatational moduli change very little with frequency. In comparison to beta-casein, the frequency response of lysozyme does not change substantially after washing the protein from the bulk solution. Apparently, it is the irreversibly adsorbed fraction that dominates the dynamic rheological response for lysozyme. Using stress relaxation after a sudden dilatational displacement and shear creep response, the characteristic time of relaxation was found to be 1000 s in both modes of deformation. The very long relaxation time for lysozyme likely results from the formation of a glassy interfacial network. This network develops at high interfacial concentrations where the molecules are highly constrained because of conformation changes that prevent desorption.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2003\n \n \n (10)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Component stress-strain behavior and small-angle neutron scattering investigation of stereoblock elastomeric polypropylene.\n \n \n \n\n\n \n Wiyatno, W; Fuller, G.; Pople, J A; Gast, A P; Chen, Z R; Waymouth, R M; and Myers, C L\n\n\n \n\n\n\n Macromolecules, 36(4): 1178–1187. 2003.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wiyatno_component_2003,\n\tAuthor = {Wiyatno, W and Fuller, G.G. and Pople, J A and Gast, A P and Chen, Z R and Waymouth, R M and Myers, C L},\n\tJournal = {Macromolecules},\n\tNumber = {4},\n\tPages = {1178--1187},\n\tTitle = {Component stress-strain behavior and small-angle neutron scattering investigation of stereoblock elastomeric polypropylene},\n\tVolume = {36},\n\tYear = {2003}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Structure and dynamics of particle monolayers at a liquid-liquid interface subjected to shear flow.\n \n \n \n \n\n\n \n Stancik, E J; Gavranovic, G T; Widenbrant, M J O; Laschitsch, A T; Vermant, J; and Fuller, G.\n\n\n \n\n\n\n Faraday Discussions, 123: 145–156. 2003.\n \n\n\n\n
\n\n\n\n \n \n \"StructurePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{stancik_structure_2003,\n\tAbstract = {The effect of shear flow on the structure and dynamics of monodisperse spherical polystyrene particles suspended at the interface between decane and water was observed. While undisturbed, the particles arrange themselves on a hexagonal lattice due to strong dipole-dipole repulsion resulting from ionizable sulfate groups on their surfaces. As the interface is subjected to shear flow, however, the lattice adopts a new semi-ordered, anisotropic state for which two distinct regimes are observed. At low particle concentrations or high shear rates, nearest neighbors in the lattice align in the flow direction and create strings of particles that slip past each other fairly readily. This results in a stretching of the overall structure and achievement of a steady state orientation in the system. In contrast, at high concentrations or low shear rates, the interparticle forces gain importance and tend to keep the particles more strongly in their lattice positions. As a result, domains within the lattice are forced to rotate, thus giving rise to movement of particles perpendicular to the flow direction. Thus a rotation, in addition to stretching, of the structure is apparent in this case.},\n\tAuthor = {Stancik, E J and Gavranovic, G T and Widenbrant, M J O and Laschitsch, A T and Vermant, J and Fuller, G.G.},\n\tJournal = {Faraday Discussions},\n\tPages = {145--156},\n\tTitle = {Structure and dynamics of particle monolayers at a liquid-liquid interface subjected to shear flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037263197&partnerID=40&md5=ca696f5ac06705ac5fed926fc96a1bed},\n\tVolume = {123},\n\tYear = {2003},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037263197&partnerID=40&md5=ca696f5ac06705ac5fed926fc96a1bed}}\n\n
\n
\n\n\n
\n The effect of shear flow on the structure and dynamics of monodisperse spherical polystyrene particles suspended at the interface between decane and water was observed. While undisturbed, the particles arrange themselves on a hexagonal lattice due to strong dipole-dipole repulsion resulting from ionizable sulfate groups on their surfaces. As the interface is subjected to shear flow, however, the lattice adopts a new semi-ordered, anisotropic state for which two distinct regimes are observed. At low particle concentrations or high shear rates, nearest neighbors in the lattice align in the flow direction and create strings of particles that slip past each other fairly readily. This results in a stretching of the overall structure and achievement of a steady state orientation in the system. In contrast, at high concentrations or low shear rates, the interparticle forces gain importance and tend to keep the particles more strongly in their lattice positions. As a result, domains within the lattice are forced to rotate, thus giving rise to movement of particles perpendicular to the flow direction. Thus a rotation, in addition to stretching, of the structure is apparent in this case.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Shearing or compressing a soft glass in 2D: Time-concentration superposition.\n \n \n \n \n\n\n \n Cicuta, P; Stancik, E J; and Fuller, G.\n\n\n \n\n\n\n Physical Review Letters, 90(23): 236101/1–236101/4. 2003.\n \n\n\n\n
\n\n\n\n \n \n \"ShearingPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{cicuta_shearing_2003,\n\tAbstract = {Viscoelasticity was investigated quantitatively on two very different systems that were confined to fluid interfaces, where flow was effectively two dimensional. Latex colloids have a nondeformable hard core, and it was shown how a soft solid was formed and how the system progressively jams as the close-packing concentration was approached. It was believed that the generality of behavior between two dimensions (2D) and three dimensions (3D) should be of value and interest for both the theoretical understanding and the numerical simulations in this field, and also sheds light onto the origin of monolayer viscoelasticity and other related glassy behavior such as long time scale stress relaxation or aging.},\n\tAuthor = {Cicuta, P and Stancik, E J and Fuller, G.G.},\n\tJournal = {Physical Review Letters},\n\tNumber = {23},\n\tPages = {236101/1--236101/4},\n\tTitle = {Shearing or compressing a soft glass in 2D: {Time}-concentration superposition},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0041810539&partnerID=40&md5=a93031f3d079d49730701ef1a9c4e3d7},\n\tVolume = {90},\n\tYear = {2003},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0041810539&partnerID=40&md5=a93031f3d079d49730701ef1a9c4e3d7}}\n\n
\n
\n\n\n
\n Viscoelasticity was investigated quantitatively on two very different systems that were confined to fluid interfaces, where flow was effectively two dimensional. Latex colloids have a nondeformable hard core, and it was shown how a soft solid was formed and how the system progressively jams as the close-packing concentration was approached. It was believed that the generality of behavior between two dimensions (2D) and three dimensions (3D) should be of value and interest for both the theoretical understanding and the numerical simulations in this field, and also sheds light onto the origin of monolayer viscoelasticity and other related glassy behavior such as long time scale stress relaxation or aging.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Microstructure evolution in magnetorheological suspensions governed by mason number.\n \n \n \n \n\n\n \n Melle, S; Calder o n , O G; Rubio, M A; and Fuller, G.\n\n\n \n\n\n\n Physical Review E - Statistical, Nonlinear, and Soft Matter Physics, 68(4 1): 415031–41503111. 2003.\n \n\n\n\n
\n\n\n\n \n \n \"MicrostructurePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{melle_microstructure_2003,\n\tAbstract = {The study of the spatiotemporal evolution of the induced structures in dilute polarizable colloidal suspensions was discussed. The technique allowed to directly visualize the dynamics and needs of very dilute suspensions. It was found that the average chain length decreases with the increasing frequency. A Mason number crossover was found above which the rotation of the field prevented the particle aggregation to form chains.},\n\tAuthor = {Melle, S and Calder o n, O G and Rubio, M A and Fuller, G.G.},\n\tJournal = {Physical Review E - Statistical, Nonlinear, and Soft Matter Physics},\n\tNumber = {4 1},\n\tPages = {415031--41503111},\n\tTitle = {Microstructure evolution in magnetorheological suspensions governed by mason number},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-1542441413&partnerID=40&md5=549d778f48b2b0431e3ac23681192e55},\n\tVolume = {68},\n\tYear = {2003},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-1542441413&partnerID=40&md5=549d778f48b2b0431e3ac23681192e55}}\n\n
\n
\n\n\n
\n The study of the spatiotemporal evolution of the induced structures in dilute polarizable colloidal suspensions was discussed. The technique allowed to directly visualize the dynamics and needs of very dilute suspensions. It was found that the average chain length decreases with the increasing frequency. A Mason number crossover was found above which the rotation of the field prevented the particle aggregation to form chains.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Microstructural changes of a binary polymer blend in simple shear flow across the phase boundary.\n \n \n \n \n\n\n \n Vlassopoulos, D; Terakawa, T; and Fuller, G.\n\n\n \n\n\n\n Journal of Rheology, 47(1): 143–161. 2003.\n \n\n\n\n
\n\n\n\n \n \n \"MicrostructuralPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{vlassopoulos_microstructural_2003,\n\tAbstract = {The influence of shear flow on the microstructure of a partially miscible binary polymer blend, both in the homogeneous pretransitional and in the phase separated regions, was studied. Data were obtained in both the flow/gradient and flow/vorticity planes by using both Couette and parallel plate flow cells. Shear suppressed the pretransitional anisotropic concentration fluctuations yielding homogenization. In the late stages of the spinodal decomposition, the anisotropic deformation of the spinodal ring and its relaxation upon cessation depend on the distance from the critical point. The detection of scattering patterns with butterfly shape, due to the coupling of fluctuations with the flow, signified the onset of instabilities.},\n\tAuthor = {Vlassopoulos, D and Terakawa, T and Fuller, G.G.},\n\tJournal = {Journal of Rheology},\n\tNumber = {1},\n\tPages = {143--161},\n\tTitle = {Microstructural changes of a binary polymer blend in simple shear flow across the phase boundary},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037286225&partnerID=40&md5=49153ed4e519626e4ec783cd8af6ae0b},\n\tVolume = {47},\n\tYear = {2003},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037286225&partnerID=40&md5=49153ed4e519626e4ec783cd8af6ae0b}}\n\n
\n
\n\n\n
\n The influence of shear flow on the microstructure of a partially miscible binary polymer blend, both in the homogeneous pretransitional and in the phase separated regions, was studied. Data were obtained in both the flow/gradient and flow/vorticity planes by using both Couette and parallel plate flow cells. Shear suppressed the pretransitional anisotropic concentration fluctuations yielding homogenization. In the late stages of the spinodal decomposition, the anisotropic deformation of the spinodal ring and its relaxation upon cessation depend on the distance from the critical point. The detection of scattering patterns with butterfly shape, due to the coupling of fluctuations with the flow, signified the onset of instabilities.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Influence of phase transition and photoisomerization on interfacial rheology.\n \n \n \n \n\n\n \n Yim, K S; and Fuller, G.\n\n\n \n\n\n\n Physical Review E - Statistical, Nonlinear, and Soft Matter Physics, 67(4 1): 416011–4160110. 2003.\n \n\n\n\n
\n\n\n\n \n \n \"InfluencePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{yim_influence_2003,\n\tAbstract = {The thermodynamic phase transition and surface rheological properties of an azobenzene-containing fatty acid monolayer were investigated. Isotherm measurements indicated a strong first-order transition between the phases in the trans- isomer, which was confirmed by Brewster angle microscopy. Photoisomerization by ultraviolet light induced isomerism from the trans- to the cis- isomer and changed molecular structure from linear to bent-shaped forms.},\n\tAuthor = {Yim, K S and Fuller, G.G.},\n\tJournal = {Physical Review E - Statistical, Nonlinear, and Soft Matter Physics},\n\tNumber = {4 1},\n\tPages = {416011--4160110},\n\tTitle = {Influence of phase transition and photoisomerization on interfacial rheology},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0038507448&partnerID=40&md5=9600fb122a9f669c3dcac6bdebf7dd21},\n\tVolume = {67},\n\tYear = {2003},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0038507448&partnerID=40&md5=9600fb122a9f669c3dcac6bdebf7dd21}}\n\n
\n
\n\n\n
\n The thermodynamic phase transition and surface rheological properties of an azobenzene-containing fatty acid monolayer were investigated. Isotherm measurements indicated a strong first-order transition between the phases in the trans- isomer, which was confirmed by Brewster angle microscopy. Photoisomerization by ultraviolet light induced isomerism from the trans- to the cis- isomer and changed molecular structure from linear to bent-shaped forms.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Flow-Induced Anisotropy and Reversible Aggregation in Two-Dimensional Suspensions.\n \n \n \n \n\n\n \n Hoekstra, H; Vermant, J; Mewis, J; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 19(22): 9134–9141. 2003.\n \n\n\n\n
\n\n\n\n \n \n \"Flow-InducedPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hoekstra_flow-induced_2003,\n\tAbstract = {The time evolution of the flow-induced changes in structure of two-dimensional suspensions have been studied by means of video microscopy. The interparticle forces were tailored to produce a two-dimensional (2D) particulate network that could be reversibly broken down by means of a shear flow. Two types of suspensions have been investigated: systems with fairly strong attractive potentials in which essentially rigid bonds develop and systems with weaker attractions in which particles can still slide over each other at contact. For both suspensions interfacial shear flow causes the floes and their spatial organization to become anisotropic at various length scales. The FFT of the real space images is used to characterize the anisotropy at large length scales. Structural anisotropy at smaller length scales is deduced from the orientational dependence of the pair distribution function and from an harmonic expansion of g(r). The mechanism leading to the anisotropy during shear flow is shown to be related to a directional dependence of breakup and re-formation of floes. Interfacial flow also affects the density of the floes. The evolution of the distribution of coordination numbers with shear rate indicates that shear flow densities the rigid floes whereas the opposite occurs for the mobile ones.},\n\tAuthor = {Hoekstra, H and Vermant, J and Mewis, J and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {22},\n\tPages = {9134--9141},\n\tTitle = {Flow-{Induced} {Anisotropy} and {Reversible} {Aggregation} in {Two}-{Dimensional} {Suspensions}},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0242409115&partnerID=40&md5=6e2d994516b6191a76c4fc251b734ebc},\n\tVolume = {19},\n\tYear = {2003},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0242409115&partnerID=40&md5=6e2d994516b6191a76c4fc251b734ebc}}\n\n
\n
\n\n\n
\n The time evolution of the flow-induced changes in structure of two-dimensional suspensions have been studied by means of video microscopy. The interparticle forces were tailored to produce a two-dimensional (2D) particulate network that could be reversibly broken down by means of a shear flow. Two types of suspensions have been investigated: systems with fairly strong attractive potentials in which essentially rigid bonds develop and systems with weaker attractions in which particles can still slide over each other at contact. For both suspensions interfacial shear flow causes the floes and their spatial organization to become anisotropic at various length scales. The FFT of the real space images is used to characterize the anisotropy at large length scales. Structural anisotropy at smaller length scales is deduced from the orientational dependence of the pair distribution function and from an harmonic expansion of g(r). The mechanism leading to the anisotropy during shear flow is shown to be related to a directional dependence of breakup and re-formation of floes. Interfacial flow also affects the density of the floes. The evolution of the distribution of coordination numbers with shear rate indicates that shear flow densities the rigid floes whereas the opposite occurs for the mobile ones.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Interfacial rheology of graft-type polymeric siloxane surfactants.\n \n \n \n \n\n\n \n Anseth, J W; Bialek, A; Hill, R M; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 19(16): 6349–6356. 2003.\n \n\n\n\n
\n\n\n\n \n \n \"InterfacialPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{anseth_interfacial_2003,\n\tAbstract = {An interfacial stress rheometer has been used to study the rheological properties of a silicone oil/water interface in the presence of siloxane surfactants that are used in the personal care industry as water-in-silicone-oil emulsifiers. The surfactants are graft-type poly(dimethylsiloxane)-poly(oxyalkylene) copolymers. They appear to stabilize water-in-silicone-oil emulsions in a fashion similar to that of Pickering emulsions, in which solid particles, such as silica or clay, accumulate in the oil/water interface forming a solidlike "eggshell" that resists coalescence. These oil-soluble copolymers, when dissolved in silicone oil and contacted with water, spontaneously form small particles of a copolymer-rich phase dispersed in the oil phase. These particles appear to accumulate in the interface and act analogously to the solid particles in Pickering emulsions by forming a solidlike network in the silicone oil/water interface that can resist coalescence. The rheology of the interface is influenced by the formation and aggregation of these surfactant-rich particles in the silicone oil/water interface. Consequently, measuring the effects of different processing parameters on interfacial rheology can be used to study their effect on the formation of these surfactant-rich particles. Furthermore, the wetting behavior of the particle surface becomes more hydrophilic when exposed to water as evidenced by changes in contact angle measurements. Presumably, when these particles are in the silicone oil/water interface they present their hydrophilic poly(oxyethylene) domains toward water, resulting in a nonuniformity of surface wettability. This nonuniform surface wettability enhances these particles' ability to reside in the silicone oil/water interface.},\n\tAuthor = {Anseth, J W and Bialek, A and Hill, R M and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {16},\n\tPages = {6349--6356},\n\tTitle = {Interfacial rheology of graft-type polymeric siloxane surfactants},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0042071145&partnerID=40&md5=a141a455e330c6551f063843fd7191fc},\n\tVolume = {19},\n\tYear = {2003},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0042071145&partnerID=40&md5=a141a455e330c6551f063843fd7191fc}}\n\n
\n
\n\n\n
\n An interfacial stress rheometer has been used to study the rheological properties of a silicone oil/water interface in the presence of siloxane surfactants that are used in the personal care industry as water-in-silicone-oil emulsifiers. The surfactants are graft-type poly(dimethylsiloxane)-poly(oxyalkylene) copolymers. They appear to stabilize water-in-silicone-oil emulsions in a fashion similar to that of Pickering emulsions, in which solid particles, such as silica or clay, accumulate in the oil/water interface forming a solidlike \"eggshell\" that resists coalescence. These oil-soluble copolymers, when dissolved in silicone oil and contacted with water, spontaneously form small particles of a copolymer-rich phase dispersed in the oil phase. These particles appear to accumulate in the interface and act analogously to the solid particles in Pickering emulsions by forming a solidlike network in the silicone oil/water interface that can resist coalescence. The rheology of the interface is influenced by the formation and aggregation of these surfactant-rich particles in the silicone oil/water interface. Consequently, measuring the effects of different processing parameters on interfacial rheology can be used to study their effect on the formation of these surfactant-rich particles. Furthermore, the wetting behavior of the particle surface becomes more hydrophilic when exposed to water as evidenced by changes in contact angle measurements. Presumably, when these particles are in the silicone oil/water interface they present their hydrophilic poly(oxyethylene) domains toward water, resulting in a nonuniformity of surface wettability. This nonuniform surface wettability enhances these particles' ability to reside in the silicone oil/water interface.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The orientation dynamics of rigid rod suspensions under extensional flow.\n \n \n \n \n\n\n \n Pignon, F; Magnin, A; Piau, J M; and Fuller, G.\n\n\n \n\n\n\n Journal of Rheology, 47(2): 371–388. 2003.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{pignon_orientation_2003,\n\tAbstract = {The orientation dynamics of a lyotropic colloidal suspension of sepiolite clay under extensional flow have been explored by combined dichroism and small angle light scattering measurements. Extensional flow was applied using a four-roll mill to a thin film of sepiolite suspension (rigid rods 1 μm long and 0.010 μm in diameter). Analysis of transient extensional flow reversals revealed important characteristics of the orientation dynamics of these suspensions: (i) the existence of a critical volume fraction separating isotropic behavior, where no orientation persists after stretching, from nematic behavior, where permanent orientations persist during the relaxation phase; and (ii) in the nematic domain, a critical strain rate separates two flow regimes corresponding to a stable, so-called strong flow regime above the critical strain rate and an unstable, so-called weak flow regime below it. These experimental observations agree with the theoretical predictions of the model proposed by Marrucci and Maffettone (1989, 1990) who have examined the two-dimensional form of the simple molecular model of Hess (1976) and Doi (1981). What is new in the present case is that the colloidal suspension is a three-dimensional system, whereas previous experimental validations of the model concerned only two-dimensional rodlike polymers systems [Maffettone et al. (1996)] [Maruyama et al. (1998a, 1998b)]. textcopyright 2003 The Society of Rheology.},\n\tAuthor = {Pignon, F and Magnin, A and Piau, J M and Fuller, G.G.},\n\tJournal = {Journal of Rheology},\n\tNumber = {2},\n\tPages = {371--388},\n\tTitle = {The orientation dynamics of rigid rod suspensions under extensional flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0038342114&partnerID=40&md5=f2458bf324a436bf8f061676efbc0ba7},\n\tVolume = {47},\n\tYear = {2003},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0038342114&partnerID=40&md5=f2458bf324a436bf8f061676efbc0ba7}}\n\n
\n
\n\n\n
\n The orientation dynamics of a lyotropic colloidal suspension of sepiolite clay under extensional flow have been explored by combined dichroism and small angle light scattering measurements. Extensional flow was applied using a four-roll mill to a thin film of sepiolite suspension (rigid rods 1 μm long and 0.010 μm in diameter). Analysis of transient extensional flow reversals revealed important characteristics of the orientation dynamics of these suspensions: (i) the existence of a critical volume fraction separating isotropic behavior, where no orientation persists after stretching, from nematic behavior, where permanent orientations persist during the relaxation phase; and (ii) in the nematic domain, a critical strain rate separates two flow regimes corresponding to a stable, so-called strong flow regime above the critical strain rate and an unstable, so-called weak flow regime below it. These experimental observations agree with the theoretical predictions of the model proposed by Marrucci and Maffettone (1989, 1990) who have examined the two-dimensional form of the simple molecular model of Hess (1976) and Doi (1981). What is new in the present case is that the colloidal suspension is a three-dimensional system, whereas previous experimental validations of the model concerned only two-dimensional rodlike polymers systems [Maffettone et al. (1996)] [Maruyama et al. (1998a, 1998b)]. textcopyright 2003 The Society of Rheology.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n RHEOLOGY OF MOBILE INTERFACES.\n \n \n \n\n\n \n Fuller, G. G\n\n\n \n\n\n\n Rheology Rev.,77–123. 2003.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_rheology_2003,\n\tAbstract = {This review discusses the rheology of interfaces that separate two, immiscible fluids. Unlike interfaces at the fluid/solid boundary, these surfaces are mobile and deform. In the presence of amphiphiles that collect at these interfaces, they can become highly structured and non-Newtonian. These nonlinear responses to flow have profound implications on many physical processes and examples can be found in nature and industry. This review summarizes experimental methods in interfacial rheometry (both mechanical and optical methods are discussed), and presents their application to numerous classes of complex fluid interfaces. These interfaces come in a wide range of forms that are often directly analogous to their three-dimensional counterparts. This review presents results on classical fatty acids and alcohols, multiphase systems, rodlike amphiphiles, flexible chain amphiphiles, surface gelation, biopolymers, and two-dimensional suspensions.},\n\tAuthor = {Fuller, Gerald G},\n\tJournal = {Rheology Rev.},\n\tPages = {77--123},\n\tTitle = {{RHEOLOGY} {OF} {MOBILE} {INTERFACES}},\n\tYear = {2003}}\n\n
\n
\n\n\n
\n This review discusses the rheology of interfaces that separate two, immiscible fluids. Unlike interfaces at the fluid/solid boundary, these surfaces are mobile and deform. In the presence of amphiphiles that collect at these interfaces, they can become highly structured and non-Newtonian. These nonlinear responses to flow have profound implications on many physical processes and examples can be found in nature and industry. This review summarizes experimental methods in interfacial rheometry (both mechanical and optical methods are discussed), and presents their application to numerous classes of complex fluid interfaces. These interfaces come in a wide range of forms that are often directly analogous to their three-dimensional counterparts. This review presents results on classical fatty acids and alcohols, multiphase systems, rodlike amphiphiles, flexible chain amphiphiles, surface gelation, biopolymers, and two-dimensional suspensions.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2002\n \n \n (12)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Chain rotational dynamics in MR suspensions.\n \n \n \n \n\n\n \n Melle, S; Calder o n , O G; Rubio, M A; and Fuller, G.\n\n\n \n\n\n\n International Journal of Modern Physics B, 16(17-18): 2293–2299. 2002.\n \n\n\n\n
\n\n\n\n \n \n \"ChainPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{melle_chain_2002,\n\tAbstract = {The dynamics of induced dipolar chains in magnetorhelogical suspensions subject to rotating magnetic fields has been experimentally studied combining scattering dichroism and video microscopy experiments. When a rotating field is imposed the chainlike aggregates rotate synchronously with the magnetic field. We found that the average size of the aggregates decreases with Mason number (ratio of viscous to magnetic forces) following a power law with exponent -0.5 being the hydrodynamic friction forces the cause of the chains break up. However the total number of aggregated particles shows two different behaviors depending on Mason number. For low Mason numbers, the total number of aggregated particles remains almost constant and above a critical Mason number, the rotation of the field prevents the particle aggregation process from taking place so the number of aggregated particles decreases with Mason number following a power law behavior with exponent -1. Athermal molecular dynamics simulations are also reported, showing good agreement with the experiments.},\n\tAuthor = {Melle, S and Calder o n, O G and Rubio, M A and Fuller, G.G.},\n\tJournal = {International Journal of Modern Physics B},\n\tNumber = {17-18},\n\tPages = {2293--2299},\n\tTitle = {Chain rotational dynamics in {MR} suspensions},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037142928&partnerID=40&md5=12bdfaae25e0a7ebe5e4ab132865279c},\n\tVolume = {16},\n\tYear = {2002},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037142928&partnerID=40&md5=12bdfaae25e0a7ebe5e4ab132865279c}}\n\n
\n
\n\n\n
\n The dynamics of induced dipolar chains in magnetorhelogical suspensions subject to rotating magnetic fields has been experimentally studied combining scattering dichroism and video microscopy experiments. When a rotating field is imposed the chainlike aggregates rotate synchronously with the magnetic field. We found that the average size of the aggregates decreases with Mason number (ratio of viscous to magnetic forces) following a power law with exponent -0.5 being the hydrodynamic friction forces the cause of the chains break up. However the total number of aggregated particles shows two different behaviors depending on Mason number. For low Mason numbers, the total number of aggregated particles remains almost constant and above a critical Mason number, the rotation of the field prevents the particle aggregation process from taking place so the number of aggregated particles decreases with Mason number following a power law behavior with exponent -1. Athermal molecular dynamics simulations are also reported, showing good agreement with the experiments.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Dynamic response of stereoblock elastomeric polypropylene studied by rheooptics and X-ray scattering. 2. Orthogonally oriented crystalline chains.\n \n \n \n\n\n \n Wiyatno, W; Pople, J A; Gast, A P; Waymouth, R M; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 35(22): 8498–8508. 2002.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wiyatno_dynamic_2002,\n\tAuthor = {Wiyatno, W and Pople, J A and Gast, A P and Waymouth, R M and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {22},\n\tPages = {8498--8508},\n\tTitle = {Dynamic response of stereoblock elastomeric polypropylene studied by rheooptics and {X}-ray scattering. 2. {Orthogonally} oriented crystalline chains},\n\tVolume = {35},\n\tYear = {2002}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Surface rheological transitions in langmuir monolayers of bi-competitive fatty acids.\n \n \n \n\n\n \n Yim, K S; Rahaii, B; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 18(17): 6597–6601. 2002.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{yim_surface_2002,\n\tAuthor = {Yim, K S and Rahaii, B and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {17},\n\tPages = {6597--6601},\n\tTitle = {Surface rheological transitions in langmuir monolayers of bi-competitive fatty acids},\n\tVolume = {18},\n\tYear = {2002}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rotational dynamics in dipolar colloidal suspensions: Video microscopy experiments and simulations results.\n \n \n \n \n\n\n \n Melle, S; Calder o n , O G; Rubio, M A; and Fuller, G.\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 102(2): 135–148. 2002.\n \n\n\n\n
\n\n\n\n \n \n \"RotationalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{melle_rotational_2002,\n\tAbstract = {The dynamics of field-induced structures in very dilute dipolar colloidal suspensions subject to rotating magnetic fields have been experimentally studied using video microscopy. When a rotating field is imposed the chain-like aggregates rotate with the magnetic field frequency. We found that the size of the induced structures at small rotational frequencies is larger than at zero rotating frequency, i.e. when an uniaxial magnetic field is applied. At higher frequencies, the average size of the aggregates decreases with frequency following a power law with exponent -0.5 as the hydrodynamic friction forces overcome the dipolar magnetic forces, causing the chains break up. A non-thermal molecular dynamics simulations are also reported, showing good agreement with the experiments. textcopyright 2002 Elsevier Science B.V. All rights reserved.},\n\tAuthor = {Melle, S and Calder o n, O G and Rubio, M A and Fuller, G.G.},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {2},\n\tPages = {135--148},\n\tTitle = {Rotational dynamics in dipolar colloidal suspensions: {Video} microscopy experiments and simulations results},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037084420&partnerID=40&md5=fbd6809cd1539f6dc99a0b5714a024ba},\n\tVolume = {102},\n\tYear = {2002},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037084420&partnerID=40&md5=fbd6809cd1539f6dc99a0b5714a024ba}}\n\n
\n
\n\n\n
\n The dynamics of field-induced structures in very dilute dipolar colloidal suspensions subject to rotating magnetic fields have been experimentally studied using video microscopy. When a rotating field is imposed the chain-like aggregates rotate with the magnetic field frequency. We found that the size of the induced structures at small rotational frequencies is larger than at zero rotating frequency, i.e. when an uniaxial magnetic field is applied. At higher frequencies, the average size of the aggregates decreases with frequency following a power law with exponent -0.5 as the hydrodynamic friction forces overcome the dipolar magnetic forces, causing the chains break up. A non-thermal molecular dynamics simulations are also reported, showing good agreement with the experiments. textcopyright 2002 Elsevier Science B.V. All rights reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Polarizable particle aggregation under rotating magnetic fields using scattering dichroism.\n \n \n \n\n\n \n Melle, S; Calder o n , O G; Fuller, G.; and Rubio, M A\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 247(1): 200–209. 2002.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{melle_polarizable_2002,\n\tAuthor = {Melle, S and Calder o n, O G and Fuller, G.G. and Rubio, M A},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {1},\n\tPages = {200--209},\n\tTitle = {Polarizable particle aggregation under rotating magnetic fields using scattering dichroism},\n\tVolume = {247},\n\tYear = {2002}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Shear-banding structure orientated in the vorticity direction observed for equimolar micellar solution.\n \n \n \n \n\n\n \n Fischer, P; Wheeler, E K; and Fuller, G.\n\n\n \n\n\n\n Rheologica Acta, 41(1): 35–44. 2002.\n \n\n\n\n
\n\n\n\n \n \n \"Shear-bandingPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fischer_shear-banding_2002,\n\tAbstract = {The non-monotonic shear flow of a viscoelastic equimolar aqueous surfactant solution (cetylpyridinium chloride-sodium salicylate) is investigated rheologically and optically in a transparent strain-controlled Taylor Couette flow cell. As reported before, this particular wormlike micellar solution exhibits first a shear thinning and then a pronounced shear-thickening behavior. Once this shear-thickening regime is reached, a transient phase separation/shear banding of the solution into turbid and clear ring-like patterns orientated perpendicular to the vorticity axis, i.e., stacked like pancakes, is observed (Wheeler et al. 1998; Fischer 2000). The solution exhibit several unique features as no induction period of the shear induced phase, no structural build-up at the inner rotating cylinder, jumping pancake structure of clear and turbid ringlike phases, and oscillating shear stresses appear once the pancake structure is present. According to our analysis this flow phenomenon is not purely a mechanical or rheological driven hydrodynamic instability but one has to take into account structural changes of the oriented micellar aggregates (flow induced non-equilibrium phase transition) as proposed by several authors. Although this particular flow behavior and the underlying mixture of shear induced phases and mechanical instabilities is not fully understood yet, some classification characteristics based on a recent theoretical approach by Schmitt et al. (1995) and Porte et al. (1997) where a strong coupling between the flow instability (non-homogeneous flow profile due to the bands) and the structural changes causes the observed transient phenomena can be derived. In reference to the presented model the observed orientation of the rings is typical for complex fluids that undergo a spinodal phase separation coupled with a thermodynamic flow instability. In contrast to other shear banding phenomena, this one is observed in parallel plate, cone-plate, and Couette flow cell as well as under controlled stress and controlled rate conditions. Therefore, it adds an additional aspect to the present discussion on shear banding phenomena, i.e., the coupling of hydrodynamics and phase transition of rheological complex fluids.},\n\tAuthor = {Fischer, P and Wheeler, E K and Fuller, G.G.},\n\tJournal = {Rheologica Acta},\n\tNumber = {1},\n\tPages = {35--44},\n\tTitle = {Shear-banding structure orientated in the vorticity direction observed for equimolar micellar solution},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0036170284&partnerID=40&md5=b64885d7cd934c66b017f41e9bf2775d},\n\tVolume = {41},\n\tYear = {2002},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0036170284&partnerID=40&md5=b64885d7cd934c66b017f41e9bf2775d}}\n\n
\n
\n\n\n
\n The non-monotonic shear flow of a viscoelastic equimolar aqueous surfactant solution (cetylpyridinium chloride-sodium salicylate) is investigated rheologically and optically in a transparent strain-controlled Taylor Couette flow cell. As reported before, this particular wormlike micellar solution exhibits first a shear thinning and then a pronounced shear-thickening behavior. Once this shear-thickening regime is reached, a transient phase separation/shear banding of the solution into turbid and clear ring-like patterns orientated perpendicular to the vorticity axis, i.e., stacked like pancakes, is observed (Wheeler et al. 1998; Fischer 2000). The solution exhibit several unique features as no induction period of the shear induced phase, no structural build-up at the inner rotating cylinder, jumping pancake structure of clear and turbid ringlike phases, and oscillating shear stresses appear once the pancake structure is present. According to our analysis this flow phenomenon is not purely a mechanical or rheological driven hydrodynamic instability but one has to take into account structural changes of the oriented micellar aggregates (flow induced non-equilibrium phase transition) as proposed by several authors. Although this particular flow behavior and the underlying mixture of shear induced phases and mechanical instabilities is not fully understood yet, some classification characteristics based on a recent theoretical approach by Schmitt et al. (1995) and Porte et al. (1997) where a strong coupling between the flow instability (non-homogeneous flow profile due to the bands) and the structural changes causes the observed transient phenomena can be derived. In reference to the presented model the observed orientation of the rings is typical for complex fluids that undergo a spinodal phase separation coupled with a thermodynamic flow instability. In contrast to other shear banding phenomena, this one is observed in parallel plate, cone-plate, and Couette flow cell as well as under controlled stress and controlled rate conditions. Therefore, it adds an additional aspect to the present discussion on shear banding phenomena, i.e., the coupling of hydrodynamics and phase transition of rheological complex fluids.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Morphology of thermoplastic elastomers: Elastomeric polypropylene.\n \n \n \n\n\n \n Sch o nherr , H; Wiyatno, W; Pople, J; Frank, C.; Fuller, G.; Gast, A P; and Waymouth, R M\n\n\n \n\n\n\n Macromolecules, 35(7): 2654–2666. 2002.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{sch_o_nherr_morphology_2002,\n\tAuthor = {Sch o nherr, H and Wiyatno, W and Pople, J and Frank, C.W. and Fuller, G.G. and Gast, A P and Waymouth, R M},\n\tJournal = {Macromolecules},\n\tNumber = {7},\n\tPages = {2654--2666},\n\tTitle = {Morphology of thermoplastic elastomers: {Elastomeric} polypropylene},\n\tVolume = {35},\n\tYear = {2002}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Development of a double-beam rheo-optical analyzer for full tensor measurement of optical anisotropy in complex fluid flow.\n \n \n \n \n\n\n \n Takahashi, T; Shirakashi, M; Miyamoto, K; and Fuller, G.\n\n\n \n\n\n\n Rheologica Acta, 41(5): 448–455. 2002.\n \n\n\n\n
\n\n\n\n \n \n \"DevelopmentPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{takahashi_development_2002,\n\tAbstract = {The design and testing of the Double-Beam Rheo-Optical Analyzers, DB-ROA is described. The DB-ROA contains the two optically modulated laser beams that are transmitted through a sheared sample in different angles. The full components of the stress tensor can be evaluated simultaneously by a single measurement of the birefringence and orientation angle of each beam. The mechanical properties are also measured simultaneously with the optical properties. This system can be applied to both steady and transient shear flows. Three types of DB-ROAs are designed. They are distinguished by the optical modulation system, that is (1) a single modulator system, (2) a synchronization of two modulators, and (3) a two individual modulator system. The performance tests were carried out and its validity was demonstrated by comparison between the optical and the mechanical measurements of the first normal stress difference. The signal-to-noise ratio was strongly affected by the choice of the oblique angle of the second beam. The design features for the different optical modulation systems are discussed.},\n\tAuthor = {Takahashi, T and Shirakashi, M and Miyamoto, K and Fuller, G.G.},\n\tJournal = {Rheologica Acta},\n\tNumber = {5},\n\tPages = {448--455},\n\tTitle = {Development of a double-beam rheo-optical analyzer for full tensor measurement of optical anisotropy in complex fluid flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0036671548&partnerID=40&md5=b364b20c075c68567b0a0659d82cae52},\n\tVolume = {41},\n\tYear = {2002},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0036671548&partnerID=40&md5=b364b20c075c68567b0a0659d82cae52}}\n\n
\n
\n\n\n
\n The design and testing of the Double-Beam Rheo-Optical Analyzers, DB-ROA is described. The DB-ROA contains the two optically modulated laser beams that are transmitted through a sheared sample in different angles. The full components of the stress tensor can be evaluated simultaneously by a single measurement of the birefringence and orientation angle of each beam. The mechanical properties are also measured simultaneously with the optical properties. This system can be applied to both steady and transient shear flows. Three types of DB-ROAs are designed. They are distinguished by the optical modulation system, that is (1) a single modulator system, (2) a synchronization of two modulators, and (3) a two individual modulator system. The performance tests were carried out and its validity was demonstrated by comparison between the optical and the mechanical measurements of the first normal stress difference. The signal-to-noise ratio was strongly affected by the choice of the oblique angle of the second beam. The design features for the different optical modulation systems are discussed.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Structure and dynamics of particle monolayers at a liquid-liquid interface subjected to extensional flow.\n \n \n \n \n\n\n \n Stancik, E J; Widenbrant, M J O; Laschitsch, A T; Vermant, J; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 18(11): 4372–4375. 2002.\n \n\n\n\n
\n\n\n\n \n \n \"StructurePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{stancik_structure_2002,\n\tAbstract = {Monodisperse spherical polystyrene particles were suspended at the interface between decane and water, and then subjected to extensional flow. Their lattice structure was observed to pass from a hexagonal array at rest, through a liquidlike state as flow was first applied, and finally to a new semi-ordered, anisotropic state during steady flow. This semi-ordered state was shown to be oriented and stretched in the flow direction relative to the original hexagonal structure. The influence of interfacial concentration and extensional rate on particle dynamics is discussed.},\n\tAuthor = {Stancik, E J and Widenbrant, M J O and Laschitsch, A T and Vermant, J and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {11},\n\tPages = {4372--4375},\n\tTitle = {Structure and dynamics of particle monolayers at a liquid-liquid interface subjected to extensional flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037188693&partnerID=40&md5=89faa26ada148f340c27c384c084bb1c},\n\tVolume = {18},\n\tYear = {2002},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037188693&partnerID=40&md5=89faa26ada148f340c27c384c084bb1c}}\n\n
\n
\n\n\n
\n Monodisperse spherical polystyrene particles were suspended at the interface between decane and water, and then subjected to extensional flow. Their lattice structure was observed to pass from a hexagonal array at rest, through a liquidlike state as flow was first applied, and finally to a new semi-ordered, anisotropic state during steady flow. This semi-ordered state was shown to be oriented and stretched in the flow direction relative to the original hexagonal structure. The influence of interfacial concentration and extensional rate on particle dynamics is discussed.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Rheology of glycocalix model at air/water interface.\n \n \n \n\n\n \n Schneider, M F; Lim, K; Fuller, G.; and Tanaka, M\n\n\n \n\n\n\n Physical Chemistry Chemical Physics, 4(10): 1949–1952. 2002.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{schneider_rheology_2002,\n\tAuthor = {Schneider, M F and Lim, K and Fuller, G.G. and Tanaka, M},\n\tJournal = {Physical Chemistry Chemical Physics},\n\tNumber = {10},\n\tPages = {1949--1952},\n\tTitle = {Rheology of glycocalix model at air/water interface},\n\tVolume = {4},\n\tYear = {2002}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Dynamic response of stereoblock elastomeric polypropylene studied by rheooptics and x-ray scattering. 1. Influence of isotacticity.\n \n \n \n\n\n \n Wiyatno, W; Pople, J A; Gast, A P; Waymouth, R M; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 35(22): 8488–8497. 2002.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wiyatno_dynamic_2002-1,\n\tAuthor = {Wiyatno, W and Pople, J A and Gast, A P and Waymouth, R M and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {22},\n\tPages = {8488--8497},\n\tTitle = {Dynamic response of stereoblock elastomeric polypropylene studied by rheooptics and x-ray scattering. 1. {Influence} of isotacticity},\n\tVolume = {35},\n\tYear = {2002}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Surface shear rheology of a polymerizable lipopolymer monolayer.\n \n \n \n\n\n \n Brooks, C.; Thiele, J; Frank, C.; O'Brien, D F; Knoll, W; Fuller, G.; and Robertson, C.\n\n\n \n\n\n\n Langmuir, 18(6): 2166–2173. 2002.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{brooks_surface_2002,\n\tAuthor = {Brooks, C.F. and Thiele, J and Frank, C.W. and O'Brien, D F and Knoll, W and Fuller, G.G. and Robertson, C.R.},\n\tJournal = {Langmuir},\n\tNumber = {6},\n\tPages = {2166--2173},\n\tTitle = {Surface shear rheology of a polymerizable lipopolymer monolayer},\n\tVolume = {18},\n\tYear = {2002}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2001\n \n \n (8)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Isotropic-nematic phase transitions of lyotropic, two-dimensional liquid crystalline polymer solutions.\n \n \n \n \n\n\n \n Yim, K S; Fuller, G.; Datko, A; and Eisenbach, C D\n\n\n \n\n\n\n Macromolecules, 34(20): 6972–6977. 2001.\n \n\n\n\n
\n\n\n\n \n \n \"Isotropic-nematicPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{yim_isotropic-nematic_2001,\n\tAbstract = {Few studies of the isotropic-nematic phase transition temperature, T IN, in two dimensions have been reported experimentally in contrast to that of liquid crystalline polymers in three dimensions. The purpose of this paper is to examine molecular orientational hydrodynamics through studies of monolayer films at the air-water interface and to understand the phenomenon of the phase transition in two dimensions. UV absorption spectroscopy was used to determine molecular orientation in hairy-rod polymeric monolayers of poly(p-phenylene) sulfonic acid (PPPSH). A well-defined extensional flow is imposed in the monolayer to study the dynamics of flow-induced anisotropy. A solvent of stearic acid (SA) at moderate concentrations is added to the polymer solution to fluidize the film, and the effect of solvent on the isotropic-nematic transition is observed. Above T IN, complete relaxation of molecules is observed after flow cessation, while a nematic, ordered phase is obtained below T IN. Measurements of the surface rheological properties were also performed to further interrogate the phase transition.},\n\tAuthor = {Yim, K S and Fuller, G.G. and Datko, A and Eisenbach, C D},\n\tJournal = {Macromolecules},\n\tNumber = {20},\n\tPages = {6972--6977},\n\tTitle = {Isotropic-nematic phase transitions of lyotropic, two-dimensional liquid crystalline polymer solutions},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0035949902&partnerID=40&md5=058464b3d34a34d99a9133f1588aa654},\n\tVolume = {34},\n\tYear = {2001},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0035949902&partnerID=40&md5=058464b3d34a34d99a9133f1588aa654}}\n\n
\n
\n\n\n
\n Few studies of the isotropic-nematic phase transition temperature, T IN, in two dimensions have been reported experimentally in contrast to that of liquid crystalline polymers in three dimensions. The purpose of this paper is to examine molecular orientational hydrodynamics through studies of monolayer films at the air-water interface and to understand the phenomenon of the phase transition in two dimensions. UV absorption spectroscopy was used to determine molecular orientation in hairy-rod polymeric monolayers of poly(p-phenylene) sulfonic acid (PPPSH). A well-defined extensional flow is imposed in the monolayer to study the dynamics of flow-induced anisotropy. A solvent of stearic acid (SA) at moderate concentrations is added to the polymer solution to fluidize the film, and the effect of solvent on the isotropic-nematic transition is observed. Above T IN, complete relaxation of molecules is observed after flow cessation, while a nematic, ordered phase is obtained below T IN. Measurements of the surface rheological properties were also performed to further interrogate the phase transition.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Electrophoresis of DNA adsorbed to a cationic supported bilayer.\n \n \n \n \n\n\n \n Olson, D J; Johnson, J M; Patel, P D; Shaqfeh, E S G; Boxer, S G; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 17(23): 7396–7401. 2001.\n \n\n\n\n
\n\n\n\n \n \n \"ElectrophoresisPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{olson_electrophoresis_2001,\n\tAbstract = {We report fluorescence microscopy studies of the electrophoresis of individual DNA molecules electrostatically adsorbed to a cationic supported lipid bilayer. Obstacles to uniform electrophoretic flow cause the 2-D chains to adopt hooked conformations similar to those previously observed in 3-D electrophoresis experiments. Analysis of the stretch-contraction dynamics allows for an estimate of the obstacle density in the bilayer. Increasing the electric field causes the DNA molecules to become more highly stretched and increases the electrophoretic mobility substantially. A comparison of the Rouse relaxation time of the polymers and the average time between chain-obstacle collisions reveals that a single-obstacle model is insufficient to describe the observed dynamics but the obstacles are not dense enough to use a reptative model. Analysis of the unhooking dynamics reveals an 80\\% increase in hydrodynamic drag as compared to free chains. Finally, we observe anomalous diffusion of the DNA chains, with a large increase in the diffusion coefficient after the repeated application of high electric fields. Implications of the flow obstacles in the engineering of separation applications are discussed.},\n\tAuthor = {Olson, D J and Johnson, J M and Patel, P D and Shaqfeh, E S G and Boxer, S G and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {23},\n\tPages = {7396--7401},\n\tTitle = {Electrophoresis of {DNA} adsorbed to a cationic supported bilayer},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0035856668&partnerID=40&md5=af79fabaabf77a089ad401e323dc887b},\n\tVolume = {17},\n\tYear = {2001},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0035856668&partnerID=40&md5=af79fabaabf77a089ad401e323dc887b}}\n\n
\n
\n\n\n
\n We report fluorescence microscopy studies of the electrophoresis of individual DNA molecules electrostatically adsorbed to a cationic supported lipid bilayer. Obstacles to uniform electrophoretic flow cause the 2-D chains to adopt hooked conformations similar to those previously observed in 3-D electrophoresis experiments. Analysis of the stretch-contraction dynamics allows for an estimate of the obstacle density in the bilayer. Increasing the electric field causes the DNA molecules to become more highly stretched and increases the electrophoretic mobility substantially. A comparison of the Rouse relaxation time of the polymers and the average time between chain-obstacle collisions reveals that a single-obstacle model is insufficient to describe the observed dynamics but the obstacles are not dense enough to use a reptative model. Analysis of the unhooking dynamics reveals an 80% increase in hydrodynamic drag as compared to free chains. Finally, we observe anomalous diffusion of the DNA chains, with a large increase in the diffusion coefficient after the repeated application of high electric fields. Implications of the flow obstacles in the engineering of separation applications are discussed.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Orientation dynamics of magnetorheological fluids subject to rotating external fields.\n \n \n \n \n\n\n \n Melle, S; Rubio, M A; and Fuller, G.\n\n\n \n\n\n\n International Journal of Modern Physics B, 15(6-7): 758–766. 2001.\n \n\n\n\n
\n\n\n\n \n \n \"OrientationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{melle_orientation_2001,\n\tAbstract = {The formation and orientation of field-induced structures in magnetorheological (MR) fluids subject to rotating magnetic fields have been studied using two optical methods: scattering dichroism and small angle light scattering (SALS). The SALS patterns show how these chain-like aggregates follow the magnetic field with the same frequency but with a retarded phase angle for all the frequencies measured. Using scattering dichroism two different behaviors for both, dichroism and phase lag, are found below or above a critical frequency. Experimental results have been reproduced by a simple model considering the torques balance on the chain-like aggregates.},\n\tAuthor = {Melle, S and Rubio, M A and Fuller, G.G.},\n\tJournal = {International Journal of Modern Physics B},\n\tNumber = {6-7},\n\tPages = {758--766},\n\tTitle = {Orientation dynamics of magnetorheological fluids subject to rotating external fields},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0035917169&partnerID=40&md5=c544762a109080cafbfc74dcf7a79b11},\n\tVolume = {15},\n\tYear = {2001},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0035917169&partnerID=40&md5=c544762a109080cafbfc74dcf7a79b11}}\n\n
\n
\n\n\n
\n The formation and orientation of field-induced structures in magnetorheological (MR) fluids subject to rotating magnetic fields have been studied using two optical methods: scattering dichroism and small angle light scattering (SALS). The SALS patterns show how these chain-like aggregates follow the magnetic field with the same frequency but with a retarded phase angle for all the frequencies measured. Using scattering dichroism two different behaviors for both, dichroism and phase lag, are found below or above a critical frequency. Experimental results have been reproduced by a simple model considering the torques balance on the chain-like aggregates.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Two-dimensional physical networks of lipopolymers at the air/water interface.\n \n \n \n \n\n\n \n Frank, C.; Naumann, C A; Knoll, W; Brooks, C.; and Fuller, G.\n\n\n \n\n\n\n Macromolecular Symposia, 166(1): 1–12. March 2001.\n \n\n\n\n
\n\n\n\n \n \n \"Two-dimensionalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{frank_two-dimensional_2001,\n\tAuthor = {Frank, C.W. and Naumann, C A and Knoll, W and Brooks, C.F. and Fuller, G.G.},\n\tDoi = {10.1002/1521-3900(200103)166:1<1::AID-MASY1>3.0.CO;2-Y},\n\tJournal = {Macromolecular Symposia},\n\tMonth = mar,\n\tNumber = {1},\n\tPages = {1--12},\n\tTitle = {Two-dimensional physical networks of lipopolymers at the air/water interface},\n\tUrl = {http://doi.wiley.com/10.1002/1521-3900%28200103%29166%3A1%3C1%3A%3AAID-MASY1%3E3.0.CO%3B2-Y},\n\tVolume = {166},\n\tYear = {2001},\n\tBdsk-Url-1 = {http://doi.wiley.com/10.1002/1521-3900%28200103%29166%3A1%3C1%3A%3AAID-MASY1%3E3.0.CO%3B2-Y},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1002/1521-3900(200103)166:1%3C1::AID-MASY1%3E3.0.CO;2-Y}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Two-dimensional physical networks of lipopolymers at the air/water interface: Correlation of molecular structure and surface rheological behavior.\n \n \n \n \n\n\n \n Naumann, C A; Brooks, C.; Fuller, G.; Lehmann, T; R u he , J; Knoll, W; Kuhn, P; Nuyken, O; and Frank, C.\n\n\n \n\n\n\n Langmuir, 17(9): 2801–2806. 2001.\n \n\n\n\n
\n\n\n\n \n \n \"Two-dimensionalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{naumann_two-dimensional_2001,\n\tAbstract = {Recent surface rheology and film balance experiments on monolayers of PEG lipopolymers and phospholipid/PEG lipopolymer mixtures at the air-water interface have revealed a new class of quasi two-dimensional physical networks. Two different kinds of associative interactions are necessary to form the network: microcondensation of alkyl chains of lipopolymers to form small clusters and water molecule mediation of the interaction between adjacent PEG clusters via hydrogen bonding. In the experiments presented here, we are interested to learn whether the physical gelation is PEG specific or whether it is a more general characteristic of lipopolymers at the air-water interface. To address this topic, we have expanded our surface rheology and film balance experiments to poly(oxazoline) lipopolymers. Our experiments indicate the occurrence of a rheological transition if the poly(oxazoline) lipopolymers consist of a dioctadecylglycerol anchor. This shows that the physical gelation among lipopolymers is not a PEG-specific phenomenon. No physical gelation is found, however, if the dioctadecylglycerol anchor of the lipopolymer is replaced by a dioctadecylamine anchor. The observed importance of the hydrophobic anchor supports our previous findings that the alkyl chain condensation should be seen as one of two kinds of physical junctions necessary for the formation of the physical network.},\n\tAuthor = {Naumann, C A and Brooks, C.F. and Fuller, G.G. and Lehmann, T and R u he, J and Knoll, W and Kuhn, P and Nuyken, O and Frank, C.W.},\n\tJournal = {Langmuir},\n\tNumber = {9},\n\tPages = {2801--2806},\n\tTitle = {Two-dimensional physical networks of lipopolymers at the air/water interface: {Correlation} of molecular structure and surface rheological behavior},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0035353375&partnerID=40&md5=865093475b0fd2be6594ff7d4850e146},\n\tVolume = {17},\n\tYear = {2001},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0035353375&partnerID=40&md5=865093475b0fd2be6594ff7d4850e146}}\n\n
\n
\n\n\n
\n Recent surface rheology and film balance experiments on monolayers of PEG lipopolymers and phospholipid/PEG lipopolymer mixtures at the air-water interface have revealed a new class of quasi two-dimensional physical networks. Two different kinds of associative interactions are necessary to form the network: microcondensation of alkyl chains of lipopolymers to form small clusters and water molecule mediation of the interaction between adjacent PEG clusters via hydrogen bonding. In the experiments presented here, we are interested to learn whether the physical gelation is PEG specific or whether it is a more general characteristic of lipopolymers at the air-water interface. To address this topic, we have expanded our surface rheology and film balance experiments to poly(oxazoline) lipopolymers. Our experiments indicate the occurrence of a rheological transition if the poly(oxazoline) lipopolymers consist of a dioctadecylglycerol anchor. This shows that the physical gelation among lipopolymers is not a PEG-specific phenomenon. No physical gelation is found, however, if the dioctadecylglycerol anchor of the lipopolymer is replaced by a dioctadecylamine anchor. The observed importance of the hydrophobic anchor supports our previous findings that the alkyl chain condensation should be seen as one of two kinds of physical junctions necessary for the formation of the physical network.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Time scaling regimes in aggregation of magnetic dipolar particles: Scattering dichroism results.\n \n \n \n\n\n \n Melle, S; Rubio, M A; and Fuller, G.\n\n\n \n\n\n\n Physical Review Letters, 87(11): 115501/1–115501/4. 2001.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{melle_time_2001,\n\tAuthor = {Melle, S and Rubio, M A and Fuller, G.G.},\n\tJournal = {Physical Review Letters},\n\tNumber = {11},\n\tPages = {115501/1--115501/4},\n\tTitle = {Time scaling regimes in aggregation of magnetic dipolar particles: {Scattering} dichroism results},\n\tVolume = {87},\n\tYear = {2001}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Rheological properties of lipopolymer-phospholipid mixtures at the air-water interface: A novel form of two-dimensional physical gelation.\n \n \n \n\n\n \n Naumann, C A; Brooks, C.; Wiyatno, W; Knoll, W; Fuller, G.; and Frank, C.\n\n\n \n\n\n\n Macromolecules, 34(9): 3024–3032. 2001.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{naumann_rheological_2001,\n\tAuthor = {Naumann, C A and Brooks, C.F. and Wiyatno, W and Knoll, W and Fuller, G.G. and Frank, C.W.},\n\tJournal = {Macromolecules},\n\tNumber = {9},\n\tPages = {3024--3032},\n\tTitle = {Rheological properties of lipopolymer-phospholipid mixtures at the air-water interface: {A} novel form of two-dimensional physical gelation},\n\tVolume = {34},\n\tYear = {2001}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rheooptical determination of aspect ratio and polydispersity of nonspherical particles.\n \n \n \n \n\n\n \n Vermant, J; Yang, H; and Fuller, G.\n\n\n \n\n\n\n AIChE Journal, 47(4): 790–798. 2001.\n \n\n\n\n
\n\n\n\n \n \n \"RheoopticalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{vermant_rheooptical_2001,\n\tAbstract = {A rheooptical method proposed rapidly determines the aspect ratio and polydispersity of small axisymmetric, nonspherical particles. The time evolution of the average orientation angle of optically isotropic, nonabsorbing particles upon inception of a shear flow was monitored by a polarization modulation method by using dilute suspensions of the particles. Since the orientation angle is a geometric property of the particles, no arbitrary assumptions on the scattering mechanism had to be made to analyze the data. The period of the damped oscillatory response of the orientation angle is related to the average aspect ratio. In the limit of strong flows the damping only depends on dispersion in particle shape. An analytical expression to relate the polydispersity of the equivalent hydrodynamic aspect ratio to the damping function is presented, as well as the applicability of the technique by studying the hydrodynamic aspect ratio and associated polydispersity in a sample containing ellipsoidal hematite particles.},\n\tAuthor = {Vermant, J and Yang, H and Fuller, G.G.},\n\tJournal = {AIChE Journal},\n\tNumber = {4},\n\tPages = {790--798},\n\tTitle = {Rheooptical determination of aspect ratio and polydispersity of nonspherical particles},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037489757&partnerID=40&md5=710fbac9e65a8243169b59244c20d9cd},\n\tVolume = {47},\n\tYear = {2001},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0037489757&partnerID=40&md5=710fbac9e65a8243169b59244c20d9cd}}\n\n
\n
\n\n\n
\n A rheooptical method proposed rapidly determines the aspect ratio and polydispersity of small axisymmetric, nonspherical particles. The time evolution of the average orientation angle of optically isotropic, nonabsorbing particles upon inception of a shear flow was monitored by a polarization modulation method by using dilute suspensions of the particles. Since the orientation angle is a geometric property of the particles, no arbitrary assumptions on the scattering mechanism had to be made to analyze the data. The period of the damped oscillatory response of the orientation angle is related to the average aspect ratio. In the limit of strong flows the damping only depends on dispersion in particle shape. An analytical expression to relate the polydispersity of the equivalent hydrodynamic aspect ratio to the damping function is presented, as well as the applicability of the technique by studying the hydrodynamic aspect ratio and associated polydispersity in a sample containing ellipsoidal hematite particles.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2000\n \n \n (7)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n On the existence of a stress-optical relation in immiscible polymer blends.\n \n \n \n \n\n\n \n Van Puyvelde, P; Moldenaers, P; Mewis, J; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 16(Washington, DC, United States): 3740–3747. 2000.\n \n\n\n\n
\n\n\n\n \n \n \"OnPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{van_puyvelde_existence_2000,\n\tAbstract = {In emulsions and immiscible polymer blends linear conservative dichroism is known to be very sensitive to flow-induced microstructural changes during flow. The same holds for the excess or interfacial contribution to the first normal stress. Similarities have already been observed between these two properties in blends. Here it is investigated whether this similarity reflects a more quantitative relationship. The interfacial normal stress can be described by means of an interface anisotropy tensor. Therefore, the measured dichroism will be compared with the calculated components of this anisotropy tensor for different flow histories. These include steady-state shear flow, sudden increase in shear rate, relaxation with droplet breakup, and oscillatory flow. In all cases a proportionality factor is obtained. For strongly deformed droplets direct theoretical evidence for a quantitative relationship between rheology and rheo-optics in blends is provided. To extend this relationship to a larger window of shear rates, the size dependence of the scattering has to be taken into account.},\n\tAuthor = {Van Puyvelde, P and Moldenaers, P and Mewis, J and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {Washington, DC, United States},\n\tPages = {3740--3747},\n\tTitle = {On the existence of a stress-optical relation in immiscible polymer blends},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033751126&partnerID=40&md5=a52a6cf61bd69942f08eb2e9a4e8a0ad},\n\tVolume = {16},\n\tYear = {2000},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033751126&partnerID=40&md5=a52a6cf61bd69942f08eb2e9a4e8a0ad}}\n\n
\n
\n\n\n
\n In emulsions and immiscible polymer blends linear conservative dichroism is known to be very sensitive to flow-induced microstructural changes during flow. The same holds for the excess or interfacial contribution to the first normal stress. Similarities have already been observed between these two properties in blends. Here it is investigated whether this similarity reflects a more quantitative relationship. The interfacial normal stress can be described by means of an interface anisotropy tensor. Therefore, the measured dichroism will be compared with the calculated components of this anisotropy tensor for different flow histories. These include steady-state shear flow, sudden increase in shear rate, relaxation with droplet breakup, and oscillatory flow. In all cases a proportionality factor is obtained. For strongly deformed droplets direct theoretical evidence for a quantitative relationship between rheology and rheo-optics in blends is provided. To extend this relationship to a larger window of shear rates, the size dependence of the scattering has to be taken into account.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Surface pressure-induced isotropic-nematic transition in polymer monolayers-effect of solvent molecules.\n \n \n \n \n\n\n \n Yim, K S; Brooks, C.; Fuller, G.; Datko, A; and Eisenbach, C D\n\n\n \n\n\n\n Langmuir, 16(Washington, DC, United States): 4319–4324. 2000.\n \n\n\n\n
\n\n\n\n \n \n \"SurfacePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{yim_surface_2000,\n\tAbstract = {The coupling of hydrodynamics to molecular orientation has been studied for a Langmuir film of the hairy-rod polymer poly(p-phenylene)sulfonic acid mixed with either arachidyl alcohol or 1,2-dimyristoyl-sn-glycero-3-phosphatidylcholine. These latter materials are small amphiphilic molecules. The orientational dynamics of the polymer under an extensional flow is examined using UV dichroism. The results reveal a surface pressure-induced isotropic-nematic transition, which is affected by polymer-solvent interactions as well as polymer-polymer interactions. An interfacial stress rheometer was used to further examine the existence of the surface pressure-induced transition. These structural and liquid crystalline transitions are difficult to observe using traditional pressure-area isotherm analysis.},\n\tAuthor = {Yim, K S and Brooks, C.F. and Fuller, G.G. and Datko, A and Eisenbach, C D},\n\tJournal = {Langmuir},\n\tNumber = {Washington, DC, United States},\n\tPages = {4319--4324},\n\tTitle = {Surface pressure-induced isotropic-nematic transition in polymer monolayers-effect of solvent molecules},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033727616&partnerID=40&md5=4bde95b962bb918946ea99d173b21f65},\n\tVolume = {16},\n\tYear = {2000},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033727616&partnerID=40&md5=4bde95b962bb918946ea99d173b21f65}}\n\n
\n
\n\n\n
\n The coupling of hydrodynamics to molecular orientation has been studied for a Langmuir film of the hairy-rod polymer poly(p-phenylene)sulfonic acid mixed with either arachidyl alcohol or 1,2-dimyristoyl-sn-glycero-3-phosphatidylcholine. These latter materials are small amphiphilic molecules. The orientational dynamics of the polymer under an extensional flow is examined using UV dichroism. The results reveal a surface pressure-induced isotropic-nematic transition, which is affected by polymer-solvent interactions as well as polymer-polymer interactions. An interfacial stress rheometer was used to further examine the existence of the surface pressure-induced transition. These structural and liquid crystalline transitions are difficult to observe using traditional pressure-area isotherm analysis.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Non-Newtonian rheology of liquid crystalline polymer monolayers.\n \n \n \n \n\n\n \n Yim, K S; Brooks, C.; Fuller, G.; Winter, D; and Eisenbach, C D\n\n\n \n\n\n\n Langmuir, 16(Washington, DC, United States): 4325–4332. 2000.\n \n\n\n\n
\n\n\n\n \n \n \"Non-NewtonianPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{yim_non-newtonian_2000,\n\tAbstract = {A rheological study on a two-dimensional Langmuir monolayer has been conducted by measuring the optical anisotropy and the mechanical surface properties. Experiments were conducted on monolayer mixtures of a liquid crystalline polymer, poly(p-phenylene)sulfonic acid (PPPSH), with an aliphatic fatty acid, stearic acid (C18). It is found that these two molecules mix ideally, as supported by studies of average molecular area, collapse pressure, and Brewster angle microscopy. Both surface moduli and surface viscosity were obtained with an interfacial stress rheometer and showed that PPPSH/stearic acid monolayer mixtures are non-Newtonian. The orientational dynamics of polymers under well-defined planar extensional flow was considered using UV dichroism, which revealed a surface pressure-induced isotropic-nematic transition. This transition is difficult to observe using traditional isotherm analysis; however, it is readily evident in measurements of the rheology of the monolayers.},\n\tAuthor = {Yim, K S and Brooks, C.F. and Fuller, G.G. and Winter, D and Eisenbach, C D},\n\tJournal = {Langmuir},\n\tNumber = {Washington, DC, United States},\n\tPages = {4325--4332},\n\tTitle = {Non-{Newtonian} rheology of liquid crystalline polymer monolayers},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033727898&partnerID=40&md5=9c3c3e4cf8efcd3f678c3362319f2475},\n\tVolume = {16},\n\tYear = {2000},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033727898&partnerID=40&md5=9c3c3e4cf8efcd3f678c3362319f2475}}\n\n
\n
\n\n\n
\n A rheological study on a two-dimensional Langmuir monolayer has been conducted by measuring the optical anisotropy and the mechanical surface properties. Experiments were conducted on monolayer mixtures of a liquid crystalline polymer, poly(p-phenylene)sulfonic acid (PPPSH), with an aliphatic fatty acid, stearic acid (C18). It is found that these two molecules mix ideally, as supported by studies of average molecular area, collapse pressure, and Brewster angle microscopy. Both surface moduli and surface viscosity were obtained with an interfacial stress rheometer and showed that PPPSH/stearic acid monolayer mixtures are non-Newtonian. The orientational dynamics of polymers under well-defined planar extensional flow was considered using UV dichroism, which revealed a surface pressure-induced isotropic-nematic transition. This transition is difficult to observe using traditional isotherm analysis; however, it is readily evident in measurements of the rheology of the monolayers.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Structure and dynamics of magnetorheological fluids in rotating magnetic fields.\n \n \n \n \n\n\n \n Melle, S; Fuller, G.; and Rubio, M A\n\n\n \n\n\n\n Physical Review E, 61(4 B): 4111–4117. 2000.\n \n\n\n\n
\n\n\n\n \n \n \"StructurePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{melle_structure_2000,\n\tAbstract = {We report on the orientation dynamics and aggregation processes of magnetorheological fluids subject to rotating magnetic fields using the technique of scattering dichroism. In the presence of stationary fields we find that the mean length of the field-induced aggregates reaches a saturation value due to finite-size effects. When a rotating field is imposed, we see the chains rotate with the magnetic field frequency (synchronous regime) but with a retarded phase angle for all the rotational frequencies applied. However, two different behaviors are found below or above a critical frequency fc. Within the first regime (low frequency values) the size of the aggregates remains almost constant, while at high frequencies this size becomes shorter due to hydrodynamic drag. Experimental results have been reproduced by a simple model considering a torque balance on the chainlike aggregates.},\n\tAuthor = {Melle, S and Fuller, G.G. and Rubio, M A},\n\tJournal = {Physical Review E},\n\tNumber = {4 B},\n\tPages = {4111--4117},\n\tTitle = {Structure and dynamics of magnetorheological fluids in rotating magnetic fields},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001441540&partnerID=40&md5=1753f1e8142fa9f6ea48f101538a4007},\n\tVolume = {61},\n\tYear = {2000},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001441540&partnerID=40&md5=1753f1e8142fa9f6ea48f101538a4007}}\n\n
\n
\n\n\n
\n We report on the orientation dynamics and aggregation processes of magnetorheological fluids subject to rotating magnetic fields using the technique of scattering dichroism. In the presence of stationary fields we find that the mean length of the field-induced aggregates reaches a saturation value due to finite-size effects. When a rotating field is imposed, we see the chains rotate with the magnetic field frequency (synchronous regime) but with a retarded phase angle for all the rotational frequencies applied. However, two different behaviors are found below or above a critical frequency fc. Within the first regime (low frequency values) the size of the aggregates remains almost constant, while at high frequencies this size becomes shorter due to hydrodynamic drag. Experimental results have been reproduced by a simple model considering a torque balance on the chainlike aggregates.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Birefringence and stress growth in uniaxial extension of polymer solutions.\n \n \n \n \n\n\n \n Sridhar, T; Nguyen, D A; and Fuller, G.\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 90(Amsterdam, Netherlands): 299–315. 2000.\n \n\n\n\n
\n\n\n\n \n \n \"BirefringencePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{sridhar_birefringence_2000,\n\tAbstract = {Simultaneous measurements of extensional stresses and birefringence are rare, especially for polymer solutions. This paper reports such measurements using the filament stretch rheometer and a phase modulated birefringence system. Both the extensional viscosity and the birefringence increase monotonically with strain and reach a plateau. Estimates of this saturation value for birefringence, using Peterlin's formula for birefringence of a fully extended polymer chain are in agreement with the experimental results. However, estimates of the saturation value of the extensional viscosity using Batchelor's formula for suspensions of elongated fibres are much higher than observed. Reasons for the inability of the flow field to fully unravel the polymer chain are examined using published Brownian dynamics simulations. It is tentatively concluded that the polymer chain forms a folded structure. Such folded chains can exhibit saturation in birefringence even though the stress is less than that expected for a fully extended molecule. Simultaneous measurements of stress and birefringence during relaxation indicate that the birefringence decays much more slowly than the stress. The stress-birefringence data show a pronounced hysteresis as predicted by bead-rod models. The failure of the stress optic coefficient in strong flows is noted. Experiments were also performed wherein the strain was increased linearly with time, then held constant for a short period before being increased again. The response of the stress and birefringence in such experiments is dramatically different and can be traced to the different configurations obtained during stretching and relaxation. The results cast doubt on the appropriateness of pre-averaging the non-linear terms in constitutive equations. (C) 2000 Elsevier Science B.V. All rights reserved. Simultaneous measurements of extensional stresses and birefringence are rare, especially for polymer solutions. This paper reports such measurements using the filament stretch rheometer and a phase modulated birefringence system. Both the extensional viscosity and the birefringence increase monotonically with strain and reach a plateau. Estimates of this saturation value for birefringence, using Peterlin's formula for birefringence of a fully extended polymer chain are in agreement with the experimental results. However, estimates of the saturation value of the extensional viscosity using Batchelor's formula for suspensions of elongated fibres are much higher than observed. Reasons for the inability of the flow field to fully unravel the polymer chain are examined using published Brownian dynamics simulations. It is tentatively concluded that the polymer chain forms a folded structure. Such folded chains can exhibit saturation in birefringence even though the stress is less than that expected for a fully extended molecule. Simultaneous measurements of stress and birefringence during relaxation indicate that the birefringence decays much more slowly than the stress. The stress-birefringence data show a pronounced hysteresis as predicted by bead-rod models. The failure of the stress optic coefficient in strong flows is noted. Experiments were also performed wherein the strain was increased linearly with time, then held constant for a short period before being increased again. The response of the stress and birefringence in such experiments is dramatically different and can be traced to the different configurations obtained during stretching and relaxation. The results cast doubt on the appropriateness of pre-averaging the non-linear terms in constitutive equations.},\n\tAuthor = {Sridhar, T and Nguyen, D A and Fuller, G.G.},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {Amsterdam, Netherlands},\n\tPages = {299--315},\n\tTitle = {Birefringence and stress growth in uniaxial extension of polymer solutions},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0034127398&partnerID=40&md5=e4993cde518868220920d0c759ab1b22},\n\tVolume = {90},\n\tYear = {2000},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0034127398&partnerID=40&md5=e4993cde518868220920d0c759ab1b22}}\n\n
\n
\n\n\n
\n Simultaneous measurements of extensional stresses and birefringence are rare, especially for polymer solutions. This paper reports such measurements using the filament stretch rheometer and a phase modulated birefringence system. Both the extensional viscosity and the birefringence increase monotonically with strain and reach a plateau. Estimates of this saturation value for birefringence, using Peterlin's formula for birefringence of a fully extended polymer chain are in agreement with the experimental results. However, estimates of the saturation value of the extensional viscosity using Batchelor's formula for suspensions of elongated fibres are much higher than observed. Reasons for the inability of the flow field to fully unravel the polymer chain are examined using published Brownian dynamics simulations. It is tentatively concluded that the polymer chain forms a folded structure. Such folded chains can exhibit saturation in birefringence even though the stress is less than that expected for a fully extended molecule. Simultaneous measurements of stress and birefringence during relaxation indicate that the birefringence decays much more slowly than the stress. The stress-birefringence data show a pronounced hysteresis as predicted by bead-rod models. The failure of the stress optic coefficient in strong flows is noted. Experiments were also performed wherein the strain was increased linearly with time, then held constant for a short period before being increased again. The response of the stress and birefringence in such experiments is dramatically different and can be traced to the different configurations obtained during stretching and relaxation. The results cast doubt on the appropriateness of pre-averaging the non-linear terms in constitutive equations. (C) 2000 Elsevier Science B.V. All rights reserved. Simultaneous measurements of extensional stresses and birefringence are rare, especially for polymer solutions. This paper reports such measurements using the filament stretch rheometer and a phase modulated birefringence system. Both the extensional viscosity and the birefringence increase monotonically with strain and reach a plateau. Estimates of this saturation value for birefringence, using Peterlin's formula for birefringence of a fully extended polymer chain are in agreement with the experimental results. However, estimates of the saturation value of the extensional viscosity using Batchelor's formula for suspensions of elongated fibres are much higher than observed. Reasons for the inability of the flow field to fully unravel the polymer chain are examined using published Brownian dynamics simulations. It is tentatively concluded that the polymer chain forms a folded structure. Such folded chains can exhibit saturation in birefringence even though the stress is less than that expected for a fully extended molecule. Simultaneous measurements of stress and birefringence during relaxation indicate that the birefringence decays much more slowly than the stress. The stress-birefringence data show a pronounced hysteresis as predicted by bead-rod models. The failure of the stress optic coefficient in strong flows is noted. Experiments were also performed wherein the strain was increased linearly with time, then held constant for a short period before being increased again. The response of the stress and birefringence in such experiments is dramatically different and can be traced to the different configurations obtained during stretching and relaxation. The results cast doubt on the appropriateness of pre-averaging the non-linear terms in constitutive equations.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Phase behavior and flow properties of `hairy-rod' monolayers.\n \n \n \n \n\n\n \n Fischer, P; Brooks, C.; Fuller, G.; Ritcey, A M; Xiao, Y; and Rahem, T\n\n\n \n\n\n\n Langmuir, 16(2): 726–734. 2000.\n \n\n\n\n
\n\n\n\n \n \n \"PhasePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fischer_phase_2000,\n\tAbstract = {Brewster angle microscopy, interfacial stress rheometry, and Π-A isotherm measurements are employed to characterize spread monolayers of two new cellulose derivatives, one containing octadecyl side chains, the other p-nitroaniline substituents. Similar monolayer phase behavior is found in both cases. At high surface area, a liquid phase is observed to coexist with a dilute gaseous phase. Monolayer compression eventually leads to complete coalescence to yield a uniform liquid phase. Surprisingly, a homogeneous phase is observed only at surface pressures significantly higher than that characterizing the gas-liquid phase equilibrium. This observation is attributed to dipolar repulsion between neighboring liquid domains. Compression beyond surface areas at which a continuous monolayer is observed provokes a phase transition, appearing as either a constant pressure plateau or as changes in slope in the surface pressure area isotherms. BAM images reveal only uniform surface films at molecular areas within the transition region. Surface rheological measurements indicate significant changes in monolayer mechanical properties during compression. Importantly, these changes can be correlated with isotherm shape. In all cases, monolayers exhibit a predominantly viscous behavior at high surface areas where two phases coexist. Homogeneous films achieved upon compression exhibit rheological properties that depend on both temperature and the nature of the side chain substituents. In the case of long alkyl side chains (HPC-C18), the uniform film is found to be predominantly viscous at high temperatures, and more elastic at low temperatures. This observation is interpreted as an indication of the partial crystallization of interdigitated side chains, made possible by bilayer formation. Highly dipolar chromophores incorporated as side chain substituents increase interlayer interactions and interdomain repulsion.},\n\tAuthor = {Fischer, P and Brooks, C.F. and Fuller, G.G. and Ritcey, A M and Xiao, Y and Rahem, T},\n\tJournal = {Langmuir},\n\tNumber = {2},\n\tPages = {726--734},\n\tTitle = {Phase behavior and flow properties of `hairy-rod' monolayers},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033639812&partnerID=40&md5=208971ac08dab917eda0654a3ab4a866},\n\tVolume = {16},\n\tYear = {2000},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033639812&partnerID=40&md5=208971ac08dab917eda0654a3ab4a866}}\n\n
\n
\n\n\n
\n Brewster angle microscopy, interfacial stress rheometry, and Π-A isotherm measurements are employed to characterize spread monolayers of two new cellulose derivatives, one containing octadecyl side chains, the other p-nitroaniline substituents. Similar monolayer phase behavior is found in both cases. At high surface area, a liquid phase is observed to coexist with a dilute gaseous phase. Monolayer compression eventually leads to complete coalescence to yield a uniform liquid phase. Surprisingly, a homogeneous phase is observed only at surface pressures significantly higher than that characterizing the gas-liquid phase equilibrium. This observation is attributed to dipolar repulsion between neighboring liquid domains. Compression beyond surface areas at which a continuous monolayer is observed provokes a phase transition, appearing as either a constant pressure plateau or as changes in slope in the surface pressure area isotherms. BAM images reveal only uniform surface films at molecular areas within the transition region. Surface rheological measurements indicate significant changes in monolayer mechanical properties during compression. Importantly, these changes can be correlated with isotherm shape. In all cases, monolayers exhibit a predominantly viscous behavior at high surface areas where two phases coexist. Homogeneous films achieved upon compression exhibit rheological properties that depend on both temperature and the nature of the side chain substituents. In the case of long alkyl side chains (HPC-C18), the uniform film is found to be predominantly viscous at high temperatures, and more elastic at low temperatures. This observation is interpreted as an indication of the partial crystallization of interdigitated side chains, made possible by bilayer formation. Highly dipolar chromophores incorporated as side chain substituents increase interlayer interactions and interdomain repulsion.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Contraction and expansion flows of Langmuir monolayers.\n \n \n \n \n\n\n \n Olson, D. J; and Fuller, G. G\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 89(1): 187–207. 2000.\n \n\n\n\n
\n\n\n\n \n \n \"ContractionPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{olson_contraction_2000,\n\tAuthor = {Olson, David J and Fuller, Gerald G},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {1},\n\tPages = {187--207},\n\tTitle = {Contraction and expansion flows of {Langmuir} monolayers},\n\tUrl = {http://www.sciencedirect.com/science/article/pii/S0377025799000245},\n\tVolume = {89},\n\tYear = {2000},\n\tBdsk-Url-1 = {http://www.sciencedirect.com/science/article/pii/S0377025799000245}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1999\n \n \n (10)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Viscoelastic properties of lipopolymers at the air-water interface: A combined interfacial stress rheometer and film balance study.\n \n \n \n\n\n \n Naumann, C A; Brooks, C.; Fuller, G.; Knoll, W; and Frank, C.\n\n\n \n\n\n\n Langmuir, 15(Washington, DC, United States): 7752–7761. 1999.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{naumann_viscoelastic_1999,\n\tAbstract = {Poly(ethylene glycol) (PEG) is a molecule that exhibits unique behavior when compared with polymers in its homologous family. Depending on its environment, it may show hydrophilic, hydrophobic, or amphiphilic properties. We have studied several PEG lipopolymers, where a PEG chain with a molecular weight (MW) of 2000 g/mol or 5000 g/mol is covalently attached to 1,2-dipalmitoyl- or 1,2-distearoyl-sn-glycero-3-phosphoethanolamine, with a Langmuir film balance and a recently developed interfacial stress rheometer. In particular, we have determined how the rheological properties of PEG molecules anchored at the air-water interface change when the polymer chains are forced into highly stretched brush conformations. Pressure-area isotherms of monolayers of PEG lipopolymers exhibit two phase transitions: a desorption transition of the PEG chains from the air-water interface at 10 mN/m and a high film pressure transition at 20-40 mN/m, but the nature of the latter transition is still poorly understood. We have observed a remarkable change of the viscoelastic properties in the range of the high-pressure transition. The monolayer is fluid below the transition, with the surface loss modulus, Gtextacutedbls, being larger than the surface storage modulus, Gtextasciiacutes, but becomes remarkably elastic above, with Gtextasciiacutes\\{\\vphantom{\\}}\\&},\n\tAuthor = {Naumann, C A and Brooks, C.F. and Fuller, G.G. and Knoll, W and Frank, C.W.},\n\tJournal = {Langmuir},\n\tNumber = {Washington, DC, United States},\n\tPages = {7752--7761},\n\tTitle = {Viscoelastic properties of lipopolymers at the air-water interface: {A} combined interfacial stress rheometer and film balance study},\n\tVolume = {15},\n\tYear = {1999}}\n\n
\n
\n\n\n
\n Poly(ethylene glycol) (PEG) is a molecule that exhibits unique behavior when compared with polymers in its homologous family. Depending on its environment, it may show hydrophilic, hydrophobic, or amphiphilic properties. We have studied several PEG lipopolymers, where a PEG chain with a molecular weight (MW) of 2000 g/mol or 5000 g/mol is covalently attached to 1,2-dipalmitoyl- or 1,2-distearoyl-sn-glycero-3-phosphoethanolamine, with a Langmuir film balance and a recently developed interfacial stress rheometer. In particular, we have determined how the rheological properties of PEG molecules anchored at the air-water interface change when the polymer chains are forced into highly stretched brush conformations. Pressure-area isotherms of monolayers of PEG lipopolymers exhibit two phase transitions: a desorption transition of the PEG chains from the air-water interface at 10 mN/m and a high film pressure transition at 20-40 mN/m, but the nature of the latter transition is still poorly understood. We have observed a remarkable change of the viscoelastic properties in the range of the high-pressure transition. The monolayer is fluid below the transition, with the surface loss modulus, Gtextacutedbls, being larger than the surface storage modulus, Gtextasciiacutes, but becomes remarkably elastic above, with Gtextasciiacutes\\p̌hantom\\&\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Interfacial stress rheometer to study rheological transitions in monolayers at the air-water interface.\n \n \n \n \n\n\n \n Brooks, C.; Fuller, G.; Frank, C.; and Robertson, C.\n\n\n \n\n\n\n Langmuir, 15(7): 2450–2459. 1999.\n \n\n\n\n
\n\n\n\n \n \n \"InterfacialPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{brooks_interfacial_1999,\n\tAuthor = {Brooks, C.F. and Fuller, G.G. and Frank, C.W. and Robertson, C.R.},\n\tJournal = {Langmuir},\n\tNumber = {7},\n\tPages = {2450--2459},\n\tTitle = {Interfacial stress rheometer to study rheological transitions in monolayers at the air-water interface},\n\tUrl = {http://pubs.acs.org/doi/abs/10.1021/la980465r},\n\tVolume = {15},\n\tYear = {1999},\n\tBdsk-Url-1 = {http://pubs.acs.org/doi/abs/10.1021/la980465r}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Transient birefringence of elastomeric polypropylene subjected to step shear strain.\n \n \n \n \n\n\n \n Carlson, E D; Fuller, G.; and Waymouth, R M\n\n\n \n\n\n\n Macromolecules, 32(Washington, DC, United States): 8094–8099. 1999.\n \n\n\n\n
\n\n\n\n \n \n \"TransientPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{carlson_transient_1999,\n\tAbstract = {The step shear strain behavior of elastomeric polypropylene (ePP) synthesized from an unbridged metallocene catalyst is studied using polarimetry. The elastomeric nature of ePP is presumed to arise from a multiblock structure of isotactic (iPP) and atactic (aPP) polypropylene blocks. At lower temperatures stereoregular blocks of iPP are able to crystallize while stereoirregular blocks of aPP remain amorphous. The flow behavior of ePP is compared with homopolymer blends of iPP and aPP that have the same isotactic content as ePP as well as solvent fractions of the parent ePP sample. Step shear strain experiments carried out on crystallized samples show that ePP does not completely relax from an applied strain and that a cross-linked network has formed. Imperfections in the network structure are revealed by birefringence measurements that show a partial relaxation of the polymer orientation. In contrast, crystallized homopolymer blend samples are able to completely relax after the application of a step shear strain. Step shear strain experiments of the solvent fractions of ePP suggest that the elastomeric behavior results from a combination of fractions that are able to cocrystallize to form a physically cross-linked network.},\n\tAuthor = {Carlson, E D and Fuller, G.G. and Waymouth, R M},\n\tJournal = {Macromolecules},\n\tNumber = {Washington, DC, United States},\n\tPages = {8094--8099},\n\tTitle = {Transient birefringence of elastomeric polypropylene subjected to step shear strain},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033309413&partnerID=40&md5=a7756533c497be6c1b2ff02062c71509},\n\tVolume = {32},\n\tYear = {1999},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033309413&partnerID=40&md5=a7756533c497be6c1b2ff02062c71509}}\n\n
\n
\n\n\n
\n The step shear strain behavior of elastomeric polypropylene (ePP) synthesized from an unbridged metallocene catalyst is studied using polarimetry. The elastomeric nature of ePP is presumed to arise from a multiblock structure of isotactic (iPP) and atactic (aPP) polypropylene blocks. At lower temperatures stereoregular blocks of iPP are able to crystallize while stereoirregular blocks of aPP remain amorphous. The flow behavior of ePP is compared with homopolymer blends of iPP and aPP that have the same isotactic content as ePP as well as solvent fractions of the parent ePP sample. Step shear strain experiments carried out on crystallized samples show that ePP does not completely relax from an applied strain and that a cross-linked network has formed. Imperfections in the network structure are revealed by birefringence measurements that show a partial relaxation of the polymer orientation. In contrast, crystallized homopolymer blend samples are able to completely relax after the application of a step shear strain. Step shear strain experiments of the solvent fractions of ePP suggest that the elastomeric behavior results from a combination of fractions that are able to cocrystallize to form a physically cross-linked network.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Large-scale bundle ordering in sterically stabilized latices.\n \n \n \n\n\n \n Vermant, J; Raynaud, L; Mewis, J; Ernst, B; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 211(2): 221–229. 1999.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{vermant_large-scale_1999,\n\tAuthor = {Vermant, J and Raynaud, L and Mewis, J and Ernst, B and Fuller, Gg},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {2},\n\tPages = {221--229},\n\tTitle = {Large-scale bundle ordering in sterically stabilized latices},\n\tVolume = {211},\n\tYear = {1999}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Elastomeric polypropylenes from unbridged 2-phenylindene zirconocene catalysts: temperature dependence of crystallinity and relaxation properties.\n \n \n \n\n\n \n Hu, Y; Carlson, E D; Fuller, G.; and Waymouth, R M\n\n\n \n\n\n\n Macromolecules, 32(Washington, DC, United States): 3334–3340. 1999.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hu_elastomeric_1999,\n\tAuthor = {Hu, Y and Carlson, E D and Fuller, G.G. and Waymouth, R M},\n\tJournal = {Macromolecules},\n\tNumber = {Washington, DC, United States},\n\tPages = {3334--3340},\n\tTitle = {Elastomeric polypropylenes from unbridged 2-phenylindene zirconocene catalysts: temperature dependence of crystallinity and relaxation properties},\n\tVolume = {32},\n\tYear = {1999}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Electric field-induced component dynamics in a binary liquid crystal mixture studied using two-dimensional Raman scattering.\n \n \n \n\n\n \n Huang, K; and Fuller, G.\n\n\n \n\n\n\n Liquid Crystals, 26(1): 1–7. 1999.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{huang_electric_1999,\n\tAbstract = {The intermolecular interactions in a mixed, binary liquid crystal system are probed using two-dimensional Raman scattering. The sample is a 50/50 by weight blend of 4-pentyl-(4-cyanophenyl)cyclohexane (PCH5) and 4-methoxybenzylidene-4-butylaniline (MBBA). The response of the Raman anisotropy of the individual components to step voltages is identical. Furthermore, cross correlation analysis of oscillatory electric field experiments shows only synchronous reorientation between the two constituents. Since neat MBBA aligns perpendicular and neat CH5 aligns parallel to the applied field, the data imply strong intermolecular interactions. Additionally, as previously observed for PCH5, frequency sweep experiments indicate that the flexible parts of the molecules can reorient independently of the rigid cores. textcopyright1999 Taylor \\{\\vphantom{\\}}\\&},\n\tAuthor = {Huang, K and Fuller, G.G.},\n\tJournal = {Liquid Crystals},\n\tNumber = {1},\n\tPages = {1--7},\n\tTitle = {Electric field-induced component dynamics in a binary liquid crystal mixture studied using two-dimensional {Raman} scattering},\n\tVolume = {26},\n\tYear = {1999}}\n\n
\n
\n\n\n
\n The intermolecular interactions in a mixed, binary liquid crystal system are probed using two-dimensional Raman scattering. The sample is a 50/50 by weight blend of 4-pentyl-(4-cyanophenyl)cyclohexane (PCH5) and 4-methoxybenzylidene-4-butylaniline (MBBA). The response of the Raman anisotropy of the individual components to step voltages is identical. Furthermore, cross correlation analysis of oscillatory electric field experiments shows only synchronous reorientation between the two constituents. Since neat MBBA aligns perpendicular and neat CH5 aligns parallel to the applied field, the data imply strong intermolecular interactions. Additionally, as previously observed for PCH5, frequency sweep experiments indicate that the flexible parts of the molecules can reorient independently of the rigid cores. textcopyright1999 Taylor \\p̌hantom\\&\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Characterization of the flow properties of sodium carboxymethylcellulose via mechanical and optical techniques.\n \n \n \n \n\n\n \n Kulicke, W M; Reinhardt, U; Fuller, G.; and Arendt, O\n\n\n \n\n\n\n Rheologica Acta, 38(1): 26–33. 1999.\n \n\n\n\n
\n\n\n\n \n \n \"CharacterizationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kulicke_characterization_1999,\n\tAbstract = {Sodium carboxymethylcellulose (NaCMC) in solution represents a complex rheological system, since it forms aggregates and associations and hence higher-level structures and, depending on the synthesis, is only found in a molecularly dispersed form in exceptional cases. Rheo-mechanical investigations of the viscoelasticity showed that the Cox-Merz rule is not fulfilled. The aim was therefore to examine whether rheo-optics could be employed to provide more detailed conclusions about the parameters that influence the flow behavior of NaCMC than has hitherto been available with mechanical methods. The flow birefringence, Δn', rises as the degree of polymerization increases, and exhibits the same dependence on molar mass as does the viscosity: Δn' Proportional to MW3.4. As the degree of polymerization increases while the shear rate remains constant, the polymer segments become more distinctly aligned in the direction of shear. Hence increasing the degree of polymerization also affects the solution structure, i.e. the interaction of the molecules with one another. The stress-optical rule only applies to a limited extent for this system. The stress-optical coefficient, C, is almost independent of the shear rate, but is strongly influenced by the concentration and attains a limiting value of 3 x 10-8 Pa-1. C was determined for a polymer in dilute solution and the curve obtained also enabled transitions in the solution structure to be recognized.},\n\tAuthor = {Kulicke, W M and Reinhardt, U and Fuller, G.G. and Arendt, O},\n\tJournal = {Rheologica Acta},\n\tNumber = {1},\n\tPages = {26--33},\n\tTitle = {Characterization of the flow properties of sodium carboxymethylcellulose via mechanical and optical techniques},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032770977&partnerID=40&md5=8e0f5fe288f64a25b6b9bde053962941},\n\tVolume = {38},\n\tYear = {1999},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032770977&partnerID=40&md5=8e0f5fe288f64a25b6b9bde053962941}}\n\n
\n
\n\n\n
\n Sodium carboxymethylcellulose (NaCMC) in solution represents a complex rheological system, since it forms aggregates and associations and hence higher-level structures and, depending on the synthesis, is only found in a molecularly dispersed form in exceptional cases. Rheo-mechanical investigations of the viscoelasticity showed that the Cox-Merz rule is not fulfilled. The aim was therefore to examine whether rheo-optics could be employed to provide more detailed conclusions about the parameters that influence the flow behavior of NaCMC than has hitherto been available with mechanical methods. The flow birefringence, Δn', rises as the degree of polymerization increases, and exhibits the same dependence on molar mass as does the viscosity: Δn' Proportional to MW3.4. As the degree of polymerization increases while the shear rate remains constant, the polymer segments become more distinctly aligned in the direction of shear. Hence increasing the degree of polymerization also affects the solution structure, i.e. the interaction of the molecules with one another. The stress-optical rule only applies to a limited extent for this system. The stress-optical coefficient, C, is almost independent of the shear rate, but is strongly influenced by the concentration and attains a limiting value of 3 x 10-8 Pa-1. C was determined for a polymer in dilute solution and the curve obtained also enabled transitions in the solution structure to be recognized.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Component relaxation processes within elastomeric polypropylene.\n \n \n \n \n\n\n \n Carlson, E D; Fuller, G.; and Waymouth, R M\n\n\n \n\n\n\n Macromolecules, 32(Washington, DC, United States): 8100–8106. 1999.\n \n\n\n\n
\n\n\n\n \n \n \"ComponentPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{carlson_component_1999,\n\tAbstract = {The behavior of specific components of elastomeric polypropylene was directly observed in situ using a dynamic, infrared polarimetry technique. The elastomeric nature of ePP is presumed to arise from a multiblock structure of isotactic (iPP) and atactic (aPP) polypropylene blocks. Elastomeric polypropylene is a heterogeneous material in terms of tacticity that can be separated into fractions differing in tacticity. A series of samples were prepared in which one fraction was labeled with deuterium. Infrared dichroism measurements made at the C-2H stretching vibration were used to follow the relaxation behavior of fractions of various tacticity. The samples were subjected to a series of step shear strain experiments at a variety of temperatures from the melt and into the crystalline region. By simultaneously measuring birefringence (which measures the bulk sample's local orientation) and IR dichroism (which measures the orientation dynamics of the deuterated fraction), it is possible to isolate the behavior of specific populations of chains. It was revealed that chains with the highest level of tacticity crystallize first, as is expected, and strongly influence the subsequent crystallization of the remaining chains. In addition, evidence of cocrystallization between the solvent fractions was observed.},\n\tAuthor = {Carlson, E D and Fuller, G.G. and Waymouth, R M},\n\tJournal = {Macromolecules},\n\tNumber = {Washington, DC, United States},\n\tPages = {8100--8106},\n\tTitle = {Component relaxation processes within elastomeric polypropylene},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033352738&partnerID=40&md5=5995649eff6f63c56e7be23c247e398c},\n\tVolume = {32},\n\tYear = {1999},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033352738&partnerID=40&md5=5995649eff6f63c56e7be23c247e398c}}\n\n
\n
\n\n\n
\n The behavior of specific components of elastomeric polypropylene was directly observed in situ using a dynamic, infrared polarimetry technique. The elastomeric nature of ePP is presumed to arise from a multiblock structure of isotactic (iPP) and atactic (aPP) polypropylene blocks. Elastomeric polypropylene is a heterogeneous material in terms of tacticity that can be separated into fractions differing in tacticity. A series of samples were prepared in which one fraction was labeled with deuterium. Infrared dichroism measurements made at the C-2H stretching vibration were used to follow the relaxation behavior of fractions of various tacticity. The samples were subjected to a series of step shear strain experiments at a variety of temperatures from the melt and into the crystalline region. By simultaneously measuring birefringence (which measures the bulk sample's local orientation) and IR dichroism (which measures the orientation dynamics of the deuterated fraction), it is possible to isolate the behavior of specific populations of chains. It was revealed that chains with the highest level of tacticity crystallize first, as is expected, and strongly influence the subsequent crystallization of the remaining chains. In addition, evidence of cocrystallization between the solvent fractions was observed.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n The rheology of two-dimensional systems.\n \n \n \n\n\n \n Fuller, G. G; Yim, K; Brooks, C. F; Olson, D; and Frank, C.\n\n\n \n\n\n\n Korea-Australia Rheology Journal, 11(4): 321–328. 1999.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_rheology_1999,\n\tAbstract = {This paper discusses the rheology of complex interfaces comprised of amphiphilic materials that are sus- ceptible to flow-induced orientation and deformation. The consequence of the coupling of the film micro- structure to flow leads to nonlinear rheology and surface fluid dynamics. Experimental methods designed to determine the mechanical rheological material functions of fluid-fluid interfaces as well as local, molec- ular and morphological responses are presented. These include a newly developed interfacial stress rhe- ometer, flow ultraviolet dichroism, and Brewster-angle microscopy. These techniques are applied to a number of complex interfaces ranging from low molecular weight amphiphiles to polymer monolayers. Nonlinear flow phenomena ranging from two-dimensional nematic responses to highly elastic surface flows that manifest surface normal stress differences and elongational viscosities are described.},\n\tAuthor = {Fuller, Gerald G and Yim, K and Brooks, Carlton F and Olson, D and Frank, C.W.},\n\tJournal = {Korea-Australia Rheology Journal},\n\tNumber = {4},\n\tPages = {321--328},\n\tTitle = {The rheology of two-dimensional systems},\n\tVolume = {11},\n\tYear = {1999}}\n\n
\n
\n\n\n
\n This paper discusses the rheology of complex interfaces comprised of amphiphilic materials that are sus- ceptible to flow-induced orientation and deformation. The consequence of the coupling of the film micro- structure to flow leads to nonlinear rheology and surface fluid dynamics. Experimental methods designed to determine the mechanical rheological material functions of fluid-fluid interfaces as well as local, molec- ular and morphological responses are presented. These include a newly developed interfacial stress rhe- ometer, flow ultraviolet dichroism, and Brewster-angle microscopy. These techniques are applied to a number of complex interfaces ranging from low molecular weight amphiphiles to polymer monolayers. Nonlinear flow phenomena ranging from two-dimensional nematic responses to highly elastic surface flows that manifest surface normal stress differences and elongational viscosities are described.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Dynamic light scattering during shear: Measurements of diffusion coefficients.\n \n \n \n \n\n\n \n Rusu, D; Genoe, D; Van Puyvelde, P; Peuvrel-Disdier, E; Navard, P; and Fuller, G.\n\n\n \n\n\n\n Polymer, 40(Exeter, United Kingdom): 1353–1357. 1999.\n \n\n\n\n
\n\n\n\n \n \n \"DynamicPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{rusu_dynamic_1999,\n\tAbstract = {A method for determining diffusion coefficients during shear by dynamic light scattering is presented. The direct determination of diffusion coefficient during flow is hampered by the domination of the convection terms over the diffusion terms except in a very narrow azimuthal scattering angle range that is impossible to reach experimentally. The method described here offers the advantage to this constraint. The technique is based on a double measurement of homodyne spectrum of light scattered from small scatterers immersed in a medium subjected to a simple shear flow. This can be applied to polymer solutions. Verification is provided using suspensions of polystyrene spheres in Newtonian media and a transparent rheometer. A method for determining diffusion coefficients during shear by dynamic light scattering is presented. The direct determination of diffusion coefficient during flow is hampered by the domination of the convection terms over the diffusion terms except in a very narrow azimuthal scattering angle range that is impossible to reach experimentally. The method described here offers the advantage to this constraint. The technique is based on a double measurement of homodyne spectrum of light scattered from small scatterers immersed in a medium subjected to a simple shear flow. This can be applied to polymer solutions. Verification is provided using suspensions of polystyrene spheres in Newtonian media and a transparent rheometer.},\n\tAuthor = {Rusu, D and Genoe, D and Van Puyvelde, P and Peuvrel-Disdier, E and Navard, P and Fuller, G.G.},\n\tJournal = {Polymer},\n\tNumber = {Exeter, United Kingdom},\n\tPages = {1353--1357},\n\tTitle = {Dynamic light scattering during shear: {Measurements} of diffusion coefficients},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033102087&partnerID=40&md5=68ffcd7b2fc07f151cebe64e53100589},\n\tVolume = {40},\n\tYear = {1999},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0033102087&partnerID=40&md5=68ffcd7b2fc07f151cebe64e53100589}}\n\n
\n
\n\n\n
\n A method for determining diffusion coefficients during shear by dynamic light scattering is presented. The direct determination of diffusion coefficient during flow is hampered by the domination of the convection terms over the diffusion terms except in a very narrow azimuthal scattering angle range that is impossible to reach experimentally. The method described here offers the advantage to this constraint. The technique is based on a double measurement of homodyne spectrum of light scattered from small scatterers immersed in a medium subjected to a simple shear flow. This can be applied to polymer solutions. Verification is provided using suspensions of polystyrene spheres in Newtonian media and a transparent rheometer. A method for determining diffusion coefficients during shear by dynamic light scattering is presented. The direct determination of diffusion coefficient during flow is hampered by the domination of the convection terms over the diffusion terms except in a very narrow azimuthal scattering angle range that is impossible to reach experimentally. The method described here offers the advantage to this constraint. The technique is based on a double measurement of homodyne spectrum of light scattered from small scatterers immersed in a medium subjected to a simple shear flow. This can be applied to polymer solutions. Verification is provided using suspensions of polystyrene spheres in Newtonian media and a transparent rheometer.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1998\n \n \n (7)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Flow-Induced Deformation and Relaxation Processes of Polydomain Structures in Langmuir Monolayer.\n \n \n \n \n\n\n \n Yim, K S; Brooks, C.; Fuller, G.; Frank, C.; and Robertson, C.\n\n\n \n\n\n\n 1998.\n \n\n\n\n
\n\n\n\n \n \n \"Flow-InducedPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@book{yim_flow-induced_1998,\n\tAbstract = {Flow-induced deformation and relaxation processes of monolayer domains at the air-water interface have been studied by using simple shear flow and extensional flow. It is found from fluorescence microscopy that deformed domains of dipalmitoyl-phosphatidylcholine (DPPC) tend to relax to circular shapes due to the line tension effects. By examining this relaxation process at low surface pressure, the value of the line tension can be calculated. In the limit of low surface pressure, the line tension was found to be λ = 1.11 textpm 0.20 texttimes 10 -12 N, which is of the same order as values found for other monolayers. A surface Capillary number Cas = ηbγ̇R2/λ is proposed to explain the two-dimensional deformation process in systems dominated by the subphase viscosity, ηb. R is the undeformed domain radius and γ̇ is the velocity gradient. For slightly deformed domains, the deformation D and orientation angle varphi are linear functions of Cas in two dimensions, and extensional flow is twice as effective in deforming domains as simple shear flow, which is similar to the case of three-dimensional droplets.},\n\tAuthor = {Yim, K S and Brooks, C.F. and Fuller, G.G. and Frank, C.W. and Robertson, C.R.},\n\tTitle = {Flow-{Induced} {Deformation} and {Relaxation} {Processes} of {Polydomain} {Structures} in {Langmuir} {Monolayer}},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0346072387&partnerID=40&md5=5d90d4108559704539b5062652a63a82},\n\tYear = {1998},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0346072387&partnerID=40&md5=5d90d4108559704539b5062652a63a82}}\n\n
\n
\n\n\n
\n Flow-induced deformation and relaxation processes of monolayer domains at the air-water interface have been studied by using simple shear flow and extensional flow. It is found from fluorescence microscopy that deformed domains of dipalmitoyl-phosphatidylcholine (DPPC) tend to relax to circular shapes due to the line tension effects. By examining this relaxation process at low surface pressure, the value of the line tension can be calculated. In the limit of low surface pressure, the line tension was found to be λ = 1.11 textpm 0.20 texttimes 10 -12 N, which is of the same order as values found for other monolayers. A surface Capillary number Cas = ηbγ̇R2/λ is proposed to explain the two-dimensional deformation process in systems dominated by the subphase viscosity, ηb. R is the undeformed domain radius and γ̇ is the velocity gradient. For slightly deformed domains, the deformation D and orientation angle varphi are linear functions of Cas in two dimensions, and extensional flow is twice as effective in deforming domains as simple shear flow, which is similar to the case of three-dimensional droplets.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rheological and thermal properties of elastomeric polypropylene.\n \n \n \n \n\n\n \n Carlson, E D; Krejchi, M T; Shah, C D; Terakawa, T; Waymouth, R M; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 31(16): 5343–5351. 1998.\n \n\n\n\n
\n\n\n\n \n \n \"RheologicalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{carlson_rheological_1998,\n\tAbstract = {The properties of elastomeric polypropylene synthesized from an unbridged metallocene catalyst are investigated by mechanical rheometry, differential scanning calorimetry, and optical depolarization. The material is compositionally heterogeneous in tacticity, and its behavior is compared to that of its solvent fractions. The behavior of elastomeric polypropylene is also compared with tacticity-matched atactic/isotactic homopolymer blends. For T below Tm, mechanical rheometry shows that elastomeric polypropylene forms a more constrained network than tacticity-matched homopolymer blends. Differential scanning calorimetry reveals that the elastomeric samples crystallize at a lower temperature than blends composed of the same isotactic content. Depolarization studies show similar transitions but reveal the presence of an additional slower time scale involved with crystallizing the elastomeric sample. Data are consistent with the presumed atactic/isotactic multiblock microstructure of elastomeric polypropylene.},\n\tAuthor = {Carlson, E D and Krejchi, M T and Shah, C D and Terakawa, T and Waymouth, R M and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {16},\n\tPages = {5343--5351},\n\tTitle = {Rheological and thermal properties of elastomeric polypropylene},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032139975&partnerID=40&md5=1972b068287a467b5ff42b16cc67163e},\n\tVolume = {31},\n\tYear = {1998},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032139975&partnerID=40&md5=1972b068287a467b5ff42b16cc67163e}}\n\n
\n
\n\n\n
\n The properties of elastomeric polypropylene synthesized from an unbridged metallocene catalyst are investigated by mechanical rheometry, differential scanning calorimetry, and optical depolarization. The material is compositionally heterogeneous in tacticity, and its behavior is compared to that of its solvent fractions. The behavior of elastomeric polypropylene is also compared with tacticity-matched atactic/isotactic homopolymer blends. For T below Tm, mechanical rheometry shows that elastomeric polypropylene forms a more constrained network than tacticity-matched homopolymer blends. Differential scanning calorimetry reveals that the elastomeric samples crystallize at a lower temperature than blends composed of the same isotactic content. Depolarization studies show similar transitions but reveal the presence of an additional slower time scale involved with crystallizing the elastomeric sample. Data are consistent with the presumed atactic/isotactic multiblock microstructure of elastomeric polypropylene.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Time-periodic flow induced structures and instabilities in a viscoelastic surfactant solution.\n \n \n \n \n\n\n \n Wheeler, E K; Fischer, P; and Fuller, G.\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 75(2-3): 193–208. 1998.\n \n\n\n\n
\n\n\n\n \n \n \"Time-periodicPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wheeler_time-periodic_1998,\n\tAbstract = {Using various optical and mechanical techniques we report on a shear induced structure (SIS) that results in instabilities during flow. The solution in this study is an equimolar solution of cetylpyridinium chloride and sodium salicylate. In this investigation we further probe the hypothesis of Rehage et al. that an SIS occurs [1]. We find that not only does an SIS occur but as the solution is subjected to stronger flows it becomes turbid before forming ring-like structures which alternate in intensity both spatially and temporally. Since the rings alternate periodically it is evident that the flow not only forms these structures but is also capable of destroying them as well. Simultaneously measuring birefringence and mechanical properties we observe the same periodic oscillations in the birefringence. On the other hand these ring-like structures are present in the parallel plate, cone and plate, and Taylor Couette flow cells. These rings are similar in appearance to those seen for elastic instabilities in Boger fluids seeded with mica flakes. In a constant stress rheometer the shear rate exhibits overshoots and fluctuations again resembling the behavior of elastic instabilities in Boger fluids. textcopyright 1998 Elsevier Science B.V. All rights reserved.},\n\tAuthor = {Wheeler, E K and Fischer, P and Fuller, G.G.},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {2-3},\n\tPages = {193--208},\n\tTitle = {Time-periodic flow induced structures and instabilities in a viscoelastic surfactant solution},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032029608&partnerID=40&md5=bd8a03366ac09cacb9bf3d26b3c0350a},\n\tVolume = {75},\n\tYear = {1998},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032029608&partnerID=40&md5=bd8a03366ac09cacb9bf3d26b3c0350a}}\n\n
\n
\n\n\n
\n Using various optical and mechanical techniques we report on a shear induced structure (SIS) that results in instabilities during flow. The solution in this study is an equimolar solution of cetylpyridinium chloride and sodium salicylate. In this investigation we further probe the hypothesis of Rehage et al. that an SIS occurs [1]. We find that not only does an SIS occur but as the solution is subjected to stronger flows it becomes turbid before forming ring-like structures which alternate in intensity both spatially and temporally. Since the rings alternate periodically it is evident that the flow not only forms these structures but is also capable of destroying them as well. Simultaneously measuring birefringence and mechanical properties we observe the same periodic oscillations in the birefringence. On the other hand these ring-like structures are present in the parallel plate, cone and plate, and Taylor Couette flow cells. These rings are similar in appearance to those seen for elastic instabilities in Boger fluids seeded with mica flakes. In a constant stress rheometer the shear rate exhibits overshoots and fluctuations again resembling the behavior of elastic instabilities in Boger fluids. textcopyright 1998 Elsevier Science B.V. All rights reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Two-dimensional Raman study of the submolecular, electric field-induced reorientation of a nematic liquid crystal.\n \n \n \n \n\n\n \n Huang, K; and Fuller, G.\n\n\n \n\n\n\n Liquid Crystals, 25(6): 745–755. 1998.\n \n\n\n\n
\n\n\n\n \n \n \"Two-dimensionalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{huang_two-dimensional_1998,\n\tAbstract = {Two-dimensional Raman scattering is presented as a technique for the monitoring of electric field-induced, submolecular reorientation in liquid crystals. The motions of the flexible part and the rigid core of 4-pentyl-(4-cyanophenyl)cyclohexane (PCH5) are independently monitored in response to both step and oscillatory electric fields. Step voltage experiments show that the flexible group reorients before the rigid core. Also, oscillatory electric field experiments demonstrate that the flexible and rigid groups reorient asynchronously. In fact, at periodicities that are shorter than the bulk reorientation times, it is observed that the reorientation of the flexible part is amplified, while the motion of the rigid core is inhibited. The data suggest that the flexible group possesses a small, local dielectric anisotropy that can couple with the electric field to induce an independent, cooperative reorientation when the mobility of the rigid core is restricted.},\n\tAuthor = {Huang, K and Fuller, G.G.},\n\tJournal = {Liquid Crystals},\n\tNumber = {6},\n\tPages = {745--755},\n\tTitle = {Two-dimensional {Raman} study of the submolecular, electric field-induced reorientation of a nematic liquid crystal},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000028458&partnerID=40&md5=2c4618a3d0da2afca6a4e25c1335d5da},\n\tVolume = {25},\n\tYear = {1998},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000028458&partnerID=40&md5=2c4618a3d0da2afca6a4e25c1335d5da}}\n\n
\n
\n\n\n
\n Two-dimensional Raman scattering is presented as a technique for the monitoring of electric field-induced, submolecular reorientation in liquid crystals. The motions of the flexible part and the rigid core of 4-pentyl-(4-cyanophenyl)cyclohexane (PCH5) are independently monitored in response to both step and oscillatory electric fields. Step voltage experiments show that the flexible group reorients before the rigid core. Also, oscillatory electric field experiments demonstrate that the flexible and rigid groups reorient asynchronously. In fact, at periodicities that are shorter than the bulk reorientation times, it is observed that the reorientation of the flexible part is amplified, while the motion of the rigid core is inhibited. The data suggest that the flexible group possesses a small, local dielectric anisotropy that can couple with the electric field to induce an independent, cooperative reorientation when the mobility of the rigid core is restricted.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The dynamics of two dimensional polymer nematics.\n \n \n \n \n\n\n \n Maruyama, T; Fuller, G.; Grosso, M; and Maffettone, P L\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 76(1-3): 233–247. 1998.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{maruyama_dynamics_1998,\n\tAbstract = {The orientation dynamics of a monolayer of a nematic polymer solution were examined for both extensional and simple shear flows. In extensional flow, a 'strong flow/weak flow' criteria was investigated and found to occur at a critical rate of strain that was predicted by the simple molecular model of Doi and Hess. In simple shear flow, evidence of director wagging and flow alignment was obtained. textcopyright 1998 Elsevier Science B.V. All rights reserved.},\n\tAuthor = {Maruyama, T and Fuller, G.G. and Grosso, M and Maffettone, P L},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {1-3},\n\tPages = {233--247},\n\tTitle = {The dynamics of two dimensional polymer nematics},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032048959&partnerID=40&md5=b1bb9720fb4668bd117658698bf6a0b1},\n\tVolume = {76},\n\tYear = {1998},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032048959&partnerID=40&md5=b1bb9720fb4668bd117658698bf6a0b1}}\n\n
\n
\n\n\n
\n The orientation dynamics of a monolayer of a nematic polymer solution were examined for both extensional and simple shear flows. In extensional flow, a 'strong flow/weak flow' criteria was investigated and found to occur at a critical rate of strain that was predicted by the simple molecular model of Doi and Hess. In simple shear flow, evidence of director wagging and flow alignment was obtained. textcopyright 1998 Elsevier Science B.V. All rights reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Anisotropy and orientation of the microstructure in viscous emulsions during shear flow.\n \n \n \n \n\n\n \n Vermant, J; Van Puyvelde, P; Moldenaers, P; Mewis, J; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 14(7): 1612–1617. 1998.\n \n\n\n\n
\n\n\n\n \n \n \"AnisotropyPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{vermant_anisotropy_1998,\n\tAbstract = {Flow small-angle light scattering and linear conservative dichroism are used to follow, in situ and time resolved, the flow-induced changes of the microstructure in viscous emulsions such as immiscible polymer blends. A dilute system consisting of poly (butadiene) droplets dispersed in a poly(isobutene) matrix has been used as a model system. Contrary to earlier rheo-optical work on such materials, the structure has been probed in the plane formed by the flow and the velocity gradient directions. In this manner, shape anisotropy as well as the orientation angle can be monitored continuously for droplet sizes where microscopic observation becomes difficult or even impossible. During steady-state shear flow the size-dependent orientation of the droplets could be detected. It is shown that the various morphological stages caused by a sudden increase in shear rate can be identified and quantified: droplet deformation and rotation, fibril formation, and development of interfacial instabilities leading to breakup and reorientation of the resulting small droplets. These latter stages result in very characteristic light-scattering patterns. Values for anisotropy and orientation are derived from dichroism and light-scattering data and compared with theoretical results for the case of affine droplet deformation.},\n\tAuthor = {Vermant, J and Van Puyvelde, P and Moldenaers, P and Mewis, J and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {7},\n\tPages = {1612--1617},\n\tTitle = {Anisotropy and orientation of the microstructure in viscous emulsions during shear flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032028802&partnerID=40&md5=e261d5e84f48e6ebb13f336dac69c32a},\n\tVolume = {14},\n\tYear = {1998},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032028802&partnerID=40&md5=e261d5e84f48e6ebb13f336dac69c32a}}\n\n
\n
\n\n\n
\n Flow small-angle light scattering and linear conservative dichroism are used to follow, in situ and time resolved, the flow-induced changes of the microstructure in viscous emulsions such as immiscible polymer blends. A dilute system consisting of poly (butadiene) droplets dispersed in a poly(isobutene) matrix has been used as a model system. Contrary to earlier rheo-optical work on such materials, the structure has been probed in the plane formed by the flow and the velocity gradient directions. In this manner, shape anisotropy as well as the orientation angle can be monitored continuously for droplet sizes where microscopic observation becomes difficult or even impossible. During steady-state shear flow the size-dependent orientation of the droplets could be detected. It is shown that the various morphological stages caused by a sudden increase in shear rate can be identified and quantified: droplet deformation and rotation, fibril formation, and development of interfacial instabilities leading to breakup and reorientation of the resulting small droplets. These latter stages result in very characteristic light-scattering patterns. Values for anisotropy and orientation are derived from dichroism and light-scattering data and compared with theoretical results for the case of affine droplet deformation.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Orientation in a fatty acid monolayer: Effect of flow type.\n \n \n \n \n\n\n \n Maruyama, T; Lauger, J; Fuller, G.; Frank, C.; and Robertson, C.\n\n\n \n\n\n\n Langmuir, 14(7): 1836–1845. 1998.\n \n\n\n\n
\n\n\n\n \n \n \"OrientationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{maruyama_orientation_1998,\n\tAbstract = {The two-dimensional fluid dynamics of the different phases of a fatty acid monolayer (docosanoic acid) were examined. Using Brewster angle microscopy, we studied the polydomain structure of two liquid condensed phases (the L2 and Ltextasciiacute2 phases) and the "solid", S, phase in situ during the application of extensional flow and simple shear flow. We found that only the L2 phase deformed nearly reversibly with a liquid-like response. Nonsymmetric domain deformations were, however, found for that phase at the lowest rate of strain studied with a four-roll mill. At higher strain rates, plots of the evolution of strain against time collapsed onto a single curve so that the rate of strain was independent of roller speed. Furthermore, a critical strain γc exists above which the rate of strain shows a stepwise increase despite the constant velocity of the rollers. The two other phases, the Ltextasciiacute2 and S phases, experienced flow-induced reorientation of the lattice onto which the molecules are arranged. The reorientation process was accompanied by the appearance of shear bands in the monolayers at textpm45textdegree to the extension axis of both types of flow. The shear bands observed in the Ltextasciiacute2 phase were modeled as a plastic flow accompanying the molecular tilt reorientation developed within elastic regions of the monolayer. This model describes the time evolution of the bandwidth quite well and provides strong evidence of the existence of an additional phase within the conventional Ltextasciiacute2 phase region. In simple shear flow, the velocity profile for the L2 phase across the gap of the shearing cell showed a nonlinear distribution of shear rates, which were highest at the center of the gap. Flow-induced breakup of domains was observed in the Ltextasciiacute2 phase subject to simple shear.},\n\tAuthor = {Maruyama, T and Lauger, J and Fuller, G.G. and Frank, C.W. and Robertson, C.R.},\n\tJournal = {Langmuir},\n\tNumber = {7},\n\tPages = {1836--1845},\n\tTitle = {Orientation in a fatty acid monolayer: {Effect} of flow type},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032024260&partnerID=40&md5=a5786ca825332f8a8fc36d72e5f109ca},\n\tVolume = {14},\n\tYear = {1998},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0032024260&partnerID=40&md5=a5786ca825332f8a8fc36d72e5f109ca}}\n\n
\n
\n\n\n
\n The two-dimensional fluid dynamics of the different phases of a fatty acid monolayer (docosanoic acid) were examined. Using Brewster angle microscopy, we studied the polydomain structure of two liquid condensed phases (the L2 and Ltextasciiacute2 phases) and the \"solid\", S, phase in situ during the application of extensional flow and simple shear flow. We found that only the L2 phase deformed nearly reversibly with a liquid-like response. Nonsymmetric domain deformations were, however, found for that phase at the lowest rate of strain studied with a four-roll mill. At higher strain rates, plots of the evolution of strain against time collapsed onto a single curve so that the rate of strain was independent of roller speed. Furthermore, a critical strain γc exists above which the rate of strain shows a stepwise increase despite the constant velocity of the rollers. The two other phases, the Ltextasciiacute2 and S phases, experienced flow-induced reorientation of the lattice onto which the molecules are arranged. The reorientation process was accompanied by the appearance of shear bands in the monolayers at textpm45textdegree to the extension axis of both types of flow. The shear bands observed in the Ltextasciiacute2 phase were modeled as a plastic flow accompanying the molecular tilt reorientation developed within elastic regions of the monolayer. This model describes the time evolution of the bandwidth quite well and provides strong evidence of the existence of an additional phase within the conventional Ltextasciiacute2 phase region. In simple shear flow, the velocity profile for the L2 phase across the gap of the shearing cell showed a nonlinear distribution of shear rates, which were highest at the center of the gap. Flow-induced breakup of domains was observed in the Ltextasciiacute2 phase subject to simple shear.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1997\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Optical rheometry of complex fluid interfaces.\n \n \n \n\n\n \n Fuller, G. G\n\n\n \n\n\n\n Technical Report 1997.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@techreport{fuller_optical_1997,\n\tAbstract = {Monolayers and other thin films can be theologically complex with highly nonlinear material properties, such as shear thinning surface shear viscosities, complex moduli with both real and imaginary components, and surface normal stress differences. These responses are related to couplings between the interfacial microstructure and flow that cause distortion and alignment of the molecular and mesoscale constituents comprising the film. To understand these phenomena, and to provide the necessary background for the development of microstructural constitutive models for interfacial rheology, optical methods can provide important measurements of in situ, flow-induced structural responses.},\n\tAuthor = {Fuller, Gerald G},\n\tPmid = {654977},\n\tTitle = {Optical rheometry of complex fluid interfaces},\n\tYear = {1997}}\n\n
\n
\n\n\n
\n Monolayers and other thin films can be theologically complex with highly nonlinear material properties, such as shear thinning surface shear viscosities, complex moduli with both real and imaginary components, and surface normal stress differences. These responses are related to couplings between the interfacial microstructure and flow that cause distortion and alignment of the molecular and mesoscale constituents comprising the film. To understand these phenomena, and to provide the necessary background for the development of microstructural constitutive models for interfacial rheology, optical methods can provide important measurements of in situ, flow-induced structural responses.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Branched viscoelastic surfactant solutions and their response to elongational flow.\n \n \n \n\n\n \n Fischer, P; Fuller, G.; and Lin, Z\n\n\n \n\n\n\n Rheologica Acta, 36(6): 632–638. 1997.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fischer_branched_1997,\n\tAuthor = {Fischer, P and Fuller, G.G. and Lin, Z},\n\tJournal = {Rheologica Acta},\n\tNumber = {6},\n\tPages = {632--638},\n\tTitle = {Branched viscoelastic surfactant solutions and their response to elongational flow},\n\tVolume = {36},\n\tYear = {1997}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Electric field studies of liquid crystal droplet suspensions.\n \n \n \n \n\n\n \n De Groot, E M; and Fuller, G.\n\n\n \n\n\n\n Liquid Crystals, 23(1): 113–126. 1997.\n \n\n\n\n
\n\n\n\n \n \n \"ElectricPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{de_groot_electric_1997,\n\tAbstract = {The responses of freely-suspended micron-sized liquid crystal droplets subjected to an alternating electric field are presented. By examining droplets of isotropic, nematic bipolar, and nematic radial configurations, we test the effect of anchoring on the droplet response. Specifically, using birefringence and scattering dichroism we measure the relaxation of electric field-induced orientation following a field pulse. Results indicate that bipolar and radial droplets in suspension orient in the field through very different mechanisms. Bipolar droplets are observed to rotate their defect axes in the field while radial droplets orient through a nematic distortion. By varying the field pulse, we observe that droplets also respond differently to the field depending on their relative sizes. In radial droplet suspensions we quantitatively measure time scales associated with the reorientation and restructuring of the defect region.},\n\tAuthor = {De Groot, E M and Fuller, G.G.},\n\tJournal = {Liquid Crystals},\n\tNumber = {1},\n\tPages = {113--126},\n\tTitle = {Electric field studies of liquid crystal droplet suspensions},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0007430592&partnerID=40&md5=abc550770fff2a49bcdffbda91b75327},\n\tVolume = {23},\n\tYear = {1997},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0007430592&partnerID=40&md5=abc550770fff2a49bcdffbda91b75327}}\n\n
\n
\n\n\n
\n The responses of freely-suspended micron-sized liquid crystal droplets subjected to an alternating electric field are presented. By examining droplets of isotropic, nematic bipolar, and nematic radial configurations, we test the effect of anchoring on the droplet response. Specifically, using birefringence and scattering dichroism we measure the relaxation of electric field-induced orientation following a field pulse. Results indicate that bipolar and radial droplets in suspension orient in the field through very different mechanisms. Bipolar droplets are observed to rotate their defect axes in the field while radial droplets orient through a nematic distortion. By varying the field pulse, we observe that droplets also respond differently to the field depending on their relative sizes. In radial droplet suspensions we quantitatively measure time scales associated with the reorientation and restructuring of the defect region.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Rheo-optical studies of shear-induced structures in semidilute polystyrene solutions.\n \n \n \n\n\n \n Kume, T; Hashimoto, T; Takahashi, T; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 30(23): 7232–7236. 1997.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kume_rheo-optical_1997,\n\tAuthor = {Kume, T and Hashimoto, T and Takahashi, T and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {23},\n\tPages = {7232--7236},\n\tTitle = {Rheo-optical studies of shear-induced structures in semidilute polystyrene solutions},\n\tVolume = {30},\n\tYear = {1997}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1996\n \n \n (12)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n In situ optical studies of flow-Induced orientation in a two-dimensional polymer solution.\n \n \n \n \n\n\n \n Friedenberg, M C; Fuller, G.; Frank, C.; and Robertson, C.\n\n\n \n\n\n\n Macromolecules, 29(2): 705–712. 1996.\n \n\n\n\n
\n\n\n\n \n \n \"InPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{friedenberg_situ_1996,\n\tAbstract = {The orientation dynamics of polymers in constrained geometries is considered through studies of monolayer films at the air-water interface. Here, in situ optical techniques are employed to probe flow orientation in monolayers of phthalocyaninatopolysiloxane dispersed in either docosanoic acid, 1,2-dimyristoyl-sn-glycero-3-phosphatidylcholine, or arachidyl alcohol. Compression of the polymer monolayer creates alignment perpendicular to the compression direction. A well-defined extensional flow is imposed in the monolayer to study the dynamics of flow-induced anisotropy. The orientation process obeys a strain-scaling law, indicating the absence of relaxation on the time scale of the flow.},\n\tAuthor = {Friedenberg, M C and Fuller, G.G. and Frank, C.W. and Robertson, C.R.},\n\tJournal = {Macromolecules},\n\tNumber = {2},\n\tPages = {705--712},\n\tTitle = {In situ optical studies of flow-{Induced} orientation in a two-dimensional polymer solution},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0006346662&partnerID=40&md5=5eff41ff41a79ba407d5a604f0b7814b},\n\tVolume = {29},\n\tYear = {1996},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0006346662&partnerID=40&md5=5eff41ff41a79ba407d5a604f0b7814b}}\n\n
\n
\n\n\n
\n The orientation dynamics of polymers in constrained geometries is considered through studies of monolayer films at the air-water interface. Here, in situ optical techniques are employed to probe flow orientation in monolayers of phthalocyaninatopolysiloxane dispersed in either docosanoic acid, 1,2-dimyristoyl-sn-glycero-3-phosphatidylcholine, or arachidyl alcohol. Compression of the polymer monolayer creates alignment perpendicular to the compression direction. A well-defined extensional flow is imposed in the monolayer to study the dynamics of flow-induced anisotropy. The orientation process obeys a strain-scaling law, indicating the absence of relaxation on the time scale of the flow.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Stress tensor measurement using birefringence in oblique transmission.\n \n \n \n \n\n\n \n Takahashi, T; and Fuller, G.\n\n\n \n\n\n\n Nihon Reoroji Gakkaishi, 24(3): 111–116. 1996.\n \n\n\n\n
\n\n\n\n \n \n \"StressPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{takahashi_stress_1996,\n\tAbstract = {A method for measuring the full sterss tensor of liquids obeying the stress-optical rule is presented. This technique is an extension of the procedure recently described by Burghardt and coworkers 4) wherein light is sent obliquely through a sample sheared between transparent plates. However, in the present development, the light is transmitted in the plane containing the velocity gradient and neutral directions, thereby reducing the necessary optical measurements by one. A polystyrene-tricresyl phosphate (TCP) solution is used as the test sample. The first and second normal stress differences in steady shear flow measured by this method show good agreement with the mechanical results measured by Magda et al. 1) using a modified cone and plate rheometer. The transient behavior of the first and second normal stress differences following the start-up of shear flow is also presented.},\n\tAuthor = {Takahashi, T and Fuller, G.G.},\n\tJournal = {Nihon Reoroji Gakkaishi},\n\tNumber = {3},\n\tPages = {111--116},\n\tTitle = {Stress tensor measurement using birefringence in oblique transmission},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0030469405&partnerID=40&md5=0e9673d347056b40a2822b2ee1448839},\n\tVolume = {24},\n\tYear = {1996},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0030469405&partnerID=40&md5=0e9673d347056b40a2822b2ee1448839}}\n\n
\n
\n\n\n
\n A method for measuring the full sterss tensor of liquids obeying the stress-optical rule is presented. This technique is an extension of the procedure recently described by Burghardt and coworkers 4) wherein light is sent obliquely through a sample sheared between transparent plates. However, in the present development, the light is transmitted in the plane containing the velocity gradient and neutral directions, thereby reducing the necessary optical measurements by one. A polystyrene-tricresyl phosphate (TCP) solution is used as the test sample. The first and second normal stress differences in steady shear flow measured by this method show good agreement with the mechanical results measured by Magda et al. 1) using a modified cone and plate rheometer. The transient behavior of the first and second normal stress differences following the start-up of shear flow is also presented.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Structure and rheology of wormlike micelles.\n \n \n \n \n\n\n \n Wheeler, E K; Izu, P; and Fuller, G.\n\n\n \n\n\n\n Rheologica Acta, 35(2): 139–149. 1996.\n \n\n\n\n
\n\n\n\n \n \n \"StructurePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wheeler_structure_1996,\n\tAbstract = {In a semi-dilute aqueous solution under certain conditions, surfactant molecules will self assemble to form wormlike micelles. The micelles are dynamic in structure since they can break and reform, providing an additional mode of relaxation. The viscoelastic properties of the wormlike micelles can be predicted using simple rheological models. For many surfactant solutions the mechanical data can be related to the optical data by the stress-optical rule. From the viscoelastic data it is possible to estimate the breaking time of the micelle. The techniques of birefringence and small angle light scattering are used to study the microstructure of a surfactant solution under simple shear and extensional flow. The sample under investigation is a solution of cetyltrimethylammonium bromide and sodium salicylate in water, with a salt to surfactant ratio of 7.7. Below a critical shear rate, the birefringence increases linearly with shear rate and the stress-optical rule is valid. The SALS patterns reveal distinctive butterfly patterns indicating that scattering is a result of concentration fluctuations that moderately couple to the flow. However, above a critical shear rate the birefringence plateaus and the stress-optical rule is no longer valid. SALS patterns show both a bright streak and a butterfly pattern. The bright streak is caused by elongated structures aligned in the direction of the flow. The oriented structures occur when the characteristic time of flow is faster than the breaking time of the micelles.},\n\tAuthor = {Wheeler, E K and Izu, P and Fuller, G.G.},\n\tJournal = {Rheologica Acta},\n\tNumber = {2},\n\tPages = {139--149},\n\tTitle = {Structure and rheology of wormlike micelles},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000970694&partnerID=40&md5=ea1be78e8465a07e656dd2e8845c3617},\n\tVolume = {35},\n\tYear = {1996},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000970694&partnerID=40&md5=ea1be78e8465a07e656dd2e8845c3617}}\n\n
\n
\n\n\n
\n In a semi-dilute aqueous solution under certain conditions, surfactant molecules will self assemble to form wormlike micelles. The micelles are dynamic in structure since they can break and reform, providing an additional mode of relaxation. The viscoelastic properties of the wormlike micelles can be predicted using simple rheological models. For many surfactant solutions the mechanical data can be related to the optical data by the stress-optical rule. From the viscoelastic data it is possible to estimate the breaking time of the micelle. The techniques of birefringence and small angle light scattering are used to study the microstructure of a surfactant solution under simple shear and extensional flow. The sample under investigation is a solution of cetyltrimethylammonium bromide and sodium salicylate in water, with a salt to surfactant ratio of 7.7. Below a critical shear rate, the birefringence increases linearly with shear rate and the stress-optical rule is valid. The SALS patterns reveal distinctive butterfly patterns indicating that scattering is a result of concentration fluctuations that moderately couple to the flow. However, above a critical shear rate the birefringence plateaus and the stress-optical rule is no longer valid. SALS patterns show both a bright streak and a butterfly pattern. The bright streak is caused by elongated structures aligned in the direction of the flow. The oriented structures occur when the characteristic time of flow is faster than the breaking time of the micelles.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Deformation and relaxation processes of mono- and bilayer domains of liquid crystalline langmuir films on water.\n \n \n \n \n\n\n \n Lauger, J; Robertson, C.; Frank, C.; and Fuller, G.\n\n\n \n\n\n\n Langmuir, 12(23): 5630–5635. 1996.\n \n\n\n\n
\n\n\n\n \n \n \"DeformationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lauger_deformation_1996,\n\tAbstract = {Flow deformation and relaxation of mono- and bilayer domains of a thermotropic liquid crystal at the air-water interface have been studied by means of simultaneous application of extensional flow fields and Brewster angle microscopy measurements. In an extentional flow, mono- and bilayer domains are found to be distorted into long, stringlike domains that are stable. This is due to the absence of Rayleigh instabilities in the two-dimensional case, which otherwise lead to the occurrence of a break-up mechanism in three dimensions. However, small fluctuations in the domain thickness lead to defects that can cause breakup of the string domains followed by a retraction of the two domain ends due to line tension effects. In relaxation experiments covering both the high deformation bola-shaped regime and the small-deformation regime of slightly deformed domains, the line tensions of mono- and bilayer domains were determined. Both deformation regimes yielded the same values for the line tension of the mono- and bilayer cases. However, the line tension of the bilayer was found to be about 10 times higher than that of the monolayer. It is believed that this large difference is a result of a strong attractive dipole-dipole interaction force in the interdigitated bilayer.},\n\tAuthor = {Lauger, J and Robertson, C.R. and Frank, C.W. and Fuller, G.G.},\n\tJournal = {Langmuir},\n\tNumber = {23},\n\tPages = {5630--5635},\n\tTitle = {Deformation and relaxation processes of mono- and bilayer domains of liquid crystalline langmuir films on water},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0040389452&partnerID=40&md5=1a9345a530ed1dee500cdfe3dc8565d2},\n\tVolume = {12},\n\tYear = {1996},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0040389452&partnerID=40&md5=1a9345a530ed1dee500cdfe3dc8565d2}}\n\n
\n
\n\n\n
\n Flow deformation and relaxation of mono- and bilayer domains of a thermotropic liquid crystal at the air-water interface have been studied by means of simultaneous application of extensional flow fields and Brewster angle microscopy measurements. In an extentional flow, mono- and bilayer domains are found to be distorted into long, stringlike domains that are stable. This is due to the absence of Rayleigh instabilities in the two-dimensional case, which otherwise lead to the occurrence of a break-up mechanism in three dimensions. However, small fluctuations in the domain thickness lead to defects that can cause breakup of the string domains followed by a retraction of the two domain ends due to line tension effects. In relaxation experiments covering both the high deformation bola-shaped regime and the small-deformation regime of slightly deformed domains, the line tensions of mono- and bilayer domains were determined. Both deformation regimes yielded the same values for the line tension of the mono- and bilayer cases. However, the line tension of the bilayer was found to be about 10 times higher than that of the monolayer. It is believed that this large difference is a result of a strong attractive dipole-dipole interaction force in the interdigitated bilayer.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Microstructural dynamics of a homopolymer melt investigated using two-dimensional Raman scattering.\n \n \n \n\n\n \n Huang, K; Archer, L A; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 29(3): 966–972. 1996.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{huang_microstructural_1996,\n\tAbstract = {The microstructural dynamics of polyisobutylene have been investigated using a new rheooptical technique, two-dimensional Raman scattering (2D Raman). 2D Raman employs intrinsic bond vibrations to selectively resolve the segmental dynamics within a material subjected to small-amplitude oscillatory deformation. In addition, interspecies interactions are probed through a cross-correlation of the orientation-induced response from distinct Raman peaks. It has been observed that the motions of the main chain and side groups in polyisobutylene are affected independently as the mechanical frequency is increased from the rubbery plateau into the glassy regime. In particular, at low frequencies (\\{\\vphantom{\\}}\\&},\n\tAuthor = {Huang, K and Archer, L A and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {3},\n\tPages = {966--972},\n\tTitle = {Microstructural dynamics of a homopolymer melt investigated using two-dimensional {Raman} scattering},\n\tVolume = {29},\n\tYear = {1996}}\n\n
\n
\n\n\n
\n The microstructural dynamics of polyisobutylene have been investigated using a new rheooptical technique, two-dimensional Raman scattering (2D Raman). 2D Raman employs intrinsic bond vibrations to selectively resolve the segmental dynamics within a material subjected to small-amplitude oscillatory deformation. In addition, interspecies interactions are probed through a cross-correlation of the orientation-induced response from distinct Raman peaks. It has been observed that the motions of the main chain and side groups in polyisobutylene are affected independently as the mechanical frequency is increased from the rubbery plateau into the glassy regime. In particular, at low frequencies (\\p̌hantom\\&\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n In-situ studies of flow-induced phenomena in Langmuir monolayers.\n \n \n \n \n\n\n \n Maruyama, T; Friedenberg, M; Fuller, G.; Frank, C.; Robertson, C.; Ferencz, A; and Wegner, G\n\n\n \n\n\n\n Thin Solid Films, 273(1-2): 76–83. 1996.\n \n\n\n\n
\n\n\n\n \n \n \"In-situPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{maruyama_-situ_1996,\n\tAbstract = {Flow-induced orientation of the hairy-rod polymer poly(phthalocyaninato siloxane) (PcPS) dissolved at 5 mol.\\% in docosanoic acid and residing in a monolayer at the air-water interface is reported. The flow was induced through two routes. In one, a four-roll mill was immersed through the monolayer to produce a well-defined extensional flow. In the second, a Langmuir-Blodgett (LB) deposition process was carried out simultaneously onto three parallel, glass plates, thereby generating a stagnation point, with extensional flow between each pair of slides. Linear dichroism was used to obtain in situ the measurements of orientation in the Langmuir films themselves, and on the glass substrates. It was found that the flows generated by the LB process induced substantial orientation within the Langmuir film, and this anisotropy was also evident in the LB films that were deposited. The velocity fields were measured using a particle tracking technique and compared against the predictions of a model assuming that the glass substrates act as line sinks of material during the LB process. It was found that the meniscus region distorts this velocity profile and leads to decreased alignment of the polymer chains.},\n\tAuthor = {Maruyama, T and Friedenberg, M and Fuller, G.G. and Frank, C.W. and Robertson, C.R. and Ferencz, A and Wegner, G},\n\tJournal = {Thin Solid Films},\n\tNumber = {1-2},\n\tPages = {76--83},\n\tTitle = {In-situ studies of flow-induced phenomena in {Langmuir} monolayers},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0030079950&partnerID=40&md5=4bbc4c4ec23224471278365ec87b0138},\n\tVolume = {273},\n\tYear = {1996},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0030079950&partnerID=40&md5=4bbc4c4ec23224471278365ec87b0138}}\n\n
\n
\n\n\n
\n Flow-induced orientation of the hairy-rod polymer poly(phthalocyaninato siloxane) (PcPS) dissolved at 5 mol.% in docosanoic acid and residing in a monolayer at the air-water interface is reported. The flow was induced through two routes. In one, a four-roll mill was immersed through the monolayer to produce a well-defined extensional flow. In the second, a Langmuir-Blodgett (LB) deposition process was carried out simultaneously onto three parallel, glass plates, thereby generating a stagnation point, with extensional flow between each pair of slides. Linear dichroism was used to obtain in situ the measurements of orientation in the Langmuir films themselves, and on the glass substrates. It was found that the flows generated by the LB process induced substantial orientation within the Langmuir film, and this anisotropy was also evident in the LB films that were deposited. The velocity fields were measured using a particle tracking technique and compared against the predictions of a model assuming that the glass substrates act as line sinks of material during the LB process. It was found that the meniscus region distorts this velocity profile and leads to decreased alignment of the polymer chains.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Direct visualization of flow-induced anisotropy in a fatty acid monolayer.\n \n \n \n \n\n\n \n Friedenberg, M C; Fuller, G.; Frank, C.; and Robertson, C.\n\n\n \n\n\n\n Langmuir, 12(6): 1594–1599. 1996.\n \n\n\n\n
\n\n\n\n \n \n \"DirectPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{friedenberg_direct_1996,\n\tAbstract = {Brewster angle microscopy is used to directly visualize the influence of an applied extensional flow on the domain structure and molecular orientation of a docosanoic acid monolayer at the air-water interface. At a surface pressure of 12 mN/m and a subphase temperature of 15 textdegreetextdegreeC (L2 phase), extensional flow causes domain elongation parallel to the extension axis. A frequency domain analysis of the Brewster angle images indicates that the domains undergo an affine deformation in response to flow. AT 20 mN/m (L2textasciiacute phase), the flow modifies not only the domain structure of the monolayer but also the azimuthal orientation of the fatty acid molecules. This flow-alignment process is strain-rate dependent. Thus, flow can couple to the monolayer order over a variety of length scales.},\n\tAuthor = {Friedenberg, M C and Fuller, G.G. and Frank, C.W. and Robertson, C.R.},\n\tJournal = {Langmuir},\n\tNumber = {6},\n\tPages = {1594--1599},\n\tTitle = {Direct visualization of flow-induced anisotropy in a fatty acid monolayer},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000489686&partnerID=40&md5=c3a92f2494e774aaa68ddf74d9cd4611},\n\tVolume = {12},\n\tYear = {1996},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000489686&partnerID=40&md5=c3a92f2494e774aaa68ddf74d9cd4611}}\n\n
\n
\n\n\n
\n Brewster angle microscopy is used to directly visualize the influence of an applied extensional flow on the domain structure and molecular orientation of a docosanoic acid monolayer at the air-water interface. At a surface pressure of 12 mN/m and a subphase temperature of 15 textdegreetextdegreeC (L2 phase), extensional flow causes domain elongation parallel to the extension axis. A frequency domain analysis of the Brewster angle images indicates that the domains undergo an affine deformation in response to flow. AT 20 mN/m (L2textasciiacute phase), the flow modifies not only the domain structure of the monolayer but also the azimuthal orientation of the fatty acid molecules. This flow-alignment process is strain-rate dependent. Thus, flow can couple to the monolayer order over a variety of length scales.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Polarization-modulated Raman scattering measurements of nematic liquid crystal orientation.\n \n \n \n \n\n\n \n Huang, K; Archer, L A; and Fuller, G.\n\n\n \n\n\n\n Rev. Sci. Instrum., 67(11): 3924–3930. 1996.\n \n\n\n\n
\n\n\n\n \n \n \"Polarization-modulatedPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{huang_polarization-modulated_1996,\n\tAbstract = {This article presents a new, automated Raman scattering experiment to measure the orientation in nematic liquid crystals. The automated experiment utilizes phase modulation to alter the polarization state of light unlike conventional, polarized Raman experiments which require the manual manipulation of polarizers and multiple measurements. Polarization modulation imparts the ability to quantify the second and fourth moments of the orientation distribution function with a single measurement. The feasibility of the new experiment is demonstrated through orientation measurements of a well characterized nematic liquid crystal, 4-pentyl-(4-cyanophenyl)-cyclohexane. Additionally, since the results of the polarization-modulated technique are in excellent agreement with conventional Raman measurements, this study supports the validity of the theory describing Raman scattering from birefringent materials. textcopyright 1996 American Institute of Physics.},\n\tAuthor = {Huang, K and Archer, L A and Fuller, G.G.},\n\tJournal = {Rev. Sci. Instrum.},\n\tNumber = {11},\n\tPages = {3924--3930},\n\tTitle = {Polarization-modulated {Raman} scattering measurements of nematic liquid crystal orientation},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0030289397&partnerID=40&md5=08fe3bf465d010aae35da78e3a69da6c},\n\tVolume = {67},\n\tYear = {1996},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0030289397&partnerID=40&md5=08fe3bf465d010aae35da78e3a69da6c}}\n\n
\n
\n\n\n
\n This article presents a new, automated Raman scattering experiment to measure the orientation in nematic liquid crystals. The automated experiment utilizes phase modulation to alter the polarization state of light unlike conventional, polarized Raman experiments which require the manual manipulation of polarizers and multiple measurements. Polarization modulation imparts the ability to quantify the second and fourth moments of the orientation distribution function with a single measurement. The feasibility of the new experiment is demonstrated through orientation measurements of a well characterized nematic liquid crystal, 4-pentyl-(4-cyanophenyl)-cyclohexane. Additionally, since the results of the polarization-modulated technique are in excellent agreement with conventional Raman measurements, this study supports the validity of the theory describing Raman scattering from birefringent materials. textcopyright 1996 American Institute of Physics.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Dynamic response of a near-critical polymer blend solution under oscillatory shear flow.\n \n \n \n \n\n\n \n Lai, J; and Fuller, G.\n\n\n \n\n\n\n Journal of Rheology, 40(1): 153–166. 1996.\n \n\n\n\n
\n\n\n\n \n \n \"DynamicPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lai_dynamic_1996,\n\tAbstract = {External flow is known to induce anisotropic growth of concentration fluctuations in polymer solutions close to the phase boundary. In this article, scattering dichroism measurements are used to investigate the structural dynamics of flow-enhanced concentration fluctuations in a near-critical, semidilute polymer blend solution subjected to small-amplitude oscillatory shear flow. Our measured frequency response reveals that the behavior of the shear-induced fluctuations is governed by the relaxation time of the concentration fluctuations. Predictions from a hydrodynamic model, in good agreement with the experimental results, also indicate that the dynamics are dominated by a single relaxation time. In addition, relaxation times of the concentration fluctuations at various temperatures are calculated from dichroism data and are found to agree with those previously obtained from small-angle light scattering experiments. textcopyright 1996 Society of Rheology.},\n\tAuthor = {Lai, J and Fuller, G.G.},\n\tJournal = {Journal of Rheology},\n\tNumber = {1},\n\tPages = {153--166},\n\tTitle = {Dynamic response of a near-critical polymer blend solution under oscillatory shear flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0029733467&partnerID=40&md5=5e76c1ce02d8f4db27175b48fdfa25a3},\n\tVolume = {40},\n\tYear = {1996},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0029733467&partnerID=40&md5=5e76c1ce02d8f4db27175b48fdfa25a3}}\n\n
\n
\n\n\n
\n External flow is known to induce anisotropic growth of concentration fluctuations in polymer solutions close to the phase boundary. In this article, scattering dichroism measurements are used to investigate the structural dynamics of flow-enhanced concentration fluctuations in a near-critical, semidilute polymer blend solution subjected to small-amplitude oscillatory shear flow. Our measured frequency response reveals that the behavior of the shear-induced fluctuations is governed by the relaxation time of the concentration fluctuations. Predictions from a hydrodynamic model, in good agreement with the experimental results, also indicate that the dynamics are dominated by a single relaxation time. In addition, relaxation times of the concentration fluctuations at various temperatures are calculated from dichroism data and are found to agree with those previously obtained from small-angle light scattering experiments. textcopyright 1996 Society of Rheology.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Flow-induced molecular orientation of a Langmuir film.\n \n \n \n\n\n \n Maruyama, T; Fuller, G; Frank, C; and Robertson, C\n\n\n \n\n\n\n Science, 274(Washington, DC, United States): 233–235. 1996.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{maruyama_flow-induced_1996,\n\tAuthor = {Maruyama, T and Fuller, G and Frank, C and Robertson, C},\n\tJournal = {Science},\n\tNumber = {Washington, DC, United States},\n\tPages = {233--235},\n\tTitle = {Flow-induced molecular orientation of a {Langmuir} film},\n\tVolume = {274},\n\tYear = {1996}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Perspective: Comments on ``The dynamic birefringence of high polymers,'' by Richard S. Stein, Shigeharu Onogi, and Daniel A. Keedy,J. polym. sci., 57, 801 (1962).\n \n \n \n \n\n\n \n Fuller, G. G\n\n\n \n\n\n\n Journal of Polymer Science, Part B: Polymer Physics, 34(9): 1505–1506. July 1996.\n \n\n\n\n
\n\n\n\n \n \n \"Perspective:Paper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_perspective:_1996,\n\tAuthor = {Fuller, Gerald G},\n\tDoi = {10.1002/polb.1996.917},\n\tJournal = {Journal of Polymer Science, Part B: Polymer Physics},\n\tMonth = jul,\n\tNumber = {9},\n\tPages = {1505--1506},\n\tTitle = {Perspective: {Comments} on ``{The} dynamic birefringence of high polymers,'' by {Richard} {S}. {Stein}, {Shigeharu} {Onogi}, and {Daniel} {A}. {Keedy},{J}. polym. sci., 57, 801 (1962)},\n\tUrl = {http://doi.wiley.com/10.1002/polb.1996.917},\n\tVolume = {34},\n\tYear = {1996},\n\tBdsk-Url-1 = {http://doi.wiley.com/10.1002/polb.1996.917},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1002/polb.1996.917}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Extensional flow of a two-dimensional polymer liquid crystal.\n \n \n \n \n\n\n \n Maffettone, P. L.; Grosso, M.; Friedenberg, M. C; and Fuller, G. G\n\n\n \n\n\n\n Macromolecules, 29(Washington, DC, United States): 8473–8478. 1996.\n \n\n\n\n
\n\n\n\n \n \n \"ExtensionalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{maffettone_extensional_1996,\n\tAbstract = {The transient behavior of a polymeric nematic liquid crystal interface in extensional flows is studied both experimentally and theoretically. Monolayers of the hairy rod polymer phthatocyan-inatopolysiloxane subjected to two-dimensional, transient, extensional flows are modeled with the two-dimensional analogue of the rigid rod model. The scalar order parameter and the director orientation are compared with the experimental observables. Two parameters appear in the model: an average rotational diffusivity and the intensity of the nematic field. The average rotational diffusivity is determined by fitting relaxation experiments. The intensity of the nematic field, which is modeled with the Onsager potential, is determined by starting from the molecular parameters. A good quantitative agreement is obtained between experiments and theoretical predictions.},\n\tAuthor = {Maffettone, Pier Luca and Grosso, Massimiliano and Friedenberg, Matthew C and Fuller, Gerald G},\n\tJournal = {Macromolecules},\n\tNumber = {Washington, DC, United States},\n\tPages = {8473--8478},\n\tTitle = {Extensional flow of a two-dimensional polymer liquid crystal},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0030415401&partnerID=40&md5=707e6fe0514dab4cf46f439c73fb89cb},\n\tVolume = {29},\n\tYear = {1996},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0030415401&partnerID=40&md5=707e6fe0514dab4cf46f439c73fb89cb}}\n\n
\n
\n\n\n
\n The transient behavior of a polymeric nematic liquid crystal interface in extensional flows is studied both experimentally and theoretically. Monolayers of the hairy rod polymer phthatocyan-inatopolysiloxane subjected to two-dimensional, transient, extensional flows are modeled with the two-dimensional analogue of the rigid rod model. The scalar order parameter and the director orientation are compared with the experimental observables. Two parameters appear in the model: an average rotational diffusivity and the intensity of the nematic field. The average rotational diffusivity is determined by fitting relaxation experiments. The intensity of the nematic field, which is modeled with the Onsager potential, is determined by starting from the molecular parameters. A good quantitative agreement is obtained between experiments and theoretical predictions.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1995\n \n \n (5)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Rheo-optical characterization (flow-birefringence and flow-dichroism) of the Tobacco Mosaic Virus.\n \n \n \n \n\n\n \n Reinhardt, U. T; de Groot, E. L M.; Fuller, G. G; and Kulicke, W.\n\n\n \n\n\n\n Macromolecular Chemistry and Physics, 196(1): 63–74. January 1995.\n \n\n\n\n
\n\n\n\n \n \n \"Rheo-opticalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{reinhardt_rheo-optical_1995,\n\tAuthor = {Reinhardt, Ulf T and de Groot, Eleanor L Meyer and Fuller, Gerald G and Kulicke, Werner-Michael},\n\tDoi = {10.1002/macp.1995.021960103},\n\tJournal = {Macromolecular Chemistry and Physics},\n\tMonth = jan,\n\tNumber = {1},\n\tPages = {63--74},\n\tTitle = {Rheo-optical characterization (flow-birefringence and flow-dichroism) of the {Tobacco} {Mosaic} {Virus}},\n\tUrl = {http://doi.wiley.com/10.1002/macp.1995.021960103},\n\tVolume = {196},\n\tYear = {1995},\n\tBdsk-Url-1 = {http://doi.wiley.com/10.1002/macp.1995.021960103},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1002/macp.1995.021960103}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Optical rheometry of multicomponent polymer liquids.\n \n \n \n \n\n\n \n Fuller, G. G\n\n\n \n\n\n\n Macromolecular Symposia, 98(1): 997–1003. July 1995.\n \n\n\n\n
\n\n\n\n \n \n \"OpticalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_optical_1995,\n\tAuthor = {Fuller, Gerald G},\n\tDoi = {10.1002/masy.19950980190},\n\tJournal = {Macromolecular Symposia},\n\tMonth = jul,\n\tNumber = {1},\n\tPages = {997--1003},\n\tTitle = {Optical rheometry of multicomponent polymer liquids},\n\tUrl = {http://doi.wiley.com/10.1002/masy.19950980190},\n\tVolume = {98},\n\tYear = {1995},\n\tBdsk-Url-1 = {http://doi.wiley.com/10.1002/masy.19950980190},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1002/masy.19950980190}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Enhancement of orientational coupling in bidisperse polybutadiene melts through the implementation of directed interactions.\n \n \n \n \n\n\n \n Seidel, U; Stadler, R; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 28(10): 3739–3740. 1995.\n \n\n\n\n
\n\n\n\n \n \n \"EnhancementPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{seidel_enhancement_1995,\n\tAuthor = {Seidel, U and Stadler, R and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {10},\n\tPages = {3739--3740},\n\tTitle = {Enhancement of orientational coupling in bidisperse polybutadiene melts through the implementation of directed interactions},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000992357&partnerID=40&md5=bf119920b502b01df7a1f1d6be818805},\n\tVolume = {28},\n\tYear = {1995},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000992357&partnerID=40&md5=bf119920b502b01df7a1f1d6be818805}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The stress jump of a semirigid macromolecule after shear: Comparison of the elastic stress to the birefringence.\n \n \n \n \n\n\n \n Smyth, S. F\n\n\n \n\n\n\n Journal of Rheology, 39(4): 659. July 1995.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{smyth_stress_1995,\n\tAuthor = {Smyth, Stuart F},\n\tDoi = {10.1122/1.550649},\n\tJournal = {Journal of Rheology},\n\tMonth = jul,\n\tNumber = {4},\n\tPages = {659},\n\tTitle = {The stress jump of a semirigid macromolecule after shear: {Comparison} of the elastic stress to the birefringence},\n\tUrl = {http://adsabs.harvard.edu/abs/1995JRheo..39..659S},\n\tVolume = {39},\n\tYear = {1995},\n\tBdsk-Url-1 = {http://adsabs.harvard.edu/abs/1995JRheo..39..659S},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1122/1.550649}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n A rheo-optical study of near-critical polymer solutions under oscillatory shear flow.\n \n \n \n \n\n\n \n Lai, J.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Rheology, 39(5): 893. September 1995.\n \n\n\n\n
\n\n\n\n \n \n \"APaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lai_rheo-optical_1995,\n\tAuthor = {Lai, Janet and Fuller, Gerald G},\n\tDoi = {10.1122/1.550623},\n\tJournal = {Journal of Rheology},\n\tMonth = sep,\n\tNumber = {5},\n\tPages = {893},\n\tTitle = {A rheo-optical study of near-critical polymer solutions under oscillatory shear flow},\n\tUrl = {http://adsabs.harvard.edu/abs/1995JRheo..39..893L},\n\tVolume = {39},\n\tYear = {1995},\n\tBdsk-Url-1 = {http://adsabs.harvard.edu/abs/1995JRheo..39..893L},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1122/1.550623}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1994\n \n \n (12)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Component dynamics in miscible blends of 1,4-polyisoprene and 1,2-polybutadiene.\n \n \n \n \n\n\n \n Zawada, J A; Fuller, G.; Colby, R H; Fetters, L J; and Roovers, J\n\n\n \n\n\n\n Macromolecules, 27(23): 6861–6870. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"ComponentPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{zawada_component_1994,\n\tAbstract = {The effects of blending on the rheology of the individual components in highly entangled miscible blends of 1,4-polyisoprene (PI) and 1,2-polybutadiene (1,2-PB) are investigated. Blend component contributions to the dynamic modulus, G*(\\${\\o}mega\\$), are recovered over the full composition range by complementing dynamic mechanical rheometry with infrared polarimetry. Distinct relaxations for each component are observed. Analysis of the modulus amplitudes of the component G*(\\${\\o}mega\\$) contributions reveals that the behavior of the slower relaxing component, 1,2-PB, compares well with constraint release scaling predictions. Each component is found to have a different average number of skeletal bonds between entanglements, in agreement with the predictions of an existing entanglement theory. The components appear to adopt a single reptation tube diameter and a mutual modulus shift factor, bT(T), in each blend. Analysis of the frequency dependence of the component G*(\\${\\o}mega\\$) contributions indicates that each component's relaxation is governed by a distinct apparent glass transition temperature, Tg, and that at a constant T - Tg blend composition only mildly influences the component relaxation times. A mild increase in the 1,2-PB relaxation time in the blend suggests a possible increase in the 1,2-PB friction factor in the presence of PI. textcopyright 1994 American Chemical Society.},\n\tAuthor = {Zawada, J A and Fuller, G.G. and Colby, R H and Fetters, L J and Roovers, J},\n\tJournal = {Macromolecules},\n\tNumber = {23},\n\tPages = {6861--6870},\n\tTitle = {Component dynamics in miscible blends of 1,4-polyisoprene and 1,2-polybutadiene},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028767787&partnerID=40&md5=c55850406267f3dbfbc46a29bffb55dc},\n\tVolume = {27},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028767787&partnerID=40&md5=c55850406267f3dbfbc46a29bffb55dc}}\n\n
\n
\n\n\n
\n The effects of blending on the rheology of the individual components in highly entangled miscible blends of 1,4-polyisoprene (PI) and 1,2-polybutadiene (1,2-PB) are investigated. Blend component contributions to the dynamic modulus, G*(${ø}mega$), are recovered over the full composition range by complementing dynamic mechanical rheometry with infrared polarimetry. Distinct relaxations for each component are observed. Analysis of the modulus amplitudes of the component G*(${ø}mega$) contributions reveals that the behavior of the slower relaxing component, 1,2-PB, compares well with constraint release scaling predictions. Each component is found to have a different average number of skeletal bonds between entanglements, in agreement with the predictions of an existing entanglement theory. The components appear to adopt a single reptation tube diameter and a mutual modulus shift factor, bT(T), in each blend. Analysis of the frequency dependence of the component G*(${ø}mega$) contributions indicates that each component's relaxation is governed by a distinct apparent glass transition temperature, Tg, and that at a constant T - Tg blend composition only mildly influences the component relaxation times. A mild increase in the 1,2-PB relaxation time in the blend suggests a possible increase in the 1,2-PB friction factor in the presence of PI. textcopyright 1994 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Formation of bilayer disks and two-dimensional foams on a collapsing/expanding liquid-crystal monolayer.\n \n \n \n \n\n\n \n Friedenberg, M C; Fuller, G.; Frank, C.; and Robertson, C.\n\n\n \n\n\n\n Langmuir, 10(4): 1251–1256. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"FormationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{friedenberg_formation_1994,\n\tAbstract = {We have visualized the collapse of a monolayer film of the thermotropic liquid-crystal 4textasciiacute-n-octyl-4-cyanobiphenyl (8CB) at the air-water interface using Brewster angle microscopy. First, the film collapses into small circular domains that represent fluid multilayer structures under the influence of a line tension. These multilayer disks grow in both area and number with further compression. Coalescence of domains occurs at high area fraction and is coupled with flow of the underlying monolayer. At still higher compression, additional coalescence leads to a homogeneous multilayer film. When the multilayer is allowed to expand, circular holes form. As these holes grow on further expansion, a two-dimensional foam is formed, stabilized by thin lamellae of the collapsed 8CB phase. The foam eventually breaks apart as the film expands back into the monolayer regime. Our results confirm the proposal by Xue et al. [Phys. Rev. Lett. 1992, 69, 474] that this collapse is a transition between the monolayer and a uniform trilayer composed of an interdigitated bilayer on top of a monolayer. textcopyright 1994 American Chemical Society.},\n\tAuthor = {Friedenberg, M C and Fuller, G.G. and Frank, C.W. and Robertson, C.R.},\n\tJournal = {Langmuir},\n\tNumber = {4},\n\tPages = {1251--1256},\n\tTitle = {Formation of bilayer disks and two-dimensional foams on a collapsing/expanding liquid-crystal monolayer},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028422485&partnerID=40&md5=276e7b4ff51931cbccb1fa0aa79a7198},\n\tVolume = {10},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028422485&partnerID=40&md5=276e7b4ff51931cbccb1fa0aa79a7198}}\n\n
\n
\n\n\n
\n We have visualized the collapse of a monolayer film of the thermotropic liquid-crystal 4textasciiacute-n-octyl-4-cyanobiphenyl (8CB) at the air-water interface using Brewster angle microscopy. First, the film collapses into small circular domains that represent fluid multilayer structures under the influence of a line tension. These multilayer disks grow in both area and number with further compression. Coalescence of domains occurs at high area fraction and is coupled with flow of the underlying monolayer. At still higher compression, additional coalescence leads to a homogeneous multilayer film. When the multilayer is allowed to expand, circular holes form. As these holes grow on further expansion, a two-dimensional foam is formed, stabilized by thin lamellae of the collapsed 8CB phase. The foam eventually breaks apart as the film expands back into the monolayer regime. Our results confirm the proposal by Xue et al. [Phys. Rev. Lett. 1992, 69, 474] that this collapse is a transition between the monolayer and a uniform trilayer composed of an interdigitated bilayer on top of a monolayer. textcopyright 1994 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Pattern and segment relaxation in a block copolymer melt following step shear flow.\n \n \n \n \n\n\n \n Archer, L A; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 27(24): 7152–7156. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"PatternPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{archer_pattern_1994,\n\tAbstract = {The relaxation dynamics of a homogeneous poly(styrene)-poly(butadiene) star diblock copolymer melt following step shear flow is studied as its temperature is lowered through the microphase separation temperature (MST). Transient Raman scattering, birefringence, and mechanical rheometry measurements are used to investigate segment and composition pattern relaxation dynamics. The birefringence was determined to be primarily form birefringence, and its relaxation rate was found to decrease dramatically in the vicinity of the MST. At temperatures close to the MST, the shear stress, first normal stress difference, and segment orientation anisotropy (from Raman scattering measurements) relaxation rates also decreased. Slowing down of the first normal stress difference and segment orientation anisotropy relaxations was found to be less dramatic than that of the birefringence relaxation but more pronounced than that of the shear stress relaxation. A decrease in the relaxation rate of segment orientation anisotropy close to the MST is proposed to result from the coupling of segmental and composition pattern relaxations due to tethering of chain segments to the domain interface. textcopyright 1994 American Chemical Society.},\n\tAuthor = {Archer, L A and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {24},\n\tPages = {7152--7156},\n\tTitle = {Pattern and segment relaxation in a block copolymer melt following step shear flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028547968&partnerID=40&md5=3def1188f25687b249c8d4a22d756028},\n\tVolume = {27},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028547968&partnerID=40&md5=3def1188f25687b249c8d4a22d756028}}\n\n
\n
\n\n\n
\n The relaxation dynamics of a homogeneous poly(styrene)-poly(butadiene) star diblock copolymer melt following step shear flow is studied as its temperature is lowered through the microphase separation temperature (MST). Transient Raman scattering, birefringence, and mechanical rheometry measurements are used to investigate segment and composition pattern relaxation dynamics. The birefringence was determined to be primarily form birefringence, and its relaxation rate was found to decrease dramatically in the vicinity of the MST. At temperatures close to the MST, the shear stress, first normal stress difference, and segment orientation anisotropy (from Raman scattering measurements) relaxation rates also decreased. Slowing down of the first normal stress difference and segment orientation anisotropy relaxations was found to be less dramatic than that of the birefringence relaxation but more pronounced than that of the shear stress relaxation. A decrease in the relaxation rate of segment orientation anisotropy close to the MST is proposed to result from the coupling of segmental and composition pattern relaxations due to tethering of chain segments to the domain interface. textcopyright 1994 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Measuring component contributions to the dynamic modulus in miscible polymer blends.\n \n \n \n \n\n\n \n Zawada, J A; Fuller, G.; Colby, R H; Fetters, L J; and Roovers, J\n\n\n \n\n\n\n Macromolecules, 27(23): 6851–6860. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"MeasuringPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{zawada_measuring_1994,\n\tAbstract = {A novel approach to investigating the rheology of miscible polymer blends is described based on complementary infrared polarimetry and mechanical rheometry measurements. The method provides a means by which the blend dynamic modulus G*(\\${\\o}mega\\$) may be separated into the individiual contributions due to each of the blend components. This separation is made possible by converting the observables of dynamic infrared 1,3-dichroism and birefringence experiments to those of dynamic shear stress experiments through the use of constitutive relations and the application of the stress-optic rule. The approach is extended to cover blend systems which exhibit orientational coupling, an effect which influences the optical anisotropies but does not contribute to the state of stress. Validity of the analysis technique is demonstrated by studying highly entangled miscible blends of 1,4-polyisoprene, PI (Mw = 75 000), and 1,2-polybutadiene, 1,2-PB (Mw = 204 000), containing a fraction of deuterium-labeled PI chains (Mw = 90 000). Orientational coupling in this blend is observed and is found to be characterized by a coupling coefficient ∈ = 0.35 textpm 0.05. textcopyright 1994 American Chemical Society.},\n\tAuthor = {Zawada, J A and Fuller, G.G. and Colby, R H and Fetters, L J and Roovers, J},\n\tJournal = {Macromolecules},\n\tNumber = {23},\n\tPages = {6851--6860},\n\tTitle = {Measuring component contributions to the dynamic modulus in miscible polymer blends},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028767847&partnerID=40&md5=d693a6a88e089805b7b2d07de9cba780},\n\tVolume = {27},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028767847&partnerID=40&md5=d693a6a88e089805b7b2d07de9cba780}}\n\n
\n
\n\n\n
\n A novel approach to investigating the rheology of miscible polymer blends is described based on complementary infrared polarimetry and mechanical rheometry measurements. The method provides a means by which the blend dynamic modulus G*(${ø}mega$) may be separated into the individiual contributions due to each of the blend components. This separation is made possible by converting the observables of dynamic infrared 1,3-dichroism and birefringence experiments to those of dynamic shear stress experiments through the use of constitutive relations and the application of the stress-optic rule. The approach is extended to cover blend systems which exhibit orientational coupling, an effect which influences the optical anisotropies but does not contribute to the state of stress. Validity of the analysis technique is demonstrated by studying highly entangled miscible blends of 1,4-polyisoprene, PI (Mw = 75 000), and 1,2-polybutadiene, 1,2-PB (Mw = 204 000), containing a fraction of deuterium-labeled PI chains (Mw = 90 000). Orientational coupling in this blend is observed and is found to be characterized by a coupling coefficient ∈ = 0.35 textpm 0.05. textcopyright 1994 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Segment orientation in a quiescent block copolymer melt studied by Raman scattering.\n \n \n \n \n\n\n \n Archer, L A; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 27(15): 4359–4363. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"SegmentPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{archer_segment_1994,\n\tAbstract = {Segment orientation of a poly(styrene)-poly(butadiene) star diblock copolymer melt is investigated using polarization-modulated laser Raman scattering (PMLRS). The PMLRS technique is based on Raman spectroscopy and allows for the simultaneous measurement of statistical segment orientation and birefringence. It is found that at temperatures below the glass transition temperature of poly(styrene), segments in both poly(styrene) and poly(butadiene) blocks preferentially align normal to the interface separating microdomains. However, at temperatures greater than the glass transition temperature of poly(styrene), segment orientation disappears. At this temperature, a dramatic decrease in birefringence is also observed. The implications of these findings on the existence of copolymer chain extension in the so-called strong segregation limit are discussed. textcopyright 1994 American Chemical Society.},\n\tAuthor = {Archer, L A and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {15},\n\tPages = {4359--4363},\n\tTitle = {Segment orientation in a quiescent block copolymer melt studied by {Raman} scattering},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028462834&partnerID=40&md5=96195d9dfa305c24f02a6b6cdbaca520},\n\tVolume = {27},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028462834&partnerID=40&md5=96195d9dfa305c24f02a6b6cdbaca520}}\n\n
\n
\n\n\n
\n Segment orientation of a poly(styrene)-poly(butadiene) star diblock copolymer melt is investigated using polarization-modulated laser Raman scattering (PMLRS). The PMLRS technique is based on Raman spectroscopy and allows for the simultaneous measurement of statistical segment orientation and birefringence. It is found that at temperatures below the glass transition temperature of poly(styrene), segments in both poly(styrene) and poly(butadiene) blocks preferentially align normal to the interface separating microdomains. However, at temperatures greater than the glass transition temperature of poly(styrene), segment orientation disappears. At this temperature, a dramatic decrease in birefringence is also observed. The implications of these findings on the existence of copolymer chain extension in the so-called strong segregation limit are discussed. textcopyright 1994 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Investigating miscible polymer blend dynamics with optical and mechanical rheometry.\n \n \n \n \n\n\n \n Fuller, G.; Zawada, J A; and Colby, R H\n\n\n \n\n\n\n Journal of Non-Crystalline Solids, 172-174(PART 2): 668–673. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"InvestigatingPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_investigating_1994,\n\tAbstract = {A method to identify component contributions to the rheology of miscible polymer blends is proposed based on complementary infrared polarimetry and mechanical rheometry. With such measurements the blend dynamic modulus, G*(\\${\\o}mega\\$), is shown to be separable into the individual contributions experienced by each of the blend components. The method relies on converting the observables of dynamic infrared 1,3 dichroism and birefringence experiments to those of dynamic shear stress experiments through the use of constitutive relations and the application of the stress-optic rule. Measures are introduced to account for local orientational correlations since such coupling effects are known to influence optical anisotropies but not the state of stress. textcopyright 1994.},\n\tAuthor = {Fuller, G.G. and Zawada, J A and Colby, R H},\n\tJournal = {Journal of Non-Crystalline Solids},\n\tNumber = {PART 2},\n\tPages = {668--673},\n\tTitle = {Investigating miscible polymer blend dynamics with optical and mechanical rheometry},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028495528&partnerID=40&md5=de555fc6a40ad058c3010f70c5fad059},\n\tVolume = {172-174},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028495528&partnerID=40&md5=de555fc6a40ad058c3010f70c5fad059}}\n\n
\n
\n\n\n
\n A method to identify component contributions to the rheology of miscible polymer blends is proposed based on complementary infrared polarimetry and mechanical rheometry. With such measurements the blend dynamic modulus, G*(${ø}mega$), is shown to be separable into the individual contributions experienced by each of the blend components. The method relies on converting the observables of dynamic infrared 1,3 dichroism and birefringence experiments to those of dynamic shear stress experiments through the use of constitutive relations and the application of the stress-optic rule. Measures are introduced to account for local orientational correlations since such coupling effects are known to influence optical anisotropies but not the state of stress. textcopyright 1994.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Optical and mechanical properties of a star diblock copolymer melt in oscillatory shear flow.\n \n \n \n \n\n\n \n Archer, L A; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 27(17): 4804–4809. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"OpticalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{archer_optical_1994,\n\tAbstract = {Mechanical rheometry and birefringence measurements are used to study the low-frequency dynamics of a homogeneous, star diblock copolymer melt as it is cooled through its microphase separation temperature (MST). Anomalous low-frequency behavior in both mechanical and optical properties are observed at temperatures close to the MST. We conclude that a general condition must be satisfied before unusual low-frequency behavior is observed, namely that the time scale of flow must be less than the relaxation time of structural features perturbed by flow. Furthermore, because of the similarities between the low-frequency dynamics of the present star diblock copolymer and those of linear diblock and triblock copolymers, we also conclude that block copolymer architecture does not play a critical role in determining the low-frequency behavior of these, materials. textcopyright 1994 American Chemical Society.},\n\tAuthor = {Archer, L A and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {17},\n\tPages = {4804--4809},\n\tTitle = {Optical and mechanical properties of a star diblock copolymer melt in oscillatory shear flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028484662&partnerID=40&md5=aac242e2077372ef8ca033ae1799161a},\n\tVolume = {27},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028484662&partnerID=40&md5=aac242e2077372ef8ca033ae1799161a}}\n\n
\n
\n\n\n
\n Mechanical rheometry and birefringence measurements are used to study the low-frequency dynamics of a homogeneous, star diblock copolymer melt as it is cooled through its microphase separation temperature (MST). Anomalous low-frequency behavior in both mechanical and optical properties are observed at temperatures close to the MST. We conclude that a general condition must be satisfied before unusual low-frequency behavior is observed, namely that the time scale of flow must be less than the relaxation time of structural features perturbed by flow. Furthermore, because of the similarities between the low-frequency dynamics of the present star diblock copolymer and those of linear diblock and triblock copolymers, we also conclude that block copolymer architecture does not play a critical role in determining the low-frequency behavior of these, materials. textcopyright 1994 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Relaxation dynamics of bidisperse temporary networks.\n \n \n \n \n\n\n \n Seidel, U; Stadler, R; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 27(8): 2066–2072. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"RelaxationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{seidel_relaxation_1994,\n\tAbstract = {Simultaneous measurement of infrared dichroism and birefringence is used to examine the component relaxation in bidisperse, fully entangled melts of polybutadienes with 1 \\% of the repeating units modified with 4-phenyl-1,2,4-triazoline-3,5-dione (urazole) groups. Since the urazole groups are capable ot forming interchain hydrogen bonds, the rheological properties of these materials differ significantly from those of the analogous polybutadiene melts without polar stickers. Samples with 30, 50, and 70 \\% long chains are examined by step strain and oscillatory shear experiments. In addition, the step strain experiments are carried out at two temperatures. All samples show a higher orientational coupling coefficient (varepsilon = 0.6-0.72) than corresponding polybutadiene melts (varepsilon = 0.47). The fact that orientational coupling depends on temperature and sample composition indicates that the relaxation of the chain containing stickers is strongly coupled. This feature is not incorporated in the current theory describing the relaxation of the chain containing stickers. 8 textcopyright 1994 American Chemical Society.},\n\tAuthor = {Seidel, U and Stadler, R and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {8},\n\tPages = {2066--2072},\n\tTitle = {Relaxation dynamics of bidisperse temporary networks},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028405799&partnerID=40&md5=ad82a2ad8907d05aa8c72a3b206607f9},\n\tVolume = {27},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028405799&partnerID=40&md5=ad82a2ad8907d05aa8c72a3b206607f9}}\n\n
\n
\n\n\n
\n Simultaneous measurement of infrared dichroism and birefringence is used to examine the component relaxation in bidisperse, fully entangled melts of polybutadienes with 1 % of the repeating units modified with 4-phenyl-1,2,4-triazoline-3,5-dione (urazole) groups. Since the urazole groups are capable ot forming interchain hydrogen bonds, the rheological properties of these materials differ significantly from those of the analogous polybutadiene melts without polar stickers. Samples with 30, 50, and 70 % long chains are examined by step strain and oscillatory shear experiments. In addition, the step strain experiments are carried out at two temperatures. All samples show a higher orientational coupling coefficient (varepsilon = 0.6-0.72) than corresponding polybutadiene melts (varepsilon = 0.47). The fact that orientational coupling depends on temperature and sample composition indicates that the relaxation of the chain containing stickers is strongly coupled. This feature is not incorporated in the current theory describing the relaxation of the chain containing stickers. 8 textcopyright 1994 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Structure and dynamics of concentration fluctuations in a polymer blend solution under shear flow.\n \n \n \n\n\n \n Lai, J.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Polymer Science, Part B: Polymer Physics, 32(New York, NY, United States): 2461–2474. 1994.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lai_structure_1994,\n\tAuthor = {Lai, Janet and Fuller, Gerald G},\n\tJournal = {Journal of Polymer Science, Part B: Polymer Physics},\n\tNumber = {New York, NY, United States},\n\tPages = {2461--2474},\n\tTitle = {Structure and dynamics of concentration fluctuations in a polymer blend solution under shear flow},\n\tVolume = {32},\n\tYear = {1994}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Structure and optical anisotropies of critical polymer solutions in electric fields.\n \n \n \n \n\n\n \n Wirtz, D; Werner, D E; and Fuller, G.\n\n\n \n\n\n\n The Journal of Chemical Physics, 101(2): 1679–1686. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"StructurePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wirtz_structure_1994,\n\tAbstract = {We present simultaneous measurements of both the small angle light scattering (SALS) structure factor and the dichroism for near-critical polymer solutions in electric fields. The theory of flow birefringence and dichroism proposed by Onuki and Doi is extended to the electric field case. This theory relates the electric-field-induced form dichroism to the second moment of the structure factor. Excellent agreement is observed between the form dichroism calculated from measured scattering intensities and the dichroism measured by optical polarimetry. In addition, a possible mechanism of dipolar coupling between the large critical concentration fluctuations and the electric field is proposed. Similar simultaneous SALS and birefringence measurements confirm this model of dipolar interaction. textcopyright 1994 American Institute of Physics.},\n\tAuthor = {Wirtz, D and Werner, D E and Fuller, G.G.},\n\tJournal = {The Journal of Chemical Physics},\n\tNumber = {2},\n\tPages = {1679--1686},\n\tTitle = {Structure and optical anisotropies of critical polymer solutions in electric fields},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001021647&partnerID=40&md5=94db56b4e1f2d34e6bc320ed7a46abee},\n\tVolume = {101},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001021647&partnerID=40&md5=94db56b4e1f2d34e6bc320ed7a46abee}}\n\n
\n
\n\n\n
\n We present simultaneous measurements of both the small angle light scattering (SALS) structure factor and the dichroism for near-critical polymer solutions in electric fields. The theory of flow birefringence and dichroism proposed by Onuki and Doi is extended to the electric field case. This theory relates the electric-field-induced form dichroism to the second moment of the structure factor. Excellent agreement is observed between the form dichroism calculated from measured scattering intensities and the dichroism measured by optical polarimetry. In addition, a possible mechanism of dipolar coupling between the large critical concentration fluctuations and the electric field is proposed. Similar simultaneous SALS and birefringence measurements confirm this model of dipolar interaction. textcopyright 1994 American Institute of Physics.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Orientation dynamics of a polymer melt studied by polarization-modulated laser Raman scattering.\n \n \n \n \n\n\n \n Archer, L. A; Huang, K.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Rheology, 38(Woodbury, NY, United States): 1101–1125. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"OrientationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{archer_orientation_1994,\n\tAbstract = {The orientation dynamics of an entangled poly(isobutylene) melt during step uniaxial extension was studied by means polarization-modulated laser Raman scattering. Results showed that anisotropies in the second and fourth moments of the orientation distribution of chain segment were recovered from the Raman-scattered light associated with the vs stretching vibration modes of carbon-carbon and carbon-hydrogen groups present in the specimen. The respective second and fourth moments recovered from carbon-carbon and carbon-hydrogen agreed with each other reasonably well. Furthermore, the strain-dependent anisotropies in the second moment compared quite well with those recovered from simultaneous birefringence measurement.},\n\tAuthor = {Archer, Lynden A and Huang, Kelly and Fuller, Gerald G},\n\tJournal = {Journal of Rheology},\n\tNumber = {Woodbury, NY, United States},\n\tPages = {1101--1125},\n\tTitle = {Orientation dynamics of a polymer melt studied by polarization-modulated laser {Raman} scattering},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028478869&partnerID=40&md5=ed03433169900ed223fd8d184cb294f8},\n\tVolume = {38},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0028478869&partnerID=40&md5=ed03433169900ed223fd8d184cb294f8}}\n\n
\n
\n\n\n
\n The orientation dynamics of an entangled poly(isobutylene) melt during step uniaxial extension was studied by means polarization-modulated laser Raman scattering. Results showed that anisotropies in the second and fourth moments of the orientation distribution of chain segment were recovered from the Raman-scattered light associated with the vs stretching vibration modes of carbon-carbon and carbon-hydrogen groups present in the specimen. The respective second and fourth moments recovered from carbon-carbon and carbon-hydrogen agreed with each other reasonably well. Furthermore, the strain-dependent anisotropies in the second moment compared quite well with those recovered from simultaneous birefringence measurement.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Monolayers of perfluoropolyethers with a hydrophilic head group.\n \n \n \n \n\n\n \n Goedel, W A; Wu, H; Friedenberg, M C; Fuller, G.; Foster, M; and Frank, C.\n\n\n \n\n\n\n Langmuir, 10(11): 4209–4218. 1994.\n \n\n\n\n
\n\n\n\n \n \n \"MonolayersPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{goedel_monolayers_1994,\n\tAbstract = {Perfluoropoly(oxypropylene),F((CF 2) 3O) n((CF 2) 2)X, with a hydrophilic head group (X = -COOH) has been investigated as a thin film spread onto the surface of water and as films transferred to solid substrates via the Langmuir-Blodgett technique. Between the onset of the isotherm and collapse, the polymer forms a continuous film of uniform thickness. The density of this film is the same as in the bulk polymer, and the film thickness depends on the area per molecule. The theory of tethered polymers predicts an entropic contribution to the free energy of the system due to deformation of the polymer coils. We propose a thermodynamic analysis in which we mathematically separate the contributions to the surface pressure from the head groups and from the stretching of the polymer chains. In addition, we investigate the influence of the subphase composition (aqueous solutions of CaCl 2, poly(ethylenimine), HCl, FeCl 3, NH 3) and compare the isotherm with a model system for the isolated head group. Based on the thermodynamic analysis and the comparison of isotherms, we conclude that-in the case of a chain length of approximately 100 σ-bonds - there is no significant entropic contribution from the polymer chains. textcopyright 1994 American Chemical Society.},\n\tAuthor = {Goedel, W A and Wu, H and Friedenberg, M C and Fuller, G.G. and Foster, M and Frank, C.W.},\n\tJournal = {Langmuir},\n\tNumber = {11},\n\tPages = {4209--4218},\n\tTitle = {Monolayers of perfluoropolyethers with a hydrophilic head group},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000609308&partnerID=40&md5=9243bbaddecd53e345b4448faa7b94a5},\n\tVolume = {10},\n\tYear = {1994},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000609308&partnerID=40&md5=9243bbaddecd53e345b4448faa7b94a5}}\n\n
\n
\n\n\n
\n Perfluoropoly(oxypropylene),F((CF 2) 3O) n((CF 2) 2)X, with a hydrophilic head group (X = -COOH) has been investigated as a thin film spread onto the surface of water and as films transferred to solid substrates via the Langmuir-Blodgett technique. Between the onset of the isotherm and collapse, the polymer forms a continuous film of uniform thickness. The density of this film is the same as in the bulk polymer, and the film thickness depends on the area per molecule. The theory of tethered polymers predicts an entropic contribution to the free energy of the system due to deformation of the polymer coils. We propose a thermodynamic analysis in which we mathematically separate the contributions to the surface pressure from the head groups and from the stretching of the polymer chains. In addition, we investigate the influence of the subphase composition (aqueous solutions of CaCl 2, poly(ethylenimine), HCl, FeCl 3, NH 3) and compare the isotherm with a model system for the isolated head group. Based on the thermodynamic analysis and the comparison of isotherms, we conclude that-in the case of a chain length of approximately 100 σ-bonds - there is no significant entropic contribution from the polymer chains. textcopyright 1994 American Chemical Society.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1993\n \n \n (7)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Flow-induced concentration fluctuations in polymer solutions: Structure/property relationships.\n \n \n \n \n\n\n \n Moldenaers, P; Yanase, H; Mewis, J; Fuller, G.; Lee, C S; and Magda, J J\n\n\n \n\n\n\n Rheologica Acta, 32(1): 1–8. 1993.\n \n\n\n\n
\n\n\n\n \n \n \"Flow-inducedPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{moldenaers_flow-induced_1993,\n\tAbstract = {Mechanical and optical rheometric measurements are reported on solutions of polystyrene dissolved in dioctyl phthalate, a solution that can undergo an apparent phase separation upon the application of shear. Solutions prepared using three molecular weights ranging from one to four million were studied. Time-temperature superposition was observed to apply for these solutions up to and including the onset of an apparent shear thickening of the steady shear and first normal stresses. Optical measurements employing turbidity and scattering dichroism determined that concentration fluctuations were enhanced by flow and grew parallel to the vorticity axis below the critical velocity gradient for the onset of the apparent shear thickening effect. Prior to the onset of thickening, the fluctuations were observed to rearrange and orient parallel to the flow direction. Second normal stress difference measurements indicate these solutions have a negative ratio of the second to the first normal stress differences. It is interesting to point out that the ratio tends to zero in the vicinity of the shear rate range at which shear thickening occurs. textcopyright 1993 Steinkopff-Verlag.},\n\tAuthor = {Moldenaers, P and Yanase, H and Mewis, J and Fuller, G.G. and Lee, C S and Magda, J J},\n\tJournal = {Rheologica Acta},\n\tNumber = {1},\n\tPages = {1--8},\n\tTitle = {Flow-induced concentration fluctuations in polymer solutions: {Structure}/property relationships},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001979872&partnerID=40&md5=7b5d10d0244ab8df8a0d1b222768b58a},\n\tVolume = {32},\n\tYear = {1993},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001979872&partnerID=40&md5=7b5d10d0244ab8df8a0d1b222768b58a}}\n\n
\n
\n\n\n
\n Mechanical and optical rheometric measurements are reported on solutions of polystyrene dissolved in dioctyl phthalate, a solution that can undergo an apparent phase separation upon the application of shear. Solutions prepared using three molecular weights ranging from one to four million were studied. Time-temperature superposition was observed to apply for these solutions up to and including the onset of an apparent shear thickening of the steady shear and first normal stresses. Optical measurements employing turbidity and scattering dichroism determined that concentration fluctuations were enhanced by flow and grew parallel to the vorticity axis below the critical velocity gradient for the onset of the apparent shear thickening effect. Prior to the onset of thickening, the fluctuations were observed to rearrange and orient parallel to the flow direction. Second normal stress difference measurements indicate these solutions have a negative ratio of the second to the first normal stress differences. It is interesting to point out that the ratio tends to zero in the vicinity of the shear rate range at which shear thickening occurs. textcopyright 1993 Steinkopff-Verlag.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Phase transitions induced by electric fields in near-critical polymer solutions.\n \n \n \n \n\n\n \n Wirtz, D; and Fuller, G.\n\n\n \n\n\n\n Physical Review Letters, 71(14): 2236–2239. 1993.\n \n\n\n\n
\n\n\n\n \n \n \"PhasePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wirtz_phase_1993,\n\tAbstract = {A time-dependent small-angle light scattering (SALS) study is presented which demonstrates the feasibility of electric-field-induced phase transitions in nonionic binary liquids near the critical point. Mixing of two-phase solutions induced by electric fields is shown to be a universal feature shared by a wide class of systems including upper and lower critical point polymer solutions and mixtures of low-molecular-weight molecules in a solvent. The shift of the critical temperature is computed from SALS measurements and is quadratic in the electric-field strength. textcopyright 1993 The American Physical Society.},\n\tAuthor = {Wirtz, D and Fuller, G.G.},\n\tJournal = {Physical Review Letters},\n\tNumber = {14},\n\tPages = {2236--2239},\n\tTitle = {Phase transitions induced by electric fields in near-critical polymer solutions},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001312769&partnerID=40&md5=ebcd726e8a783f4b36d1cbb8c74e390e},\n\tVolume = {71},\n\tYear = {1993},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001312769&partnerID=40&md5=ebcd726e8a783f4b36d1cbb8c74e390e}}\n\n
\n
\n\n\n
\n A time-dependent small-angle light scattering (SALS) study is presented which demonstrates the feasibility of electric-field-induced phase transitions in nonionic binary liquids near the critical point. Mixing of two-phase solutions induced by electric fields is shown to be a universal feature shared by a wide class of systems including upper and lower critical point polymer solutions and mixtures of low-molecular-weight molecules in a solvent. The shift of the critical temperature is computed from SALS measurements and is quadratic in the electric-field strength. textcopyright 1993 The American Physical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Electric Field Induced Structure in Dense Suspensions.\n \n \n \n \n\n\n \n Smith, K L; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 155(1): 183–190. 1993.\n \n\n\n\n
\n\n\n\n \n \n \"ElectricPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{smith_electric_1993,\n\tAbstract = {This paper presents measurements of the structure of dense suspensions of colloidal spheres subject to electric fields. Scattering dichroism measurements, performed as a function of wavelength, probed the anisotropy in the pair distribution function of submicrometer silica spheres, following the application of electric fields. The results indicate that at lower concentrations, high aspect ratio fibrils are formed upon application of the electric fields. As the concentration is increased, branching of the fibrillar structure reduces the anisotropy of the suspension with a subsequent decrease of the measured dichroism. The time scales of structure formation and relaxation are also considered. The response time of the fluid decreases as the applied field is increased and as the particle volume fraction is decreased. The relaxation time increases with increasing particle size but is insensitive to concentration. textcopyright 1993 Academic Press. All rights reserved.},\n\tAuthor = {Smith, K L and Fuller, G.G.},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {1},\n\tPages = {183--190},\n\tTitle = {Electric {Field} {Induced} {Structure} in {Dense} {Suspensions}},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0042317232&partnerID=40&md5=d95f45a052221b51822911b168bb01a3},\n\tVolume = {155},\n\tYear = {1993},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0042317232&partnerID=40&md5=d95f45a052221b51822911b168bb01a3}}\n\n
\n
\n\n\n
\n This paper presents measurements of the structure of dense suspensions of colloidal spheres subject to electric fields. Scattering dichroism measurements, performed as a function of wavelength, probed the anisotropy in the pair distribution function of submicrometer silica spheres, following the application of electric fields. The results indicate that at lower concentrations, high aspect ratio fibrils are formed upon application of the electric fields. As the concentration is increased, branching of the fibrillar structure reduces the anisotropy of the suspension with a subsequent decrease of the measured dichroism. The time scales of structure formation and relaxation are also considered. The response time of the fluid decreases as the applied field is increased and as the particle volume fraction is decreased. The relaxation time increases with increasing particle size but is insensitive to concentration. textcopyright 1993 Academic Press. All rights reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Concentration fluctuation enhancement in polymer solutions by extensional flow.\n \n \n \n \n\n\n \n Van Egmond, J W; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 26(26): 7182–7188. 1993.\n \n\n\n\n
\n\n\n\n \n \n \"ConcentrationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{van_egmond_concentration_1993,\n\tAbstract = {When subjected to flow, various polymer solutions undergo an "apparent" change of phase, which is manifested by a dramatic increase in the turbidity. This phenomenon is the result of anisotropic flow-induced growth of concentration fluctuations, which can be predicted qualitatively by the coupled equations of motion for the concentration and velocity fields. We have investigated the growth of concentration fluctuations of poor and near-\\$þeta\\$ semidilute solutions of polystyrene in dioctyl phthalate (PS/DOP) subject to plane extensional flow at temperatures above the quiescent cloud point. Scattering dichroism and small-angle light scattering (SALS) indicate that concentration fluctuations grow perpendicular to the principal axis of extension for low strain rates. For higher extensional rates, fourfold symmetry appears in the structure factor, with intensity maxima on the axes at 45textdegree to the principal axes. This fourfold symmetry is predicted by a recent theoretical model. We also show that the strength of flow-induced scattering is linear to first order in the ratio of viscoelastic stress to osmotic pressure. textcopyright 1993 American Chemical Society.},\n\tAuthor = {Van Egmond, J W and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {26},\n\tPages = {7182--7188},\n\tTitle = {Concentration fluctuation enhancement in polymer solutions by extensional flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0027840348&partnerID=40&md5=61fedcf01c8c8b48c570705fb0d03fb4},\n\tVolume = {26},\n\tYear = {1993},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0027840348&partnerID=40&md5=61fedcf01c8c8b48c570705fb0d03fb4}}\n\n
\n
\n\n\n
\n When subjected to flow, various polymer solutions undergo an \"apparent\" change of phase, which is manifested by a dramatic increase in the turbidity. This phenomenon is the result of anisotropic flow-induced growth of concentration fluctuations, which can be predicted qualitatively by the coupled equations of motion for the concentration and velocity fields. We have investigated the growth of concentration fluctuations of poor and near-$þeta$ semidilute solutions of polystyrene in dioctyl phthalate (PS/DOP) subject to plane extensional flow at temperatures above the quiescent cloud point. Scattering dichroism and small-angle light scattering (SALS) indicate that concentration fluctuations grow perpendicular to the principal axis of extension for low strain rates. For higher extensional rates, fourfold symmetry appears in the structure factor, with intensity maxima on the axes at 45textdegree to the principal axes. This fourfold symmetry is predicted by a recent theoretical model. We also show that the strength of flow-induced scattering is linear to first order in the ratio of viscoelastic stress to osmotic pressure. textcopyright 1993 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Scattering Dichroism Measurements of Flow-Induced Structure of a Shear Thickening Suspension.\n \n \n \n \n\n\n \n D'Haene, P; Mewis, J; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 156(2): 350–358. 1993.\n \n\n\n\n
\n\n\n\n \n \n \"ScatteringPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{dhaene_scattering_1993,\n\tAbstract = {Shear thickening in concentrated colloidal dispersions of sterically stabilized PMMA particles has been investigated systematically, both mechanically and optically. Two types of shear thickening were distinguished. At the highest volume fractions, a discontinuous increase of the viscosity occurs at some critical point (in either the value of the stress or the velocity gradient). A further increase of the applied stress results in an erratic flow behavior. At lower concentrations the viscosity increases more gradually and the flow remains smooth. Dichroism experiments performed on a flowing dispersion and probing the structure in the 1-2 plane of simple shear flow do not show marked differences between a dispersion sheared in the upper Newtonian region and a continuously shear thickening sample. Moreover, it was also demonstrated that the type of shear thickening has no influence on the equilibrium values of the dichroism. However, a discontinuously shear thickening sample, sheared beyond the critical conditions, shows erratic, time-dependent transitions. The relaxation of the dichroism, after cessation of the flow, shows a marked increase in time scale compared to the relaxation time for shear rates smaller than the critical value. This enhanced relaxation process suggests that the discontinuity in viscosity is associated with the formation of very large, anisotropic aggregates of particles, which relax by both disorientation and disintegration. This behavior is closely related to the relaxation of the shear stress for a sample sheared beyond the critical conditions, which also shows two relaxation mechanisms. textcopyright 1993 Academic Press. All rights reserved.},\n\tAuthor = {D'Haene, P and Mewis, J and Fuller, G.G.},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {2},\n\tPages = {350--358},\n\tTitle = {Scattering {Dichroism} {Measurements} of {Flow}-{Induced} {Structure} of a {Shear} {Thickening} {Suspension}},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001569995&partnerID=40&md5=0ec02113be7d1965b69a553b9bc829d9},\n\tVolume = {156},\n\tYear = {1993},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001569995&partnerID=40&md5=0ec02113be7d1965b69a553b9bc829d9}}\n\n
\n
\n\n\n
\n Shear thickening in concentrated colloidal dispersions of sterically stabilized PMMA particles has been investigated systematically, both mechanically and optically. Two types of shear thickening were distinguished. At the highest volume fractions, a discontinuous increase of the viscosity occurs at some critical point (in either the value of the stress or the velocity gradient). A further increase of the applied stress results in an erratic flow behavior. At lower concentrations the viscosity increases more gradually and the flow remains smooth. Dichroism experiments performed on a flowing dispersion and probing the structure in the 1-2 plane of simple shear flow do not show marked differences between a dispersion sheared in the upper Newtonian region and a continuously shear thickening sample. Moreover, it was also demonstrated that the type of shear thickening has no influence on the equilibrium values of the dichroism. However, a discontinuously shear thickening sample, sheared beyond the critical conditions, shows erratic, time-dependent transitions. The relaxation of the dichroism, after cessation of the flow, shows a marked increase in time scale compared to the relaxation time for shear rates smaller than the critical value. This enhanced relaxation process suggests that the discontinuity in viscosity is associated with the formation of very large, anisotropic aggregates of particles, which relax by both disorientation and disintegration. This behavior is closely related to the relaxation of the shear stress for a sample sheared beyond the critical conditions, which also shows two relaxation mechanisms. textcopyright 1993 Academic Press. All rights reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Orientation dynamics of side chain polymers subject to electric fields. Part I. Steady state.\n \n \n \n\n\n \n Hoffmann, C L; Man, H T; and Fuller, G.\n\n\n \n\n\n\n Acta Polymerica, 44(1): 39–49. 1993.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hoffmann_orientation_1993,\n\tAbstract = {Many polymeric glasses must be orientationally ordered for the materials to exhibit useful characteristics. The objective of this work was to understand the dynamics of the orientation and relaxation processes and to relate these behaviours to the structural characteristics of the polymeric material. In this study we considered materials suitable for nonlinear optical applications. In particular, thin polymer films, either doped with nonlinear optical chromophores or films with the chromophores covalently bonded to side groups were studied. Orientation of these chromophores was achieved by the application of an external electric field. Ellipsometry was used to measure the birefringence resulting from chromophore orientation and the steady state and transient behaviours were studied. Materials characteristics studied included method of chromophore incorporation, chromophore concentration, spacer length. and molecular weight. Our results showed that the polymer structure, specifically the tacticity, was the dominant factor determining the steady state and transient behaviours. The relationship between monomer and polymer structure and the resulting effect on the orientation dynamics is discussed. Qualitative agreement with theoretical predictions was obtained for the simpler guest/host system, but the model was shown to be inadequate to explain the behaviours of the more complex systems studied. The results reported here have provided valuable insight into the molecular processes that occur during the chromophore orientation and relaxation, and the effects of structural characteristics of the polymers on these mechanisms.},\n\tAuthor = {Hoffmann, C L and Man, H T and Fuller, G.G.},\n\tJournal = {Acta Polymerica},\n\tNumber = {1},\n\tPages = {39--49},\n\tTitle = {Orientation dynamics of side chain polymers subject to electric fields. {Part} {I}. {Steady} state},\n\tVolume = {44},\n\tYear = {1993}}\n\n
\n
\n\n\n
\n Many polymeric glasses must be orientationally ordered for the materials to exhibit useful characteristics. The objective of this work was to understand the dynamics of the orientation and relaxation processes and to relate these behaviours to the structural characteristics of the polymeric material. In this study we considered materials suitable for nonlinear optical applications. In particular, thin polymer films, either doped with nonlinear optical chromophores or films with the chromophores covalently bonded to side groups were studied. Orientation of these chromophores was achieved by the application of an external electric field. Ellipsometry was used to measure the birefringence resulting from chromophore orientation and the steady state and transient behaviours were studied. Materials characteristics studied included method of chromophore incorporation, chromophore concentration, spacer length. and molecular weight. Our results showed that the polymer structure, specifically the tacticity, was the dominant factor determining the steady state and transient behaviours. The relationship between monomer and polymer structure and the resulting effect on the orientation dynamics is discussed. Qualitative agreement with theoretical predictions was obtained for the simpler guest/host system, but the model was shown to be inadequate to explain the behaviours of the more complex systems studied. The results reported here have provided valuable insight into the molecular processes that occur during the chromophore orientation and relaxation, and the effects of structural characteristics of the polymers on these mechanisms.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Investigation of xanthan gum solution behavior under shear flow using rheooptical techniques.\n \n \n \n \n\n\n \n Meyer, E L; Fuller, G.; Clark, R C; and Kulicke, W M\n\n\n \n\n\n\n Macromolecules, 26(3): 504–511. 1993.\n \n\n\n\n
\n\n\n\n \n \n \"InvestigationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{meyer_investigation_1993,\n\tAbstract = {The complementary rheooptical techniques of birefringence and dichroism are used to investigate the dynamics of xanthan gum in dilute/semidilute solution. Because xanthan self-aggregates, there are a range of very different length scales of molecules in solution which the optical methods probe. Birefringence results give information about segmental orientation in flow while scattering dichroism allows interrogation of the much larger aggregate orientation. Two xanthan samples are studied and the results are compared with previous studies. Birefringence and dichroism measurements show very different orientational and relaxational responses in shear flow, indicative of the differences between the structures examined by the two techniques. Dichroism steady-state orientation angle and birefringence/dichroism relaxation results also indicate a critical concentration at approximately 2000 ppm which coincides with that found in the literature by several other investigators and explained by some to be a critical concentration of aggregation. Dichroism measurements detect the presence of aggregates though at concentrations as low as 50 ppm, far below this critical concentration. Overshoots in dichroism, birefringence, and stress signals at approximately 2 strain units are also discussed in light of this aggregation phenomenon. The addition of salt was found to greatly enhance the dichroism signal, providing further evidence that the dichroism signal monitors aggregate dynamics. textcopyright 1993 American Chemical Society.},\n\tAuthor = {Meyer, E L and Fuller, G.G. and Clark, R C and Kulicke, W M},\n\tJournal = {Macromolecules},\n\tNumber = {3},\n\tPages = {504--511},\n\tTitle = {Investigation of xanthan gum solution behavior under shear flow using rheooptical techniques},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0027540891&partnerID=40&md5=bd2001e2961e52950f370590ab74d691},\n\tVolume = {26},\n\tYear = {1993},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0027540891&partnerID=40&md5=bd2001e2961e52950f370590ab74d691}}\n\n
\n
\n\n\n
\n The complementary rheooptical techniques of birefringence and dichroism are used to investigate the dynamics of xanthan gum in dilute/semidilute solution. Because xanthan self-aggregates, there are a range of very different length scales of molecules in solution which the optical methods probe. Birefringence results give information about segmental orientation in flow while scattering dichroism allows interrogation of the much larger aggregate orientation. Two xanthan samples are studied and the results are compared with previous studies. Birefringence and dichroism measurements show very different orientational and relaxational responses in shear flow, indicative of the differences between the structures examined by the two techniques. Dichroism steady-state orientation angle and birefringence/dichroism relaxation results also indicate a critical concentration at approximately 2000 ppm which coincides with that found in the literature by several other investigators and explained by some to be a critical concentration of aggregation. Dichroism measurements detect the presence of aggregates though at concentrations as low as 50 ppm, far below this critical concentration. Overshoots in dichroism, birefringence, and stress signals at approximately 2 strain units are also discussed in light of this aggregation phenomenon. The addition of salt was found to greatly enhance the dichroism signal, providing further evidence that the dichroism signal monitors aggregate dynamics. textcopyright 1993 American Chemical Society.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1992\n \n \n (7)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Oligomers as molecular probes of orientational coupling interactions in polymer melts and networks.\n \n \n \n \n\n\n \n Ylitalo, C M; Zawada, J A; Fuller, G.; Abetz, V; and Stadler, R\n\n\n \n\n\n\n Polymer, 33(14): 2949–2960. 1992.\n \n\n\n\n
\n\n\n\n \n \n \"OligomersPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{ylitalo_oligomers_1992,\n\tAbstract = {Simultaneous measurement of infrared dichroism and birefringence was used to study orientational coupling effects between a series of monodisperse polybutadiene oligomers and the surrounding matrix, which consisted of either polybutadiene polymer melt or polybutadiene network. The oligomers had molecular weights ranging from well below the critical molecular weight for entanglement Me, to several times Me. For the oligomer/network system, the magnitude of the coupling coefficient was found to remain constant at about unity over the range of network crosslink densities, oligomer molecular weights and degrees of swelling examined, thereby confirming the D n.m.r. experiments performed on the same samples. For the oligomer/polymer system, orientational coupling effects were examined as a function of oligomer molecular weight and temperature. It was found that unentangled oligomers experienced almost complete coupling, while fully entangled melts were subject to a much lower coupling. In addition, it was found that the magnitude of the coupling is independent of temperature, indicating that these orientation correlations are entropic in nature, arising from the packing entropy of chain segments. An orientational dependent lattice model, proposed originally by DiMarzio, was found suitable for predicting experimental results. textcopyright 1992.},\n\tAuthor = {Ylitalo, C M and Zawada, J A and Fuller, G.G. and Abetz, V and Stadler, R},\n\tJournal = {Polymer},\n\tNumber = {14},\n\tPages = {2949--2960},\n\tTitle = {Oligomers as molecular probes of orientational coupling interactions in polymer melts and networks},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-37949000959&partnerID=40&md5=877276ff9a6235437327e43868cdd3ff},\n\tVolume = {33},\n\tYear = {1992},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-37949000959&partnerID=40&md5=877276ff9a6235437327e43868cdd3ff}}\n\n
\n
\n\n\n
\n Simultaneous measurement of infrared dichroism and birefringence was used to study orientational coupling effects between a series of monodisperse polybutadiene oligomers and the surrounding matrix, which consisted of either polybutadiene polymer melt or polybutadiene network. The oligomers had molecular weights ranging from well below the critical molecular weight for entanglement Me, to several times Me. For the oligomer/network system, the magnitude of the coupling coefficient was found to remain constant at about unity over the range of network crosslink densities, oligomer molecular weights and degrees of swelling examined, thereby confirming the D n.m.r. experiments performed on the same samples. For the oligomer/polymer system, orientational coupling effects were examined as a function of oligomer molecular weight and temperature. It was found that unentangled oligomers experienced almost complete coupling, while fully entangled melts were subject to a much lower coupling. In addition, it was found that the magnitude of the coupling is independent of temperature, indicating that these orientation correlations are entropic in nature, arising from the packing entropy of chain segments. An orientational dependent lattice model, proposed originally by DiMarzio, was found suitable for predicting experimental results. textcopyright 1992.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Time-dependent small-angle light scattering of shear-induced concentration fluctuations in polymer solutions.\n \n \n \n \n\n\n \n Van Egmond, J W; Werner, D E; and Fuller, G.\n\n\n \n\n\n\n The Journal of Chemical Physics, 96(10): 7742–7757. 1992.\n \n\n\n\n
\n\n\n\n \n \n \"Time-dependentPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{van_egmond_time-dependent_1992,\n\tAbstract = {A small-angle light-scattering experiment probing the time-dependent behavior of a semidilute polymer solution under simple shear is presented. Specifically, this paper investigates time-dependent anisotropic growth and orientation of concentration fluctuations in a semidilute solution of polystyrene in dioctyl phthalate above the quiescent cloud-point temperature. Two flow cells, which permitted investigation of the structure and dynamics of concentration fluctuations in both the flow-shear and the flow-vorticity planes, were used. Concentration fluctuations were found to be enhanced anisotropically by the shear flow in both planes. In the flow-vorticity plane, shear-enhanced concentration fluctuations orient along the voracity direction, while in the flow-shear plane, the axis of average orientation is at an angle between 135textdegree and 45textdegree to the direction of flow. Qualitative trends of this anisotropic growth of concentration fluctuations are in agreement with the predictions of the Helfand-Fredrickson model in the flow-shear plane. Furthermore, the Onuki-Doi theory relating the scattering dichroism to the structure factor has been verified experimentally by comparing scattering dichroism calculated from scattering patterns with previous experimentally measured dichroism. Scattering dichroism, both calculated from structure factors and experimentally determined over a range of shear rates and temperatures, is demonstrated to depend only on the Weissenberg number. Relaxation of shear-enhanced fluctuations on cessation of flow allowed for the calculation of a diffusion coefficient. textcopyright 1992 American Institute of Physics.},\n\tAuthor = {Van Egmond, J W and Werner, D E and Fuller, G.G.},\n\tJournal = {The Journal of Chemical Physics},\n\tNumber = {10},\n\tPages = {7742--7757},\n\tTitle = {Time-dependent small-angle light scattering of shear-induced concentration fluctuations in polymer solutions},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-3743065618&partnerID=40&md5=ff2ce0c64b36861d07cab95190adc8a4},\n\tVolume = {96},\n\tYear = {1992},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-3743065618&partnerID=40&md5=ff2ce0c64b36861d07cab95190adc8a4}}\n\n
\n
\n\n\n
\n A small-angle light-scattering experiment probing the time-dependent behavior of a semidilute polymer solution under simple shear is presented. Specifically, this paper investigates time-dependent anisotropic growth and orientation of concentration fluctuations in a semidilute solution of polystyrene in dioctyl phthalate above the quiescent cloud-point temperature. Two flow cells, which permitted investigation of the structure and dynamics of concentration fluctuations in both the flow-shear and the flow-vorticity planes, were used. Concentration fluctuations were found to be enhanced anisotropically by the shear flow in both planes. In the flow-vorticity plane, shear-enhanced concentration fluctuations orient along the voracity direction, while in the flow-shear plane, the axis of average orientation is at an angle between 135textdegree and 45textdegree to the direction of flow. Qualitative trends of this anisotropic growth of concentration fluctuations are in agreement with the predictions of the Helfand-Fredrickson model in the flow-shear plane. Furthermore, the Onuki-Doi theory relating the scattering dichroism to the structure factor has been verified experimentally by comparing scattering dichroism calculated from scattering patterns with previous experimentally measured dichroism. Scattering dichroism, both calculated from structure factors and experimentally determined over a range of shear rates and temperatures, is demonstrated to depend only on the Weissenberg number. Relaxation of shear-enhanced fluctuations on cessation of flow allowed for the calculation of a diffusion coefficient. textcopyright 1992 American Institute of Physics.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rheo-Optiics for Multicomponent Liquids.\n \n \n \n \n\n\n \n Fuller, G; van Egmond, J; Zawada, J; and Archer, L\n\n\n \n\n\n\n MRS Proceedings, 248: 139. February 1992.\n \n\n\n\n
\n\n\n\n \n \n \"Rheo-OptiicsPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_rheo-optiics_1992,\n\tAuthor = {Fuller, G and van Egmond, J and Zawada, J and Archer, L},\n\tDoi = {10.1557/PROC-248-139},\n\tJournal = {MRS Proceedings},\n\tLanguage = {English},\n\tMonth = feb,\n\tPages = {139},\n\tTitle = {Rheo-{Optiics} for {Multicomponent} {Liquids}},\n\tUrl = {http://journals.cambridge.org/abstract_S1946427400482382},\n\tVolume = {248},\n\tYear = {1992},\n\tBdsk-Url-1 = {http://journals.cambridge.org/abstract_S1946427400482382},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1557/PROC-248-139}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Dynamics of polymeric liquids using polarization-modulated laser Raman scattering.\n \n \n \n \n\n\n \n Archer, L A; Fuller, G.; and Nunnelley, L\n\n\n \n\n\n\n Polymer, 33(17): 3574–3581. 1992.\n \n\n\n\n
\n\n\n\n \n \n \"DynamicsPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{archer_dynamics_1992,\n\tAbstract = {A new experimental technique, 'polarization-modulated laser Raman scattering' (PMLRS), is presented for the study of the dynamics of polymeric materials subjected to transient flows. A detailed analysis of the experiment based on Jones-Mueller matrix calculus is presented. Jones and Mueller matrices that take into account the integrated effect of the deformed sample's birefringence are developed for a Raman scattering element, using either the 0textdegree or 180textdegree scattering geometries. To test the predictions of the theory a Raman scattering system, based on high-speed phase modulation (50 kHz) of the incident light coupled with phase-sensitive detection, has been developed. This system is also capable of simultaneously measuring birefringence. Single-step reversal extension and step strain experiments were conducted using a polyisobutylene melt. Experiments were done using the Raman scattered light associated with the v stretching vibration of C-H groups present in the specimen. The results obtained indicate that the Raman signal ratios all depend on the second- and fourth-order moments of the orientation distribution function of polymer segments and on the sample's birefringence. textcopyright 1992.},\n\tAuthor = {Archer, L A and Fuller, G.G. and Nunnelley, L},\n\tJournal = {Polymer},\n\tNumber = {17},\n\tPages = {3574--3581},\n\tTitle = {Dynamics of polymeric liquids using polarization-modulated laser {Raman} scattering},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0342968345&partnerID=40&md5=de8f9f42688cf01c58a398beb500dcd4},\n\tVolume = {33},\n\tYear = {1992},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0342968345&partnerID=40&md5=de8f9f42688cf01c58a398beb500dcd4}}\n\n
\n
\n\n\n
\n A new experimental technique, 'polarization-modulated laser Raman scattering' (PMLRS), is presented for the study of the dynamics of polymeric materials subjected to transient flows. A detailed analysis of the experiment based on Jones-Mueller matrix calculus is presented. Jones and Mueller matrices that take into account the integrated effect of the deformed sample's birefringence are developed for a Raman scattering element, using either the 0textdegree or 180textdegree scattering geometries. To test the predictions of the theory a Raman scattering system, based on high-speed phase modulation (50 kHz) of the incident light coupled with phase-sensitive detection, has been developed. This system is also capable of simultaneously measuring birefringence. Single-step reversal extension and step strain experiments were conducted using a polyisobutylene melt. Experiments were done using the Raman scattered light associated with the v stretching vibration of C-H groups present in the specimen. The results obtained indicate that the Raman signal ratios all depend on the second- and fourth-order moments of the orientation distribution function of polymer segments and on the sample's birefringence. textcopyright 1992.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Comparison of numerical simulations and birefringence measurements in viscoelastic flow between eccentric rotating cylinders.\n \n \n \n \n\n\n \n Rajagopalan, D.\n\n\n \n\n\n\n Journal of Rheology, 36(7): 1349. October 1992.\n \n\n\n\n
\n\n\n\n \n \n \"ComparisonPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{rajagopalan_comparison_1992,\n\tAbstract = {Finite-element predictions of steady, two-dimensional viscoelasticflow based on the elastic-viscous split stress formulation of Rajagopalan et al. [Rajagopalan, D., R. C. Armstrong, and R. A. Brown, ``Finite Element Methods for Calculation of Steady, ViscoelasticFlow Using Constitutive Equations with a Newtonian Viscosity,'' J. Non-Newt. Fluid Mech. 36, 159--192 (1990)] are compared to birefringencemeasurements of stress for flow between eccentric rotating cylinders of a solution of polystyrene in tricresyl phosphate. Steady and small-amplitude oscillatory shearing measurements are used to characterize the shear flow rheology of the fluid, and a four-mode Giesekus model with a Newtonian solvent viscosity is fit to the rheological data. Comparison of the stress fields measured by flow-induced birefringence in the one-dimensional flow between concentric cylinders to the mechanically measured viscometric data show that the optical data compare reasonably well with the mechanical data. Comparison of optical measurements and finite-element predictions in the two-dimensional, eccentric cylinder flow, show good pointwise agreement throughout most regions of the flow.},\n\tAuthor = {Rajagopalan, Dilip},\n\tDoi = {10.1122/1.550266},\n\tJournal = {Journal of Rheology},\n\tMonth = oct,\n\tNumber = {7},\n\tPages = {1349},\n\tTitle = {Comparison of numerical simulations and birefringence measurements in viscoelastic flow between eccentric rotating cylinders},\n\tUrl = {http://scitation.aip.org/content/sor/journal/jor2/36/7/10.1122/1.550266},\n\tVolume = {36},\n\tYear = {1992},\n\tBdsk-Url-1 = {http://scitation.aip.org/content/sor/journal/jor2/36/7/10.1122/1.550266},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1122/1.550266}}\n\n
\n
\n\n\n
\n Finite-element predictions of steady, two-dimensional viscoelasticflow based on the elastic-viscous split stress formulation of Rajagopalan et al. [Rajagopalan, D., R. C. Armstrong, and R. A. Brown, ``Finite Element Methods for Calculation of Steady, ViscoelasticFlow Using Constitutive Equations with a Newtonian Viscosity,'' J. Non-Newt. Fluid Mech. 36, 159–192 (1990)] are compared to birefringencemeasurements of stress for flow between eccentric rotating cylinders of a solution of polystyrene in tricresyl phosphate. Steady and small-amplitude oscillatory shearing measurements are used to characterize the shear flow rheology of the fluid, and a four-mode Giesekus model with a Newtonian solvent viscosity is fit to the rheological data. Comparison of the stress fields measured by flow-induced birefringence in the one-dimensional flow between concentric cylinders to the mechanically measured viscometric data show that the optical data compare reasonably well with the mechanical data. Comparison of optical measurements and finite-element predictions in the two-dimensional, eccentric cylinder flow, show good pointwise agreement throughout most regions of the flow.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Component relaxation dynamics in a miscible polymer blend: poly(ethylene oxide)/poly(methyl methacrylate).\n \n \n \n \n\n\n \n Zawada, J. A; Ylitalo, C. M; Fuller, G. G; Colby, R. H; and Long, T. E\n\n\n \n\n\n\n Macromolecules, 25(11): 2896–2902. May 1992.\n \n\n\n\n
\n\n\n\n \n \n \"ComponentPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{zawada_component_1992,\n\tAuthor = {Zawada, Jeffrey A and Ylitalo, Caroline M and Fuller, Gerald G and Colby, Ralph H and Long, Timothy E},\n\tDoi = {10.1021/ma00037a017},\n\tJournal = {Macromolecules},\n\tMonth = may,\n\tNumber = {11},\n\tPages = {2896--2902},\n\tTitle = {Component relaxation dynamics in a miscible polymer blend: poly(ethylene oxide)/poly(methyl methacrylate)},\n\tUrl = {http://dx.doi.org/10.1021/ma00037a017},\n\tVolume = {25},\n\tYear = {1992},\n\tBdsk-Url-1 = {http://dx.doi.org/10.1021/ma00037a017}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Electric-field-induced structure in polymer solutions near the critical point.\n \n \n \n \n\n\n \n Wirtz, D; Berend, K; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 25(26): 7234–7246. 1992.\n \n\n\n\n
\n\n\n\n \n \n \"Electric-field-inducedPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wirtz_electric-field-induced_1992,\n\tAbstract = {In the vicinity of the coexistence curve, binary solutions exhibit large concentration fluctuations. Such fluctuations can be anisotropically distorted through the application of an electric field, which induces both electric birefringence and enhanced forward light scattering as the system approaches the coexistence curve. Using time-dependent small-angle light scattering (SALS) and form dichroism measurements on semidilute polystyrene/cyclohexane systems (molecular weight range 400 000-1 800 000), two new phenomena have been observed: electric scattering dichroism in the one-phase region and electric-field-induced remixing in the two-phase region. A mean-field theory that predicts the evolution of the scattering patterns in the presence of an electric field above the coexistence curve has been developed. In particular, it is predicted that, in the plane perpendicular to the propagation of the incident light, circular scattering patterns become elliptical patterns having the minor axis along the direction of the applied field. As a consequence, scattering dichroism is shown to be induced in the vicinity of the critical point. Moreover, the same phenomenological model predicts that the presence of an electric field lowers the coexistence curve and results in electric-field-induced remixing of binary mixtures at a temperature below the quiescent coexistence curve. The influence of temperature, molecular weight, and concentration of the polymer in solution as well as the influence of the electric field strength on SALS patterns and induced scattering dichroism measurements is studied. The experimental observations are in fairly good agreement with the trends predicted by a phenomenologically-based, mean-field theory. textcopyright 1992 American Chemical Society.},\n\tAuthor = {Wirtz, D and Berend, K and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {26},\n\tPages = {7234--7246},\n\tTitle = {Electric-field-induced structure in polymer solutions near the critical point},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001390506&partnerID=40&md5=2d7e51e7826a4bcee0f257bd4c7f2c25},\n\tVolume = {25},\n\tYear = {1992},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001390506&partnerID=40&md5=2d7e51e7826a4bcee0f257bd4c7f2c25}}\n\n
\n
\n\n\n
\n In the vicinity of the coexistence curve, binary solutions exhibit large concentration fluctuations. Such fluctuations can be anisotropically distorted through the application of an electric field, which induces both electric birefringence and enhanced forward light scattering as the system approaches the coexistence curve. Using time-dependent small-angle light scattering (SALS) and form dichroism measurements on semidilute polystyrene/cyclohexane systems (molecular weight range 400 000-1 800 000), two new phenomena have been observed: electric scattering dichroism in the one-phase region and electric-field-induced remixing in the two-phase region. A mean-field theory that predicts the evolution of the scattering patterns in the presence of an electric field above the coexistence curve has been developed. In particular, it is predicted that, in the plane perpendicular to the propagation of the incident light, circular scattering patterns become elliptical patterns having the minor axis along the direction of the applied field. As a consequence, scattering dichroism is shown to be induced in the vicinity of the critical point. Moreover, the same phenomenological model predicts that the presence of an electric field lowers the coexistence curve and results in electric-field-induced remixing of binary mixtures at a temperature below the quiescent coexistence curve. The influence of temperature, molecular weight, and concentration of the polymer in solution as well as the influence of the electric field strength on SALS patterns and induced scattering dichroism measurements is studied. The experimental observations are in fairly good agreement with the trends predicted by a phenomenologically-based, mean-field theory. textcopyright 1992 American Chemical Society.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1991\n \n \n (10)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Molecular weight dependence of component dynamics in bidisperse melt rheology.\n \n \n \n \n\n\n \n Ylitalo, C M; Kornfield, J A; Fuller, G.; and Pearson, D S\n\n\n \n\n\n\n Macromolecules, 24(3): 749–758. 1991.\n \n\n\n\n
\n\n\n\n \n \n \"MolecularPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{ylitalo_molecular_1991,\n\tAbstract = {Simultaneous measurement of infrared dichroism and birefringence is used to study component relaxation in bimodal molecular weight distribution melts. Nearly monodisperse poly(ethylenepropylene) samples of molecular weights 53K, 125K, and 370K, all above the critical molecular weight for entanglement, were used. Results for the step-strain relaxation of each component and of the total sample are presented and discussed for binary blends of 10, 20, 30, 50, and 75\\% by volume of the higher molecular weight species for three sets of blends: 53K/125K, 53K/370K, and 125K/370K. For each sample, the relaxation dynamics of the blend and of each component depend upon the two polymer relaxation times and the blend composition. Effects of intermolecular orientational coupling interactions were observed, and a coupling strength of 0.45 was measured. The results of these experiments are compared to a reptation-based constraint release model, and a qualitative agreement is found. textcopyright 1991 American Chemical Society.},\n\tAuthor = {Ylitalo, C M and Kornfield, J A and Fuller, G.G. and Pearson, D S},\n\tJournal = {Macromolecules},\n\tNumber = {3},\n\tPages = {749--758},\n\tTitle = {Molecular weight dependence of component dynamics in bidisperse melt rheology},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026415674&partnerID=40&md5=96e467384e867c5bcbd305c03fa3ab7d},\n\tVolume = {24},\n\tYear = {1991},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026415674&partnerID=40&md5=96e467384e867c5bcbd305c03fa3ab7d}}\n\n
\n
\n\n\n
\n Simultaneous measurement of infrared dichroism and birefringence is used to study component relaxation in bimodal molecular weight distribution melts. Nearly monodisperse poly(ethylenepropylene) samples of molecular weights 53K, 125K, and 370K, all above the critical molecular weight for entanglement, were used. Results for the step-strain relaxation of each component and of the total sample are presented and discussed for binary blends of 10, 20, 30, 50, and 75% by volume of the higher molecular weight species for three sets of blends: 53K/125K, 53K/370K, and 125K/370K. For each sample, the relaxation dynamics of the blend and of each component depend upon the two polymer relaxation times and the blend composition. Effects of intermolecular orientational coupling interactions were observed, and a coupling strength of 0.45 was measured. The results of these experiments are compared to a reptation-based constraint release model, and a qualitative agreement is found. textcopyright 1991 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Field-induced anisotropy in concentrated systems of rigid particles and macromolecules.\n \n \n \n \n\n\n \n Fuller, G.; Smith, K; and Burghardt, W R\n\n\n \n\n\n\n Journal of Statistical Physics, 62(5-6): 1025–1039. 1991.\n \n\n\n\n
\n\n\n\n \n \n \"Field-inducedPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_field-induced_1991,\n\tAbstract = {This paper presents experimental results on the use of spectroscopic optical polarimetry to study structure in dense systems of rigid particles and rigid polymer liquid crystals. These measurements probe microstructural anisotropy induced by the application of electric fields in the case of dense suspensions of rigid spheres, or flow fields in the case of polymer liquid crystals. It is demonstrated that conservative linear dichroism can measure moments of the particle pair distribution function in dense suspensions. In liquid crystals, the dichroism is a result of field-induced anisotropy in the defect structure of the material. textcopyright 1991 Plenum Publishing Corporation.},\n\tAuthor = {Fuller, G.G. and Smith, K and Burghardt, W R},\n\tJournal = {Journal of Statistical Physics},\n\tNumber = {5-6},\n\tPages = {1025--1039},\n\tTitle = {Field-induced anisotropy in concentrated systems of rigid particles and macromolecules},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-34249919776&partnerID=40&md5=040aa89c652087e0ed29829b7da1d78c},\n\tVolume = {62},\n\tYear = {1991},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-34249919776&partnerID=40&md5=040aa89c652087e0ed29829b7da1d78c}}\n\n
\n
\n\n\n
\n This paper presents experimental results on the use of spectroscopic optical polarimetry to study structure in dense systems of rigid particles and rigid polymer liquid crystals. These measurements probe microstructural anisotropy induced by the application of electric fields in the case of dense suspensions of rigid spheres, or flow fields in the case of polymer liquid crystals. It is demonstrated that conservative linear dichroism can measure moments of the particle pair distribution function in dense suspensions. In liquid crystals, the dichroism is a result of field-induced anisotropy in the defect structure of the material. textcopyright 1991 Plenum Publishing Corporation.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Infrared polarimetry studies for multicomponent polymer melts.\n \n \n \n \n\n\n \n Fuller, G.; and Ylitalo, C M\n\n\n \n\n\n\n Journal of Non-Crystalline Solids, 131-133(PART 2): 676–684. 1991.\n \n\n\n\n
\n\n\n\n \n \n \"InfraredPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_infrared_1991,\n\tAbstract = {A rheo-optical technique to measure dynamic infrared dichroism was used with deuterium labeling to study component relaxation in a variety of multicomponent polymer melts. For each sample overall relaxation was observed simultaneously using birefringence measurements. Nearly monodisperse poly(ethylene propylene) samples were used to construct three different sets of bidisperse blends. For each sample, results for step-strain relaxation of each component and of the total sample are presented and discussed. Effects of intermolecular orientational coupling were observed and a coupling strength of 0.45 was measured. The results of these experiments were compared to a reptation-based constraint release model, and a qualitative agreement was found. Using nearly monodisperse polybutadiene block copolymers, selected segment relaxations following step strains were examined. It was found that segments located at chain ends relax faster than segments located at chain centers, and that both the Rouse model and the tube model of Doi and Edwards correctly predict the qualitative features of segmental relaxation. The tube model, however, gave exact quantitative agreement with experiment when the effects of orientational coupling interactions of strength 0.4 were incorporated into the model. Additional studies on orientational coupling effects in polymer melts were conducted using a series of monodisperse polybutadiene oligomers dissolved in a polybutadiene polymer matrix. The oligomers had molecular weights ranging from well below the entanglement molecular weight, Me, to several times Me. The results indicate that orientational coupling interactions are significantly reduced for oligomers with molecular weights above Me. This technique can be further applied to study component dynamics in blends of two different polymers where the relaxation of each polymer in the melt can be observed separately. textcopyright 1991.},\n\tAuthor = {Fuller, G.G. and Ylitalo, C M},\n\tJournal = {Journal of Non-Crystalline Solids},\n\tNumber = {PART 2},\n\tPages = {676--684},\n\tTitle = {Infrared polarimetry studies for multicomponent polymer melts},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026413749&partnerID=40&md5=ed7f67a2fdbfda2041df083ca58d13e9},\n\tVolume = {131-133},\n\tYear = {1991},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026413749&partnerID=40&md5=ed7f67a2fdbfda2041df083ca58d13e9}}\n\n
\n
\n\n\n
\n A rheo-optical technique to measure dynamic infrared dichroism was used with deuterium labeling to study component relaxation in a variety of multicomponent polymer melts. For each sample overall relaxation was observed simultaneously using birefringence measurements. Nearly monodisperse poly(ethylene propylene) samples were used to construct three different sets of bidisperse blends. For each sample, results for step-strain relaxation of each component and of the total sample are presented and discussed. Effects of intermolecular orientational coupling were observed and a coupling strength of 0.45 was measured. The results of these experiments were compared to a reptation-based constraint release model, and a qualitative agreement was found. Using nearly monodisperse polybutadiene block copolymers, selected segment relaxations following step strains were examined. It was found that segments located at chain ends relax faster than segments located at chain centers, and that both the Rouse model and the tube model of Doi and Edwards correctly predict the qualitative features of segmental relaxation. The tube model, however, gave exact quantitative agreement with experiment when the effects of orientational coupling interactions of strength 0.4 were incorporated into the model. Additional studies on orientational coupling effects in polymer melts were conducted using a series of monodisperse polybutadiene oligomers dissolved in a polybutadiene polymer matrix. The oligomers had molecular weights ranging from well below the entanglement molecular weight, Me, to several times Me. The results indicate that orientational coupling interactions are significantly reduced for oligomers with molecular weights above Me. This technique can be further applied to study component dynamics in blends of two different polymers where the relaxation of each polymer in the melt can be observed separately. textcopyright 1991.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Temperature effects on the magnitude of orientational coupling interactions in polymer melts.\n \n \n \n \n\n\n \n Ylitalo, C M; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 24(20): 5736–5737. 1991.\n \n\n\n\n
\n\n\n\n \n \n \"TemperaturePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{ylitalo_temperature_1991,\n\tAuthor = {Ylitalo, C M and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {20},\n\tPages = {5736--5737},\n\tTitle = {Temperature effects on the magnitude of orientational coupling interactions in polymer melts},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026220408&partnerID=40&md5=cf51d15a86f34cf1c61b3fb213b85da8},\n\tVolume = {24},\n\tYear = {1991},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026220408&partnerID=40&md5=cf51d15a86f34cf1c61b3fb213b85da8}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Linear infrared dichroism by a double modulation technique.\n \n \n \n \n\n\n \n Abetz, V.; Fuller, G. G; and Stadler, R.\n\n\n \n\n\n\n Makromolekulare Chemie. Macromolecular Symposia, 52(1): 23–40. December 1991.\n \n\n\n\n
\n\n\n\n \n \n \"LinearPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{abetz_linear_1991,\n\tAuthor = {Abetz, Volker and Fuller, Gerald G and Stadler, Reimund},\n\tDoi = {10.1002/masy.19910520105},\n\tJournal = {Makromolekulare Chemie. Macromolecular Symposia},\n\tMonth = dec,\n\tNumber = {1},\n\tPages = {23--40},\n\tTitle = {Linear infrared dichroism by a double modulation technique},\n\tUrl = {http://doi.wiley.com/10.1002/masy.19910520105},\n\tVolume = {52},\n\tYear = {1991},\n\tBdsk-Url-1 = {http://doi.wiley.com/10.1002/masy.19910520105},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1002/masy.19910520105}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Third normal stress difference and component relaxation spectra for bidisperse melts under oscillatory shear.\n \n \n \n \n\n\n \n Kornfield, J A; Fuller, G.; and Pearson, D S\n\n\n \n\n\n\n Macromolecules, 24(19): 5429–5441. 1991.\n \n\n\n\n
\n\n\n\n \n \n \"ThirdPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kornfield_third_1991,\n\tAbstract = {A recently developed rheooptical technique that simultaneously measures infrared dichroism and birefringence is used to study bidisperse, linear polymer melts in an oscillatory shear. The method was used to individually measure the contribution of the high and low molecular weight components to the third normal stress difference. This was accomplished by measuring the infrared dichroism due to deuterium-labeled polymers. Binary blends are prepared from nearly monodisperse pairs of hydrogenated and deuteriated polyisoprenes of molecular weights 53 000 and 370 000 with volume fractions of long polymer, ΦL, of 0.10, 0.20, 0.30, 0.50, and 0.75. The long-chain relaxation in the blends shows the appearance of a peak for compositions with ΦL le 0.50 at approximately the same frequency as a similar peak in the pure short polymer. The response of the short-chain relaxation in the blends shows the appearance of a low-frequency shoulder that becomes surprisingly large as the volume fraction of long chains increases. The results are used to test a modified reptation model that includes constraint release and a short-range orientational coupling between chain segments. textcopyright 1991 American Chemical Society.},\n\tAuthor = {Kornfield, J A and Fuller, G.G. and Pearson, D S},\n\tJournal = {Macromolecules},\n\tNumber = {19},\n\tPages = {5429--5441},\n\tTitle = {Third normal stress difference and component relaxation spectra for bidisperse melts under oscillatory shear},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026221264&partnerID=40&md5=63619082916c156a87fb2ccaad5dbdf7},\n\tVolume = {24},\n\tYear = {1991},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026221264&partnerID=40&md5=63619082916c156a87fb2ccaad5dbdf7}}\n\n
\n
\n\n\n
\n A recently developed rheooptical technique that simultaneously measures infrared dichroism and birefringence is used to study bidisperse, linear polymer melts in an oscillatory shear. The method was used to individually measure the contribution of the high and low molecular weight components to the third normal stress difference. This was accomplished by measuring the infrared dichroism due to deuterium-labeled polymers. Binary blends are prepared from nearly monodisperse pairs of hydrogenated and deuteriated polyisoprenes of molecular weights 53 000 and 370 000 with volume fractions of long polymer, ΦL, of 0.10, 0.20, 0.30, 0.50, and 0.75. The long-chain relaxation in the blends shows the appearance of a peak for compositions with ΦL le 0.50 at approximately the same frequency as a similar peak in the pure short polymer. The response of the short-chain relaxation in the blends shows the appearance of a low-frequency shoulder that becomes surprisingly large as the volume fraction of long chains increases. The results are used to test a modified reptation model that includes constraint release and a short-range orientational coupling between chain segments. textcopyright 1991 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Role of director tumbling in the rheology of polymer liquid crystal solutions.\n \n \n \n \n\n\n \n Burghardt, W. R; and Fuller, G. G\n\n\n \n\n\n\n Macromolecules, 24(9): 2546–2555. 1991.\n \n\n\n\n
\n\n\n\n \n \n \"RolePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{burghardt_role_1991,\n\tAbstract = {The role of director tumbling in the dynamic response of polymer liquid crystals is investigated. A flow experiment on an oriented monodomain confirms director tumbling in liquid crystalline poly(benzyl glutamate) solutions. Mechanical and optical rheometric techniques are used to investigate the behavior of textured liquid crystal solutions in slow shear flows. Director tumbling in this system is shown to play a major role in determining the rheological behavior, particularly in that it results in large distortions in the director profile, leading to significant distortional elastic effects. The rheological data are interpreted in terms of general scaling arguments based on the assumption that the director field responds over a length scale associated with the texture rather than the macroscopic flow dimension. Comparison of our results with model calculations using the Leslie-Ericksen continuum model further suggests that the texture length scale is refined in response to increased shear rates, as a mechanism for limiting distortional free energy in the shear flow; rheooptical evidence for this texture refinement is described. A broad range of experimental data may be explained by these arguments, including a widely observed relaxation scaling law.},\n\tAuthor = {Burghardt, Wesley R and Fuller, Gerald G},\n\tJournal = {Macromolecules},\n\tNumber = {9},\n\tPages = {2546--2555},\n\tTitle = {Role of director tumbling in the rheology of polymer liquid crystal solutions},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026140629&partnerID=40&md5=b1a9632b1b06b95fdc2c679d2fb283ea},\n\tVolume = {24},\n\tYear = {1991},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026140629&partnerID=40&md5=b1a9632b1b06b95fdc2c679d2fb283ea}}\n\n
\n
\n\n\n
\n The role of director tumbling in the dynamic response of polymer liquid crystals is investigated. A flow experiment on an oriented monodomain confirms director tumbling in liquid crystalline poly(benzyl glutamate) solutions. Mechanical and optical rheometric techniques are used to investigate the behavior of textured liquid crystal solutions in slow shear flows. Director tumbling in this system is shown to play a major role in determining the rheological behavior, particularly in that it results in large distortions in the director profile, leading to significant distortional elastic effects. The rheological data are interpreted in terms of general scaling arguments based on the assumption that the director field responds over a length scale associated with the texture rather than the macroscopic flow dimension. Comparison of our results with model calculations using the Leslie-Ericksen continuum model further suggests that the texture length scale is refined in response to increased shear rates, as a mechanism for limiting distortional free energy in the shear flow; rheooptical evidence for this texture refinement is described. A broad range of experimental data may be explained by these arguments, including a widely observed relaxation scaling law.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Polymers as Rheology Modifiers.\n \n \n \n \n\n\n \n Fuller, G. G; and Cathey, C. A\n\n\n \n\n\n\n Volume 462 of ACS Symposium SeriesAmerican Chemical Society, Washington, DC, May 1991.\n \n\n\n\n
\n\n\n\n \n \n \"PolymersPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@book{fuller_polymers_1991,\n\tAbstract = {This paper reviews recent progress in understanding of the extensional flow behavior of polymer solutions. Classification of flows with extensional character according to flow strength is discussed along with a summary the current state of molecular modeling in this area. Different experimental approaches to measurement of the extensional viscosity and macromolecular conformation are discussed. Applications of extensional flow measurements to macromolecular characterization, chain scission and fluid dynamics are presented. This paper reviews recent progress in understanding of the extensional flow behavior of polymer solutions. Classification of flows with extensional character according to flow strength is discussed along with a summary the current state of molecular modeling in this area. Different experimental approaches to measurement of the extensional viscosity and macromolecular conformation are discussed. Applications of extensional flow measurements to macromolecular characterization, chain scission and fluid dynamics are presented.},\n\tAddress = {Washington, DC},\n\tAuthor = {Fuller, Gerald G and Cathey, Cheryl A},\n\tEditor = {Schulz, Donald N and Glass, J Edward},\n\tIsbn = {0-8412-2009-3},\n\tMonth = may,\n\tPublisher = {American Chemical Society},\n\tSeries = {{ACS} {Symposium} {Series}},\n\tTitle = {Polymers as {Rheology} {Modifiers}},\n\tUrl = {http://dx.doi.org/10.1021/bk-1991-0462.ch003},\n\tVolume = {462},\n\tYear = {1991},\n\tBdsk-Url-1 = {http://dx.doi.org/10.1021/bk-1991-0462.ch003}}\n\n
\n
\n\n\n
\n This paper reviews recent progress in understanding of the extensional flow behavior of polymer solutions. Classification of flows with extensional character according to flow strength is discussed along with a summary the current state of molecular modeling in this area. Different experimental approaches to measurement of the extensional viscosity and macromolecular conformation are discussed. Applications of extensional flow measurements to macromolecular characterization, chain scission and fluid dynamics are presented. This paper reviews recent progress in understanding of the extensional flow behavior of polymer solutions. Classification of flows with extensional character according to flow strength is discussed along with a summary the current state of molecular modeling in this area. Different experimental approaches to measurement of the extensional viscosity and macromolecular conformation are discussed. Applications of extensional flow measurements to macromolecular characterization, chain scission and fluid dynamics are presented.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Shear wave propagation after stepwise increase of shear rate.\n \n \n \n \n\n\n \n Lee, J S; and Fuller, G.\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 39(1): 1–15. 1991.\n \n\n\n\n
\n\n\n\n \n \n \"ShearPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lee_shear_1991,\n\tAbstract = {Spatially resolved flow birefringence is used to investigate shear wave propagation in polymeric liquids following a step increase in boundary velocities. Under the same flow conditions and with suitable parameters, numerical calculations using the Giesekus constitutive equation are presented. Comparisons indicate good qualitative agreement in wave speeds and wave reflection phenomena for various combinations of initial and final boundary velocities. The results are also compared with recent data by Takahashi and co-workers on the effect of shear-rate changes on entanglement density in concentrated solutions. textcopyright 1991.},\n\tAuthor = {Lee, J S and Fuller, G.G.},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {1},\n\tPages = {1--15},\n\tTitle = {Shear wave propagation after stepwise increase of shear rate},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026116802&partnerID=40&md5=6149bf4de5945391dc718ceac9647919},\n\tVolume = {39},\n\tYear = {1991},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0026116802&partnerID=40&md5=6149bf4de5945391dc718ceac9647919}}\n\n
\n
\n\n\n
\n Spatially resolved flow birefringence is used to investigate shear wave propagation in polymeric liquids following a step increase in boundary velocities. Under the same flow conditions and with suitable parameters, numerical calculations using the Giesekus constitutive equation are presented. Comparisons indicate good qualitative agreement in wave speeds and wave reflection phenomena for various combinations of initial and final boundary velocities. The results are also compared with recent data by Takahashi and co-workers on the effect of shear-rate changes on entanglement density in concentrated solutions. textcopyright 1991.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Structure and dynamics of a polymer solution subject to flow-induced phase separation.\n \n \n \n \n\n\n \n Yanase, H; Moldenaers, P; Mewis, J; Abetz, V; van Egmond, J; and Fuller, G.\n\n\n \n\n\n\n Rheologica Acta, 30(1): 89–97. 1991.\n \n\n\n\n
\n\n\n\n \n \n \"StructurePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{yanase_structure_1991,\n\tAbstract = {Experiments combining mechanical rheometry with polarimetry (birefringence and scattering dichroism) have been conducted on a 6\\% solution of polystyrene (1.86x106 molecular weight) in dioctyl phthalate. Birefringence is used to measure the extent of segmental orientation, whereas the dichroism is sensitive to orientation and deformation of concentration fluctuations associated with the process of flow-induced phase separation. The results indicate that these fluctuations grow predominately along the neutral (or vorticity axis) of a simple shear flow. At higher rates of shear, orientation in the flow direction is favored. The transition in orientation direction is accompanied by time-dependent behavior in the optical properties of the solution during shear and the onset of shear thickening of the viscosity and the first normal stress difference coefficient. textcopyright 1991 Steinkopff-Verlag.},\n\tAuthor = {Yanase, H and Moldenaers, P and Mewis, J and Abetz, V and van Egmond, J and Fuller, G.G.},\n\tJournal = {Rheologica Acta},\n\tNumber = {1},\n\tPages = {89--97},\n\tTitle = {Structure and dynamics of a polymer solution subject to flow-induced phase separation},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0002098908&partnerID=40&md5=80f7b028c0f6715835bc5b52660a4b47},\n\tVolume = {30},\n\tYear = {1991},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0002098908&partnerID=40&md5=80f7b028c0f6715835bc5b52660a4b47}}\n\n
\n
\n\n\n
\n Experiments combining mechanical rheometry with polarimetry (birefringence and scattering dichroism) have been conducted on a 6% solution of polystyrene (1.86x106 molecular weight) in dioctyl phthalate. Birefringence is used to measure the extent of segmental orientation, whereas the dichroism is sensitive to orientation and deformation of concentration fluctuations associated with the process of flow-induced phase separation. The results indicate that these fluctuations grow predominately along the neutral (or vorticity axis) of a simple shear flow. At higher rates of shear, orientation in the flow direction is favored. The transition in orientation direction is accompanied by time-dependent behavior in the optical properties of the solution during shear and the onset of shear thickening of the viscosity and the first normal stress difference coefficient. textcopyright 1991 Steinkopff-Verlag.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1990\n \n \n (11)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Optical anisotropy in colloidal crystals.\n \n \n \n \n\n\n \n Monovoukas, Y; Fuller, G.; and Gast, A P\n\n\n \n\n\n\n The Journal of Chemical Physics, 93(11): 8294–8299. 1990.\n \n\n\n\n
\n\n\n\n \n \n \"OpticalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{monovoukas_optical_1990,\n\tAbstract = {We use the theory of dynamical diffraction to interpret polarization-dependent light attenuation phenomena in thin colloidal crystals. We measure light attenuation anisotropies, resulting from coherent and incoherent scattering from a suspension of monodisperse aqueous polystyrene microspheres, with rotary polarization modulating polarimetry. Measured anisotropies are in excellent agreement with dynamical diffraction predictions. These measurements provide a new means to quantify crystal orientation in thin cells. We relate anisotropy to crystal color observed between crossed polarizers and to crystal structure, volume fraction and orientation. textcopyright 1990 American Institute of Physics.},\n\tAuthor = {Monovoukas, Y and Fuller, G.G. and Gast, A P},\n\tJournal = {The Journal of Chemical Physics},\n\tNumber = {11},\n\tPages = {8294--8299},\n\tTitle = {Optical anisotropy in colloidal crystals},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-36549094387&partnerID=40&md5=b9837fbc8eecb5507b9e2c87b0612cf6},\n\tVolume = {93},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-36549094387&partnerID=40&md5=b9837fbc8eecb5507b9e2c87b0612cf6}}\n\n
\n
\n\n\n
\n We use the theory of dynamical diffraction to interpret polarization-dependent light attenuation phenomena in thin colloidal crystals. We measure light attenuation anisotropies, resulting from coherent and incoherent scattering from a suspension of monodisperse aqueous polystyrene microspheres, with rotary polarization modulating polarimetry. Measured anisotropies are in excellent agreement with dynamical diffraction predictions. These measurements provide a new means to quantify crystal orientation in thin cells. We relate anisotropy to crystal color observed between crossed polarizers and to crystal structure, volume fraction and orientation. textcopyright 1990 American Institute of Physics.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Infrared linear dichroism spectroscopy by a double modulation technique.\n \n \n \n \n\n\n \n Abetz, V; Fuller, G.; and Stadler, R\n\n\n \n\n\n\n Polymer Bulletin, 23(4): 447–454. 1990.\n \n\n\n\n
\n\n\n\n \n \n \"InfraredPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{abetz_infrared_1990,\n\tAbstract = {A linear IR dichroism experiment has been devised using a FTIR instrument in combination with a photoelastic modulator. Based on the Mueller-Stokes calculus, a method to directly obtain the absorbance spectra A vparallel and A vperpendicular from the modulated interferogram is developped, which allows the sensitive evaluation of the dichroic ratio. The photoelastic modulator, in combination with a FTIR instrument, offers high sensitivity, high speed, excellent signal to noise ratio and a broad spectral range from 3300 to 850 cm-1. First experimental results obstained from SBS block copolymer are reported. textcopyright 1990 Springer-Verlag.},\n\tAuthor = {Abetz, V and Fuller, G.G. and Stadler, R},\n\tJournal = {Polymer Bulletin},\n\tNumber = {4},\n\tPages = {447--454},\n\tTitle = {Infrared linear dichroism spectroscopy by a double modulation technique},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0025414282&partnerID=40&md5=6ec2ae73fbcc2c60422576ad752a3890},\n\tVolume = {23},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0025414282&partnerID=40&md5=6ec2ae73fbcc2c60422576ad752a3890}}\n\n
\n
\n\n\n
\n A linear IR dichroism experiment has been devised using a FTIR instrument in combination with a photoelastic modulator. Based on the Mueller-Stokes calculus, a method to directly obtain the absorbance spectra A vparallel and A vperpendicular from the modulated interferogram is developped, which allows the sensitive evaluation of the dichroic ratio. The photoelastic modulator, in combination with a FTIR instrument, offers high sensitivity, high speed, excellent signal to noise ratio and a broad spectral range from 3300 to 850 cm-1. First experimental results obstained from SBS block copolymer are reported. textcopyright 1990 Springer-Verlag.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Optical rheometry.\n \n \n \n \n\n\n \n Fuller, G.\n\n\n \n\n\n\n Annual Review of Fluid Mechanics, 22(1): 387–417. 1990.\n \n\n\n\n
\n\n\n\n \n \n \"OpticalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_optical_1990,\n\tAuthor = {Fuller, G.G.},\n\tJournal = {Annual Review of Fluid Mechanics},\n\tNumber = {1},\n\tPages = {387--417},\n\tTitle = {Optical rheometry},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001323459&partnerID=40&md5=27ead415d4e741ce1d84d523198b19c3},\n\tVolume = {22},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001323459&partnerID=40&md5=27ead415d4e741ce1d84d523198b19c3}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The optical and mechanical response of flexible polymer solutions to extensional flow.\n \n \n \n \n\n\n \n Cathey, C A; and Fuller, G.\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 34(1): 63–88. 1990.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{cathey_optical_1990,\n\tAbstract = {The optical and mechanical responses of dilute and semidulute solutions of flexible polymer molecules to extensional flow are measured. The optical property of interest, the birefringence, is sensitive to the local orientation of the polymer coil, whereas the mechanical property investigated, the effective extensional viscosity, is sensitive to the overall deformed length of the molecule. These two properties are simultaneously measured as a function of the rate of strain by using an opposing jets apparatus. Solutions ranging from 50 to 300 ppm by weight polystyrene dissolved in both tri-cresyl phosphate, a good solvent at 22textdegreeC, and di-octyl phthalate, a theta solvent at 22textdegreeC, were studied. The results show that the response of the flexible polymer is dependent on the solvent quality. Both the local orientation and the deformed length of the molecule increase with increasing strain rate at low rates of strain for both solvents. However, in the theta solvent at high rates of strain the birefringence saturates, while the effective extensional viscosity drops with increasing strain rate. This indicates a decrease in the overall deformed length of the molecule at high strain rates due to the decrease in the residence time of the coil in the flow field. In the good solvent, molecular entanglements begin to affect the extensional viscosity at high strain rates if the concentration and molecular weight are sufficiently large. The response of flexible polymers to extensional flow is qualitatively compared to numerical simulations based on bead-spring and bead-rod models. Although these models are able to capture the saturation of birefringence under conditions where the extensional viscosity is still changing, they do not predict the observed maxima in extensional viscosity as a function of strain rate. textcopyright 1990.},\n\tAuthor = {Cathey, C A and Fuller, G.G.},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {1},\n\tPages = {63--88},\n\tTitle = {The optical and mechanical response of flexible polymer solutions to extensional flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0025225075&partnerID=40&md5=ed56f1e3061b7433687bcc091d897795},\n\tVolume = {34},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0025225075&partnerID=40&md5=ed56f1e3061b7433687bcc091d897795}}\n\n
\n
\n\n\n
\n The optical and mechanical responses of dilute and semidulute solutions of flexible polymer molecules to extensional flow are measured. The optical property of interest, the birefringence, is sensitive to the local orientation of the polymer coil, whereas the mechanical property investigated, the effective extensional viscosity, is sensitive to the overall deformed length of the molecule. These two properties are simultaneously measured as a function of the rate of strain by using an opposing jets apparatus. Solutions ranging from 50 to 300 ppm by weight polystyrene dissolved in both tri-cresyl phosphate, a good solvent at 22textdegreeC, and di-octyl phthalate, a theta solvent at 22textdegreeC, were studied. The results show that the response of the flexible polymer is dependent on the solvent quality. Both the local orientation and the deformed length of the molecule increase with increasing strain rate at low rates of strain for both solvents. However, in the theta solvent at high rates of strain the birefringence saturates, while the effective extensional viscosity drops with increasing strain rate. This indicates a decrease in the overall deformed length of the molecule at high strain rates due to the decrease in the residence time of the coil in the flow field. In the good solvent, molecular entanglements begin to affect the extensional viscosity at high strain rates if the concentration and molecular weight are sufficiently large. The response of flexible polymers to extensional flow is qualitatively compared to numerical simulations based on bead-spring and bead-rod models. Although these models are able to capture the saturation of birefringence under conditions where the extensional viscosity is still changing, they do not predict the observed maxima in extensional viscosity as a function of strain rate. textcopyright 1990.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Relaxation dynamics of selected polymer chain segments and comparison with theoretical models.\n \n \n \n \n\n\n \n Ylitalo, C M; Fuller, G.; Abetz, V; Stadler, R; and Pearson, D S\n\n\n \n\n\n\n Rheologica Acta, 29(6): 543–555. 1990.\n \n\n\n\n
\n\n\n\n \n \n \"RelaxationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{ylitalo_relaxation_1990,\n\tAbstract = {Simultaneous measurement of infrared dichroism and birefringence is used to study selected polymer segment dynamics in isotopically labeled block copolymers. Two different polymers were studied: polybutadiene and poly (ethylene propylene). The first type consisted of a triblock with a short middle block labeled and a diblock with a short end block labeled, while the second type consisted of a triblock with three equal blocks and the end blocks labeled. Results of step strain experiments at -10textdegreeC for polybutadiene and at room temperature for poly(ethylene propylene) indicated that segments located at chain ends relax faster than segments located at chain centers. These experimental data were compared to the predictions of two molecular models: the bead-spring model of Rouse and the tube model of Doi and Edwards, and it was found that both models correctly predict the qualitative features of segmental relaxation. However, the tube-model predictions were closer to the experimental results. In addition, when the effects of orientational coupling interactions between segments in the melt were incorporated into this model, its predictions quantitatively agreed with the experimental results. The orientational coupling coefficient for poly(ethylene propylene) was 0.45 as measured from previous work, and for polybutadiene it was found to be 0.4. textcopyright 1990 Steinkopff.},\n\tAuthor = {Ylitalo, C M and Fuller, G.G. and Abetz, V and Stadler, R and Pearson, D S},\n\tJournal = {Rheologica Acta},\n\tNumber = {6},\n\tPages = {543--555},\n\tTitle = {Relaxation dynamics of selected polymer chain segments and comparison with theoretical models},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001664108&partnerID=40&md5=dcc9ec4d5cc3393b61a758914d1853e4},\n\tVolume = {29},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0001664108&partnerID=40&md5=dcc9ec4d5cc3393b61a758914d1853e4}}\n\n
\n
\n\n\n
\n Simultaneous measurement of infrared dichroism and birefringence is used to study selected polymer segment dynamics in isotopically labeled block copolymers. Two different polymers were studied: polybutadiene and poly (ethylene propylene). The first type consisted of a triblock with a short middle block labeled and a diblock with a short end block labeled, while the second type consisted of a triblock with three equal blocks and the end blocks labeled. Results of step strain experiments at -10textdegreeC for polybutadiene and at room temperature for poly(ethylene propylene) indicated that segments located at chain ends relax faster than segments located at chain centers. These experimental data were compared to the predictions of two molecular models: the bead-spring model of Rouse and the tube model of Doi and Edwards, and it was found that both models correctly predict the qualitative features of segmental relaxation. However, the tube-model predictions were closer to the experimental results. In addition, when the effects of orientational coupling interactions between segments in the melt were incorporated into this model, its predictions quantitatively agreed with the experimental results. The orientational coupling coefficient for poly(ethylene propylene) was 0.45 as measured from previous work, and for polybutadiene it was found to be 0.4. textcopyright 1990 Steinkopff.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Two-color rotary modulated flow birefringence.\n \n \n \n \n\n\n \n Abetz, V; and Fuller, G.\n\n\n \n\n\n\n Rheologica Acta, 29(1): 11–15. January 1990.\n \n\n\n\n
\n\n\n\n \n \n \"Two-colorPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{abetz_two-color_1990,\n\tAbstract = {Abstract A method is described that solves the problem of determining the correct birefringence and orientation angle of samples having multiple orders of retardation. The approach simultaneously uses two wavelengths of light combined with modulation of the ...},\n\tAuthor = {Abetz, V and Fuller, G.G.},\n\tDoi = {10.1007/BF01331796},\n\tJournal = {Rheologica Acta},\n\tLanguage = {English},\n\tMonth = jan,\n\tNumber = {1},\n\tPages = {11--15},\n\tTitle = {Two-color rotary modulated flow birefringence},\n\tUrl = {http://link.springer.com/10.1007/BF01331796},\n\tVolume = {29},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://link.springer.com/10.1007/BF01331796},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1007/BF01331796}}\n\n
\n
\n\n\n
\n Abstract A method is described that solves the problem of determining the correct birefringence and orientation angle of samples having multiple orders of retardation. The approach simultaneously uses two wavelengths of light combined with modulation of the ...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Stress optical measurement of the third normal stress difference in polymer melts under oscillatory shear.\n \n \n \n \n\n\n \n Kornfield, J A; Fuller, G.; and Pearson, D S\n\n\n \n\n\n\n Rheologica Acta, 29(2): 105–116. 1990.\n \n\n\n\n
\n\n\n\n \n \n \"StressPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kornfield_stress_1990,\n\tAbstract = {Results are reported for the dynamic moduli, Gtextasciiacute and Gtextacutedbl, measured mechanically, and the dynamic third normal stress difference, measured optically, of a series bidisperse linear polymer melts under oscillatory shear. Nearly monodisperse hydrogenated polyisoprenes of molecular weights 53000 and 370000 were used to prepare blends with a volume fraction of long polymer, ΦL, of 0.10, 0.20, 0.30, 0.50, and 0.75. The results demonstrate the applicability of birefringence measurements to solve the longstanding problem of measuring the third normal stress difference in oscillatory flow. The relationship between the third normal stress difference and the shear stress observed for these entangled polymer melts is in agreement with a widely predicted constitutive relationship: the relationship between the first normal stress difference and the shear stress is that of a simple fluid, and the second normal stress difference is proportional to the first. These results demonstrate the potential use of 1,3-birefringence to measure the third normal stress difference in oscillatory flow. Further, the general constitutive equation supported by the present results may be used to determine the dynamic moduli from the measured third normal stress difference in small amplitude oscillatory shear. Directions for future research, including the use of birefringence measurements to determine N2/N1 in oscillatory shear, are described. textcopyright 1990 Steinkopff.},\n\tAuthor = {Kornfield, J A and Fuller, G.G. and Pearson, D S},\n\tJournal = {Rheologica Acta},\n\tNumber = {2},\n\tPages = {105--116},\n\tTitle = {Stress optical measurement of the third normal stress difference in polymer melts under oscillatory shear},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000640507&partnerID=40&md5=dc164049b641c15450418f1d402b043f},\n\tVolume = {29},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000640507&partnerID=40&md5=dc164049b641c15450418f1d402b043f}}\n\n
\n
\n\n\n
\n Results are reported for the dynamic moduli, Gtextasciiacute and Gtextacutedbl, measured mechanically, and the dynamic third normal stress difference, measured optically, of a series bidisperse linear polymer melts under oscillatory shear. Nearly monodisperse hydrogenated polyisoprenes of molecular weights 53000 and 370000 were used to prepare blends with a volume fraction of long polymer, ΦL, of 0.10, 0.20, 0.30, 0.50, and 0.75. The results demonstrate the applicability of birefringence measurements to solve the longstanding problem of measuring the third normal stress difference in oscillatory flow. The relationship between the third normal stress difference and the shear stress observed for these entangled polymer melts is in agreement with a widely predicted constitutive relationship: the relationship between the first normal stress difference and the shear stress is that of a simple fluid, and the second normal stress difference is proportional to the first. These results demonstrate the potential use of 1,3-birefringence to measure the third normal stress difference in oscillatory flow. Further, the general constitutive equation supported by the present results may be used to determine the dynamic moduli from the measured third normal stress difference in small amplitude oscillatory shear. Directions for future research, including the use of birefringence measurements to determine N2/N1 in oscillatory shear, are described. textcopyright 1990 Steinkopff.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Dynamics of colloidal particles in sheared, non-Newtonian fluids.\n \n \n \n \n\n\n \n Johnson, S J; Salem, A J; and Fuller, G.\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 34(1): 89–121. 1990.\n \n\n\n\n
\n\n\n\n \n \n \"DynamicsPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{johnson_dynamics_1990,\n\tAbstract = {The dynamics of colloidal, axisymmetric particles suspended in non-Newtonian, polymeric liquids are examined by using optical techniques that are capable of following their average orientation. Polarimetry measurements of birefringence and dichroism are used on dispersions of colloidal iron oxide spheroids and small-angle light scattering is used on hardened, prolate human red blood cells. A variety of suspending liquids are used including Boger fluids designed to have some of the attributes of a second order fluid under steady state flow conditions, and concentrated polystyrene solutions with rheological properties similar to more commonly encountered polymer liquids. As a result of the non-Newtonian properties of the polymeric suspending fluids, and in particular the presence of normal stresses, the particles were observed to drift out of the orbits associated with their motion in Newtonian fluids. As predicted by Cohen and coworkers, in flows where the elastic forces on the particles dominate Brownian forces, the particles tend to drift away from orientations favoring the flow direction and towards the vorticity axis. When Brownian forces are large compared with the elastic forces, alignments in the flow direction are more probable. This trend, however, was only observed when the Boger fluids were used as suspending fluids. Although the polystyrene solutions possessed shear viscosities and first normal stress difference coefficients that were comparable in magnitude to those of the Boger fluids, only orientations along the flow direction were found. textcopyright 1990.},\n\tAuthor = {Johnson, S J and Salem, A J and Fuller, G.G.},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {1},\n\tPages = {89--121},\n\tTitle = {Dynamics of colloidal particles in sheared, non-{Newtonian} fluids},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0025112097&partnerID=40&md5=a69f36021617fc34bab8339893a1b187},\n\tVolume = {34},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0025112097&partnerID=40&md5=a69f36021617fc34bab8339893a1b187}}\n\n
\n
\n\n\n
\n The dynamics of colloidal, axisymmetric particles suspended in non-Newtonian, polymeric liquids are examined by using optical techniques that are capable of following their average orientation. Polarimetry measurements of birefringence and dichroism are used on dispersions of colloidal iron oxide spheroids and small-angle light scattering is used on hardened, prolate human red blood cells. A variety of suspending liquids are used including Boger fluids designed to have some of the attributes of a second order fluid under steady state flow conditions, and concentrated polystyrene solutions with rheological properties similar to more commonly encountered polymer liquids. As a result of the non-Newtonian properties of the polymeric suspending fluids, and in particular the presence of normal stresses, the particles were observed to drift out of the orbits associated with their motion in Newtonian fluids. As predicted by Cohen and coworkers, in flows where the elastic forces on the particles dominate Brownian forces, the particles tend to drift away from orientations favoring the flow direction and towards the vorticity axis. When Brownian forces are large compared with the elastic forces, alignments in the flow direction are more probable. This trend, however, was only observed when the Boger fluids were used as suspending fluids. Although the polystyrene solutions possessed shear viscosities and first normal stress difference coefficients that were comparable in magnitude to those of the Boger fluids, only orientations along the flow direction were found. textcopyright 1990.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n On-line tribology measurements on lubricated rigid disks.\n \n \n \n \n\n\n \n Nunnelley, L L; Burleson, M A; and Fuller, G.\n\n\n \n\n\n\n 1990 International Magnetics Conference - INTERMAG, 26(5): 2679–2681. 1990.\n \n\n\n\n
\n\n\n\n \n \n \"On-linePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{nunnelley_-line_1990,\n\tAbstract = {A high-speed ellipsometer has been devised for online wear measurements of recording media. The instrument can either determine spatially resolved film-thickness profiles across a wear track or make high-speed measurements of the angular dependence of wear damage as the test specimen rotates. In its present configuration, data-acquisition rates of 300 Hz are possible. Experiments on thin film demonstrate the ability of the technique to monitor thin-film profiles and to reveal the progression of wear from the ablation of the lubricant layer to the removal of the magnetic film.},\n\tAuthor = {Nunnelley, L L and Burleson, M A and Fuller, G.G.},\n\tJournal = {1990 International Magnetics Conference - INTERMAG},\n\tNumber = {5},\n\tPages = {2679--2681},\n\tTitle = {On-line tribology measurements on lubricated rigid disks},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0025491030&partnerID=40&md5=b6551d7abfec62a77e4c41157ef22c8c},\n\tVolume = {26},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0025491030&partnerID=40&md5=b6551d7abfec62a77e4c41157ef22c8c}}\n\n
\n
\n\n\n
\n A high-speed ellipsometer has been devised for online wear measurements of recording media. The instrument can either determine spatially resolved film-thickness profiles across a wear track or make high-speed measurements of the angular dependence of wear damage as the test specimen rotates. In its present configuration, data-acquisition rates of 300 Hz are possible. Experiments on thin film demonstrate the ability of the technique to monitor thin-film profiles and to reveal the progression of wear from the ablation of the lubricant layer to the removal of the magnetic film.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Mechanical and optical responses of the M1 fluid subject to stagnation point flow.\n \n \n \n \n\n\n \n Schweizer, T; Mikkelsen, K; Cathey, C; and Fuller, G\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 35(2-3): 277–286. 1990.\n \n\n\n\n
\n\n\n\n \n \n \"MechanicalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{schweizer_mechanical_1990,\n\tAbstract = {In this paper, the flow properties of the M1 fluid are studied in a stagnation point flow generated using opposed jets. Mechanical measurements of an effective extensional viscosity and optical birefringence measurements are presented for rates of strain ranging from 0.02 to 50 s -1 and at temperatures from 20 to 40 textdegree C. A transition from a Newtonian plateau to strain thickening behavior was identified and time-temperature superposition of the effective viscosity data was found to apply. Comparison of the optical and mechanical results indicates that a stress-optical rule does not apply in extension. At rates of strain above 50 s -1, the flow became time-dependent and unstable. This was demonstrated using both birefringence and flow visualization techniques. textcopyright 1990.},\n\tAuthor = {Schweizer, T and Mikkelsen, K and Cathey, C and Fuller, G},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {2-3},\n\tPages = {277--286},\n\tTitle = {Mechanical and optical responses of the {M}1 fluid subject to stagnation point flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0025456793&partnerID=40&md5=1ce80d28ba806d088ff0ee01170b0016},\n\tVolume = {35},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0025456793&partnerID=40&md5=1ce80d28ba806d088ff0ee01170b0016}}\n\n
\n
\n\n\n
\n In this paper, the flow properties of the M1 fluid are studied in a stagnation point flow generated using opposed jets. Mechanical measurements of an effective extensional viscosity and optical birefringence measurements are presented for rates of strain ranging from 0.02 to 50 s -1 and at temperatures from 20 to 40 textdegree C. A transition from a Newtonian plateau to strain thickening behavior was identified and time-temperature superposition of the effective viscosity data was found to apply. Comparison of the optical and mechanical results indicates that a stress-optical rule does not apply in extension. At rates of strain above 50 s -1, the flow became time-dependent and unstable. This was demonstrated using both birefringence and flow visualization techniques. textcopyright 1990.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Transient shear flow of nematic liquid crystals: Manifestations of director tumbling.\n \n \n \n \n\n\n \n Burghardt, W. R\n\n\n \n\n\n\n Journal of Rheology, 34(6): 959. August 1990.\n \n\n\n\n
\n\n\n\n \n \n \"TransientPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{burghardt_transient_1990,\n\tAbstract = {Calculations of a variety of transient flow phenomena based on the Leslie--Ericksen model for nematic liquid crystals are presented. Emphasis is placed on the behavior of nematics subject to director tumbling. A wide range of complicated transient phenomena such as oscillatory responses are predicted when the Ericksen number is sufficiently high to cause significant rotation of the director at steady state. Particular attention is given to relaxation processes such as constrained elastic recoil and stress relaxation that arise from the interaction of the relaxing director profile with the viscous response of the nematic. These relaxation processes are dramatically different in tumbling and flow aligning systems, due to the saturation of the director profile at high Ericksen number for the flow aligning case. A qualitative physical model of a textured polymer liquid crystal is presented in which the dynamics on a local scale governed by the distance between defects are assumed to be similar to the macroscopic dynamics presented in these calculations. By assuming (i) director tumbling and (ii) the existence of a limiting Ericksen number in polymer systems, a wide range of published observations in lyotropic liquid crystalline polymer solutions may be qualitatively reproduced.},\n\tAuthor = {Burghardt, Wesley R},\n\tDoi = {10.1122/1.550151},\n\tJournal = {Journal of Rheology},\n\tMonth = aug,\n\tNumber = {6},\n\tPages = {959},\n\tTitle = {Transient shear flow of nematic liquid crystals: {Manifestations} of director tumbling},\n\tUrl = {http://scitation.aip.org/content/sor/journal/jor2/34/6/10.1122/1.550151},\n\tVolume = {34},\n\tYear = {1990},\n\tBdsk-Url-1 = {http://scitation.aip.org/content/sor/journal/jor2/34/6/10.1122/1.550151},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1122/1.550151}}\n\n
\n
\n\n\n
\n Calculations of a variety of transient flow phenomena based on the Leslie–Ericksen model for nematic liquid crystals are presented. Emphasis is placed on the behavior of nematics subject to director tumbling. A wide range of complicated transient phenomena such as oscillatory responses are predicted when the Ericksen number is sufficiently high to cause significant rotation of the director at steady state. Particular attention is given to relaxation processes such as constrained elastic recoil and stress relaxation that arise from the interaction of the relaxing director profile with the viscous response of the nematic. These relaxation processes are dramatically different in tumbling and flow aligning systems, due to the saturation of the director profile at high Ericksen number for the flow aligning case. A qualitative physical model of a textured polymer liquid crystal is presented in which the dynamics on a local scale governed by the distance between defects are assumed to be similar to the macroscopic dynamics presented in these calculations. By assuming (i) director tumbling and (ii) the existence of a limiting Ericksen number in polymer systems, a wide range of published observations in lyotropic liquid crystalline polymer solutions may be qualitatively reproduced.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1989\n \n \n (6)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Mechanical and optical rheometry of polymer liquid-crystal domain structure.\n \n \n \n \n\n\n \n Moldenaers, P; Fuller, G; and Mewis, J\n\n\n \n\n\n\n Macromolecules, 22(2): 960–965. 1989.\n \n\n\n\n
\n\n\n\n \n \n \"MechanicalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{moldenaers_mechanical_1989,\n\tAbstract = {Complementary mechanical and optical rheometric techniques are used to examine the domain structure of two lyotropic polymer liquid-crystalline samples. The mechanical methods involved both shear and first normal stress measurements, and the optical experiments employed conservative linear dichroism. The experimental protocol used transient simple shear flows where the shear rate was either suddenly reversed in sign or stepped up to a higher value. In either case, pronounced oscillations of the same period were observed in both the mechanical and optical measurements. The periods, if measured in strain units, were observed to be independent of strain rate and suggested a suspension-like behavior. Since the conservative linear dichroism can be identified with the polarization-dependent scattering of light by the defect structure in the samples, these results offer strong evidence that this structure dominates the rheological response. textcopyright 1989 American Chemical Society.},\n\tAuthor = {Moldenaers, P and Fuller, G and Mewis, J},\n\tJournal = {Macromolecules},\n\tNumber = {2},\n\tPages = {960--965},\n\tTitle = {Mechanical and optical rheometry of polymer liquid-crystal domain structure},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024607559&partnerID=40&md5=8e323e39cf43401996b2f7a16aacc3ab},\n\tVolume = {22},\n\tYear = {1989},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024607559&partnerID=40&md5=8e323e39cf43401996b2f7a16aacc3ab}}\n\n
\n
\n\n\n
\n Complementary mechanical and optical rheometric techniques are used to examine the domain structure of two lyotropic polymer liquid-crystalline samples. The mechanical methods involved both shear and first normal stress measurements, and the optical experiments employed conservative linear dichroism. The experimental protocol used transient simple shear flows where the shear rate was either suddenly reversed in sign or stepped up to a higher value. In either case, pronounced oscillations of the same period were observed in both the mechanical and optical measurements. The periods, if measured in strain units, were observed to be independent of strain rate and suggested a suspension-like behavior. Since the conservative linear dichroism can be identified with the polarization-dependent scattering of light by the defect structure in the samples, these results offer strong evidence that this structure dominates the rheological response. textcopyright 1989 American Chemical Society.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n End effects in flow birefringence measurements.\n \n \n \n \n\n\n \n Burghardt, W R; and Fuller, G.\n\n\n \n\n\n\n Journal of Rheology, 33(5): 771–779. 1989.\n \n\n\n\n
\n\n\n\n \n \n \"EndPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{burghardt_end_1989,\n\tAbstract = {Flow birefringence techniques are becoming increasingly valuable tools in studying the response of polymeric fluids to deformation. The authors address the issue by modeling the optical properties of a solution undergoing confined shear flow in a planar Couette geometry. They believe that the approach outlined in this note is of value since the optical properties of the sample are the observables in a flow birefringence experiment. Similarly, while other investigators have calculated velocity profiles in such confined shearing geometries, the further modeling of the optical properties allows a more direct assessment of how end effects influence the flow birefringence measurements.},\n\tAuthor = {Burghardt, W R and Fuller, G.G.},\n\tJournal = {Journal of Rheology},\n\tNumber = {5},\n\tPages = {771--779},\n\tTitle = {End effects in flow birefringence measurements},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024702305&partnerID=40&md5=0ea4c2589df15cdf8d065ccd10945a08},\n\tVolume = {33},\n\tYear = {1989},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024702305&partnerID=40&md5=0ea4c2589df15cdf8d065ccd10945a08}}\n\n
\n
\n\n\n
\n Flow birefringence techniques are becoming increasingly valuable tools in studying the response of polymeric fluids to deformation. The authors address the issue by modeling the optical properties of a solution undergoing confined shear flow in a planar Couette geometry. They believe that the approach outlined in this note is of value since the optical properties of the sample are the observables in a flow birefringence experiment. Similarly, while other investigators have calculated velocity profiles in such confined shearing geometries, the further modeling of the optical properties allows a more direct assessment of how end effects influence the flow birefringence measurements.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rheologically interesting polysaccharides from yeasts.\n \n \n \n \n\n\n \n Petersen, G R; Nelson, G A; Cathey, C A; and Fuller, G.\n\n\n \n\n\n\n Applied Biochemistry and Biotechnology, 20-21(1): 845–867. 1989.\n \n\n\n\n
\n\n\n\n \n \n \"RheologicallyPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{petersen_rheologically_1989,\n\tAbstract = {We have examined the relationships between primary, secondary, and tertiary structures of polysaccharides exhibiting the rheological property of friction (drag) reduction in turbulent flows. We found an example of an exopolysaccharide from the yeast Cryptococcus laurentii that possessed high molecular weight but exhibited lower than expected drag reducing activity. Earlier correlations by Hoyt (8,10) showing that β1 textrightarrow 3, βtextrightarrow4, and αl textrightarrow 3 linkages in polysaccharides favored drag reduction were expanded to include correlations to secondary structure. The effect of sidechains in a series of gellan gums was shown to be related to sidechain length and position. Disruption of secondary structure in drag reducing polysaccharides reduced drag reducing activity for some but not all exopolysaccharides. The polymer from C. laurentii was shown to be more stable than xanthan gum and other exopolysaccharides under the most vigorous of denaturing conditions. We also showed a direct relationship between extensional viscosity measurements and the drag reducing coefficient for four exopolysaccharides. textcopyright 1989 Humana Press Inc.},\n\tAuthor = {Petersen, G R and Nelson, G A and Cathey, C A and Fuller, G.G.},\n\tJournal = {Applied Biochemistry and Biotechnology},\n\tNumber = {1},\n\tPages = {845--867},\n\tTitle = {Rheologically interesting polysaccharides from yeasts},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024959215&partnerID=40&md5=db0f2bd891327772cd262f317cf0ce7d},\n\tVolume = {20-21},\n\tYear = {1989},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024959215&partnerID=40&md5=db0f2bd891327772cd262f317cf0ce7d}}\n\n
\n
\n\n\n
\n We have examined the relationships between primary, secondary, and tertiary structures of polysaccharides exhibiting the rheological property of friction (drag) reduction in turbulent flows. We found an example of an exopolysaccharide from the yeast Cryptococcus laurentii that possessed high molecular weight but exhibited lower than expected drag reducing activity. Earlier correlations by Hoyt (8,10) showing that β1 textrightarrow 3, βtextrightarrow4, and αl textrightarrow 3 linkages in polysaccharides favored drag reduction were expanded to include correlations to secondary structure. The effect of sidechains in a series of gellan gums was shown to be related to sidechain length and position. Disruption of secondary structure in drag reducing polysaccharides reduced drag reducing activity for some but not all exopolysaccharides. The polymer from C. laurentii was shown to be more stable than xanthan gum and other exopolysaccharides under the most vigorous of denaturing conditions. We also showed a direct relationship between extensional viscosity measurements and the drag reducing coefficient for four exopolysaccharides. textcopyright 1989 Humana Press Inc.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Optical rheometry using a rotary polarization modulator.\n \n \n \n \n\n\n \n Fuller, G. G; and Mikkelsen, K. J\n\n\n \n\n\n\n Journal of Rheology, 33(5): 761–769. 1989.\n \n\n\n\n
\n\n\n\n \n \n \"OpticalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_optical_1989,\n\tAbstract = {The purpose of this study is to present a new design f an optical rheometer which incorporates a rotary polarization modulator. It is demonstrated that this type of modulation produces information comparable to that obtainable using photoelastic modulation and yet requires simpler calibration and demodulation procedures. The rotary modulator is also far less expensive and a good deal more robust than alternative devices. On the other hand, it is capable of achieving the high modulation frequencies which are necessary for fast rheological tests.},\n\tAuthor = {Fuller, Gerald G and Mikkelsen, Kirk J},\n\tJournal = {Journal of Rheology},\n\tNumber = {5},\n\tPages = {761--769},\n\tTitle = {Optical rheometry using a rotary polarization modulator},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024702557&partnerID=40&md5=2ae77b1eb3a3fdd52b01e085cc96213b},\n\tVolume = {33},\n\tYear = {1989},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024702557&partnerID=40&md5=2ae77b1eb3a3fdd52b01e085cc96213b}}\n\n
\n
\n\n\n
\n The purpose of this study is to present a new design f an optical rheometer which incorporates a rotary polarization modulator. It is demonstrated that this type of modulation produces information comparable to that obtainable using photoelastic modulation and yet requires simpler calibration and demodulation procedures. The rotary modulator is also far less expensive and a good deal more robust than alternative devices. On the other hand, it is capable of achieving the high modulation frequencies which are necessary for fast rheological tests.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Effect of nematic interaction in the orientational relaxation of polymer melts.\n \n \n \n \n\n\n \n Doi, M.; Pearson, D.; Kornfield, J.; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 22(3): 1488–1490. 1989.\n \n\n\n\n
\n\n\n\n \n \n \"EffectPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{doi_effect_1989,\n\tAbstract = {The authors model the entanglements with slip links and assume that the polymer is a freely jointed chain with constant bond length, b. It is shown that the modification does not affect the previous reptation theory in any fundamental way. Also it is consistent with the stress optical law. The theory enables us to study the reptation dynamics, especially the effect of constraint release by spectroscopic techniques.},\n\tAuthor = {Doi, Masao and Pearson, Dale and Kornfield, Julie and Fuller, Gerry},\n\tJournal = {Macromolecules},\n\tNumber = {3},\n\tPages = {1488--1490},\n\tTitle = {Effect of nematic interaction in the orientational relaxation of polymer melts},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024621578&partnerID=40&md5=620c45b4ebed1c4757ee005646928fef},\n\tVolume = {22},\n\tYear = {1989},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024621578&partnerID=40&md5=620c45b4ebed1c4757ee005646928fef}}\n\n
\n
\n\n\n
\n The authors model the entanglements with slip links and assume that the polymer is a freely jointed chain with constant bond length, b. It is shown that the modification does not affect the previous reptation theory in any fundamental way. Also it is consistent with the stress optical law. The theory enables us to study the reptation dynamics, especially the effect of constraint release by spectroscopic techniques.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Infrared dichroism measurements of molecular relaxation in binary blend melt rheology.\n \n \n \n \n\n\n \n Kornfield, J A; Fuller, G.; and Pearson, D S\n\n\n \n\n\n\n Macromolecules, 22(3): 1334–1345. 1989.\n \n\n\n\n
\n\n\n\n \n \n \"InfraredPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kornfield_infrared_1989,\n\tAbstract = {A new rheooptical technique to measure dynamic infrared dichroism was used with deuterium labeling to observe the relaxation of each molecular weight component in a bidisperse entangled polymer melt. The bulk relaxation is measured simultaneously by using birefringence. The relaxation dynamics of the bulk and of each component depend upon the two polymer relaxation times and the blend ratio. Nearly monodisperse pairs of hydrogenated and deuteriated polyisoprenes of molecular weights 53 000 and 370 000, both several times the entanglement molecular weight, were used. Results for the linear viscoelastic step shear strain relaxation of each component and of the bulk for blends containing 10, 20, 30, 50, and 75 vol \\% long chains are presented. The component dynamics are qualitatively predicted by current molecular theories of polymer melt rheology; however, some important differences are observed. In particular, it was found that polydispersity can strongly retard the orientational relaxation of the low molecular weight component and decrease the longest relaxation time of the high molecular weight component. textcopyright 1989 American Chemical Society.},\n\tAuthor = {Kornfield, J A and Fuller, G.G. and Pearson, D S},\n\tJournal = {Macromolecules},\n\tNumber = {3},\n\tPages = {1334--1345},\n\tTitle = {Infrared dichroism measurements of molecular relaxation in binary blend melt rheology},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024622714&partnerID=40&md5=da890022710975364303480fa3911a4a},\n\tVolume = {22},\n\tYear = {1989},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024622714&partnerID=40&md5=da890022710975364303480fa3911a4a}}\n\n
\n
\n\n\n
\n A new rheooptical technique to measure dynamic infrared dichroism was used with deuterium labeling to observe the relaxation of each molecular weight component in a bidisperse entangled polymer melt. The bulk relaxation is measured simultaneously by using birefringence. The relaxation dynamics of the bulk and of each component depend upon the two polymer relaxation times and the blend ratio. Nearly monodisperse pairs of hydrogenated and deuteriated polyisoprenes of molecular weights 53 000 and 370 000, both several times the entanglement molecular weight, were used. Results for the linear viscoelastic step shear strain relaxation of each component and of the bulk for blends containing 10, 20, 30, 50, and 75 vol % long chains are presented. The component dynamics are qualitatively predicted by current molecular theories of polymer melt rheology; however, some important differences are observed. In particular, it was found that polydispersity can strongly retard the orientational relaxation of the low molecular weight component and decrease the longest relaxation time of the high molecular weight component. textcopyright 1989 American Chemical Society.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1988\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Optical measurements of particle orientation in magnetic media.\n \n \n \n \n\n\n \n Nunnelley, L; and Fuller, G.\n\n\n \n\n\n\n Journal of Applied Physics, 63(5): 1687–1690. 1988.\n \n\n\n\n
\n\n\n\n \n \n \"OpticalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{nunnelley_optical_1988,\n\tAbstract = {Not Available},\n\tAuthor = {Nunnelley, L and Fuller, G.G.},\n\tDoi = {10.1063/1.339902},\n\tJournal = {Journal of Applied Physics},\n\tNumber = {5},\n\tPages = {1687--1690},\n\tTitle = {Optical measurements of particle orientation in magnetic media},\n\tUrl = {http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=1988JAP....63.1687N&link_type=EJOURNAL},\n\tVolume = {63},\n\tYear = {1988},\n\tBdsk-Url-1 = {http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=1988JAP....63.1687N&link_type=EJOURNAL},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1063/1.339902}}\n\n
\n
\n\n\n
\n Not Available\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The optical anisotropy of sheared hematite suspensions.\n \n \n \n \n\n\n \n Johnson, S J; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 124(2): 441–451. 1988.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{johnson_optical_1988,\n\tAbstract = {Simultaneous linear dichroism and linear birefringence measurements were used to study sheared suspensions of colloidal hematite (α-Fe2O3) particles. Particles of three different axis ratios were synthesized. The particles were characterized using electron microscopy, and the three samples had axis ratios of 2.6 textpm 0.14, 3.61 textpm 0.24, and 5.82 textpm 0.58 with mean lengths of 0.315, 0.360, and 0.485 μm, repsectively. The particles were suspended in two Newtonian fluids, glycerin (viscosity = 17 P) and low molecular weight polybutene (viscosity = 220 P), and each suspension was subjected to transient shear flow in a Couette flow cell. The role of polydispersity in particle size on the relative orientations of the real and imaginary parts of the refractive index tensor was examined. It was demonstrated that when polydispersity in particle size exists, an asymmetric orienting field such as a simple shear flow generally results in birefringence and dichroism that are not coaxial. It was also found that at a wavelength of 632.8 nm, the particles can be adequately modeled as Rayleigh scatterers, and absorption can be neglected. textcopyright 1988.},\n\tAuthor = {Johnson, S J and Fuller, G.G.},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {2},\n\tPages = {441--451},\n\tTitle = {The optical anisotropy of sheared hematite suspensions},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024067675&partnerID=40&md5=2f30b6e13022fd684648b029ed824ed4},\n\tVolume = {124},\n\tYear = {1988},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024067675&partnerID=40&md5=2f30b6e13022fd684648b029ed824ed4}}\n\n
\n
\n\n\n
\n Simultaneous linear dichroism and linear birefringence measurements were used to study sheared suspensions of colloidal hematite (α-Fe2O3) particles. Particles of three different axis ratios were synthesized. The particles were characterized using electron microscopy, and the three samples had axis ratios of 2.6 textpm 0.14, 3.61 textpm 0.24, and 5.82 textpm 0.58 with mean lengths of 0.315, 0.360, and 0.485 μm, repsectively. The particles were suspended in two Newtonian fluids, glycerin (viscosity = 17 P) and low molecular weight polybutene (viscosity = 220 P), and each suspension was subjected to transient shear flow in a Couette flow cell. The role of polydispersity in particle size on the relative orientations of the real and imaginary parts of the refractive index tensor was examined. It was demonstrated that when polydispersity in particle size exists, an asymmetric orienting field such as a simple shear flow generally results in birefringence and dichroism that are not coaxial. It was also found that at a wavelength of 632.8 nm, the particles can be adequately modeled as Rayleigh scatterers, and absorption can be neglected. textcopyright 1988.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The dichroism and birefringence of a hard-sphere suspension under shear.\n \n \n \n \n\n\n \n Wagner, N J; Fuller, G.; and Russel, W B\n\n\n \n\n\n\n The Journal of Chemical Physics, 89(3): 1580–1587. 1988.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{wagner_dichroism_1988,\n\tAbstract = {Optical measurements were used to detect structural anisotropy in concentrated dispersions over a range of Peclet numbers. Silica spheres of 49 and 130 run radii with grafted octadecyl chains were dispersed in cyclohexane at volume fractions from 0.1 to 0.4. The apparatus consisted of a Couette cell with the dispersion in the annulus probed by a HeNe laser beam parallel to the axis of rotation. The dichroism and birefringence of the transmitted beam varied linearly with shear rate at low Peclet numbers with an orientation coincident with the principle direction of shear. Increasing the Peclet number, by increasing the particle size, produced a nonlinear response with the orientation tending to align in the direction of flow. A theory coupling the nonequilibrium microstructure under shear to the optical properties of the suspension enables direct interpretation of the dichroism. Comparison of sample dichroism calculations for two different forms of the theory demonstrates that the optical technique can be used to discriminate between theories which predict the microstructure. textcopyright 1988 American Institute of Physics.},\n\tAuthor = {Wagner, N J and Fuller, G.G. and Russel, W B},\n\tJournal = {The Journal of Chemical Physics},\n\tNumber = {3},\n\tPages = {1580--1587},\n\tTitle = {The dichroism and birefringence of a hard-sphere suspension under shear},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000900567&partnerID=40&md5=b7cb0a1a9e9246396dff3a8aa97d41a9},\n\tVolume = {89},\n\tYear = {1988},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000900567&partnerID=40&md5=b7cb0a1a9e9246396dff3a8aa97d41a9}}\n\n
\n
\n\n\n
\n Optical measurements were used to detect structural anisotropy in concentrated dispersions over a range of Peclet numbers. Silica spheres of 49 and 130 run radii with grafted octadecyl chains were dispersed in cyclohexane at volume fractions from 0.1 to 0.4. The apparatus consisted of a Couette cell with the dispersion in the annulus probed by a HeNe laser beam parallel to the axis of rotation. The dichroism and birefringence of the transmitted beam varied linearly with shear rate at low Peclet numbers with an orientation coincident with the principle direction of shear. Increasing the Peclet number, by increasing the particle size, produced a nonlinear response with the orientation tending to align in the direction of flow. A theory coupling the nonequilibrium microstructure under shear to the optical properties of the suspension enables direct interpretation of the dichroism. Comparison of sample dichroism calculations for two different forms of the theory demonstrates that the optical technique can be used to discriminate between theories which predict the microstructure. textcopyright 1988 American Institute of Physics.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Uniaxial and biaxial extensional viscosity measurements of dilute and semi-dilute solutions of rigid rod polymers.\n \n \n \n \n\n\n \n Cathey, C A; and Fuller, G.\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 30(2-3): 303–316. 1988.\n \n\n\n\n
\n\n\n\n \n \n \"UniaxialPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{cathey_uniaxial_1988,\n\tAbstract = {Effective uniaxial extensional and biaxial extensional viscosities of dilute and semi-dilute solutions of collagen, a rigid rod molecule, have been measured with an opposing jet apparatus. The concentration of collagen in the glycerin/water solvent ranged from 50 to 2300 ppm. The data agree quantitatively with a theory developed by Batchelor describing the extensional viscosity of perfectly aligned rigid rods. The viscosity measured for the dilute rigid rod solutions is independent of the rate of strain as predicted by Batchelor's theory. Data taken on the semi-dilute, strain-thinning solutions at strain rates sufficiently high to align the rods in the extension direction also agree with the predictions of Batchelor's theory. The measured viscosity of semi-dilute solutions at low strain rates agree qualitatively with a theory developed by Doi and Edwards describing the strain-thinning behavior of semi-dilute rigid rod solutions. textcopyright 1988.},\n\tAuthor = {Cathey, C A and Fuller, G.G.},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {2-3},\n\tPages = {303--316},\n\tTitle = {Uniaxial and biaxial extensional viscosity measurements of dilute and semi-dilute solutions of rigid rod polymers},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024103882&partnerID=40&md5=28b4f029711f3f57d639baba829a03b1},\n\tVolume = {30},\n\tYear = {1988},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0024103882&partnerID=40&md5=28b4f029711f3f57d639baba829a03b1}}\n\n
\n
\n\n\n
\n Effective uniaxial extensional and biaxial extensional viscosities of dilute and semi-dilute solutions of collagen, a rigid rod molecule, have been measured with an opposing jet apparatus. The concentration of collagen in the glycerin/water solvent ranged from 50 to 2300 ppm. The data agree quantitatively with a theory developed by Batchelor describing the extensional viscosity of perfectly aligned rigid rods. The viscosity measured for the dilute rigid rod solutions is independent of the rate of strain as predicted by Batchelor's theory. Data taken on the semi-dilute, strain-thinning solutions at strain rates sufficiently high to align the rods in the extension direction also agree with the predictions of Batchelor's theory. The measured viscosity of semi-dilute solutions at low strain rates agree qualitatively with a theory developed by Doi and Edwards describing the strain-thinning behavior of semi-dilute rigid rod solutions. textcopyright 1988.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1987\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n EXTENSIONAL VISCOSITY MEASUREMENTS FOR LOW-VISCOSITY FLUIDS.\n \n \n \n \n\n\n \n Fuller, G. G; Cathey, C. A; Hubbard, B.; and Zebrowski, B. E\n\n\n \n\n\n\n Journal of Rheology, 31(3): 235–249. 1987.\n \n\n\n\n
\n\n\n\n \n \n \"EXTENSIONALPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_extensional_1987,\n\tAbstract = {A design for an extensional viscometer is described which is useful for studying low-viscosity fluids at high strain rates. The configuration consists of opposing nozzles through which the liquid is either sucked in or blown out. Measurement of the force required to keep the nozzles at a fixed distance apart as a function of flow rate directly yields a determination of the extensional viscosity. Data are presented on both Newtonian and non-Newtonian liquids. For Newtonian liquids, Trouton's rule is verified for both uniaxial extensional and compressive flows. Results are also presented on dilute solutions of polyacrylamide and Xanthan gum which indicate substantial enhancements of the extensional viscosity upon addition of these polymers. Qualitative differences were observed, however, for the extensional flow behavior of the flexible polyacrylamide chains and the more rigid and rodlike Xanthan gum chains.},\n\tAuthor = {Fuller, Gerald G and Cathey, Cheryl A and Hubbard, Brent and Zebrowski, Beth E},\n\tJournal = {Journal of Rheology},\n\tNumber = {3},\n\tPages = {235--249},\n\tTitle = {{EXTENSIONAL} {VISCOSITY} {MEASUREMENTS} {FOR} {LOW}-{VISCOSITY} {FLUIDS}.},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0023325347&partnerID=40&md5=81c5f104fda83b2fe7b73fde7dd33a41},\n\tVolume = {31},\n\tYear = {1987},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0023325347&partnerID=40&md5=81c5f104fda83b2fe7b73fde7dd33a41}}\n\n
\n
\n\n\n
\n A design for an extensional viscometer is described which is useful for studying low-viscosity fluids at high strain rates. The configuration consists of opposing nozzles through which the liquid is either sucked in or blown out. Measurement of the force required to keep the nozzles at a fixed distance apart as a function of flow rate directly yields a determination of the extensional viscosity. Data are presented on both Newtonian and non-Newtonian liquids. For Newtonian liquids, Trouton's rule is verified for both uniaxial extensional and compressive flows. Results are also presented on dilute solutions of polyacrylamide and Xanthan gum which indicate substantial enhancements of the extensional viscosity upon addition of these polymers. Qualitative differences were observed, however, for the extensional flow behavior of the flexible polyacrylamide chains and the more rigid and rodlike Xanthan gum chains.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The spatial development of transient couette flow and shear wave propagation in polymeric liquids by flow birefringence.\n \n \n \n \n\n\n \n Lee, J S; and Fuller, G.\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 26(1): 57–76. 1987.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lee_spatial_1987,\n\tAbstract = {In this paper the method of polarization modulated flow birefringence is us used to analyze the spatial and temporal dependence of the inception and cessation of Couette flow of polymeric liquids. Data are presented for both a Newtonian but birefringent polybutene oil and two solutions of poly(oxyethylene). The stress-optical rule is used to obtain the corresponding 'shear and normal stresses as functions of both time and space. The data for the Newtonian fluid are compared with analytical solutions to the momentum equations and are found to accurately follow the theoretical predictions. The polymer solutions, on the other hand, show a complex response with evidence for shear stress propagation as a wave of finite, constant speed. The measured wave speeds agree well with recent data reported by Joseph. Riccius and Arney [JFM 171(1986)309] on the same polymer solutions. textcopyright 1987.},\n\tAuthor = {Lee, J S and Fuller, G.G.},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {1},\n\tPages = {57--76},\n\tTitle = {The spatial development of transient couette flow and shear wave propagation in polymeric liquids by flow birefringence},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0023455625&partnerID=40&md5=0a394dfc7966654f5bfa97be198ab1c5},\n\tVolume = {26},\n\tYear = {1987},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0023455625&partnerID=40&md5=0a394dfc7966654f5bfa97be198ab1c5}}\n\n
\n
\n\n\n
\n In this paper the method of polarization modulated flow birefringence is us used to analyze the spatial and temporal dependence of the inception and cessation of Couette flow of polymeric liquids. Data are presented for both a Newtonian but birefringent polybutene oil and two solutions of poly(oxyethylene). The stress-optical rule is used to obtain the corresponding 'shear and normal stresses as functions of both time and space. The data for the Newtonian fluid are compared with analytical solutions to the momentum equations and are found to accurately follow the theoretical predictions. The polymer solutions, on the other hand, show a complex response with evidence for shear stress propagation as a wave of finite, constant speed. The measured wave speeds agree well with recent data reported by Joseph. Riccius and Arney [JFM 171(1986)309] on the same polymer solutions. textcopyright 1987.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The dynamics of colloidal particles suspended in a second-order fluid.\n \n \n \n \n\n\n \n Johnson, S J; and Fuller, G.\n\n\n \n\n\n\n Faraday Discussions of the Chemical Society, 83: 271–285. 1987.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{johnson_dynamics_1987,\n\tAbstract = {The transient flow response of colloidal particles in sheared suspensions has been examined with linear dichroism measurements. Data are reported for dilute suspensions of spheroidal haematite (α-Fe2O3) particles in both a Newtonian and a non-Newtonian Boger fluid consisting of 1000 ppm poly-isobutylene in low-molecular-weight polybutene. Haematite samples with average axis ratios of 2.66, 3.61 and 5.82 were used in the studies. Flow reversal and flow start-up experiments were conducted for shear rates up to 10 s-1 using both Couette and parallel-plate flow devices. The experimental data for suspensions in the non-Newtonian fluid are in qualitative agreement with predictions from the theory for particle motion in a second-order fluid. The particles experienced orbit drift and tended toward alignment along the vorticity axis of the flow. Also the period of rotation of the particles relative to that in the Newtonian suspending fluid increased with increasing shear rate. The predictions do not explicitly account for particle geometry, so experimental trends with axis ratio cannot be compared directly with the theory. In addition, Brownian motion is not considered by the theory, and we expect that the presence of Brownian motion in our suspensions strongly contributed to the steady-state distribution of particle orientation.},\n\tAuthor = {Johnson, S J and Fuller, G.G.},\n\tJournal = {Faraday Discussions of the Chemical Society},\n\tPages = {271--285},\n\tTitle = {The dynamics of colloidal particles suspended in a second-order fluid},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-37049078080&partnerID=40&md5=72e9edbc870b1b6e05a1c4530d3c246b},\n\tVolume = {83},\n\tYear = {1987},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-37049078080&partnerID=40&md5=72e9edbc870b1b6e05a1c4530d3c246b}}\n\n
\n
\n\n\n
\n The transient flow response of colloidal particles in sheared suspensions has been examined with linear dichroism measurements. Data are reported for dilute suspensions of spheroidal haematite (α-Fe2O3) particles in both a Newtonian and a non-Newtonian Boger fluid consisting of 1000 ppm poly-isobutylene in low-molecular-weight polybutene. Haematite samples with average axis ratios of 2.66, 3.61 and 5.82 were used in the studies. Flow reversal and flow start-up experiments were conducted for shear rates up to 10 s-1 using both Couette and parallel-plate flow devices. The experimental data for suspensions in the non-Newtonian fluid are in qualitative agreement with predictions from the theory for particle motion in a second-order fluid. The particles experienced orbit drift and tended toward alignment along the vorticity axis of the flow. Also the period of rotation of the particles relative to that in the Newtonian suspending fluid increased with increasing shear rate. The predictions do not explicitly account for particle geometry, so experimental trends with axis ratio cannot be compared directly with the theory. In addition, Brownian motion is not considered by the theory, and we expect that the presence of Brownian motion in our suspensions strongly contributed to the steady-state distribution of particle orientation.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Conservative dichroism of a sheared suspension in the Rayleigh-Gans light scattering approximation.\n \n \n \n \n\n\n \n Frattini, P L; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 119(2): 335–351. 1987.\n \n\n\n\n
\n\n\n\n \n \n \"ConservativePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{frattini_conservative_1987,\n\tAbstract = {A polarization modulation linear dichroism technique has recently been developed and applied to studies of the particle microdynamics associated with subjecting dilute colloidal suspensions to time-dependent flows. In this paper, we compare the abilities of the Rayleigh and Rayleigh-Gans light scattering approximations to represent experimental dichroism data in terms of calculations based on known hydrodynamic models. In particular, the importance of considering finite particle size in the light scattering theory is investigated. Calculations of the conservative linear dichroism and of the average particle orientation angle are carried out for dilute suspensions subjected to transient simple shear flow in the absence of Brownian motion and in the presence of weak Brownian motion. For both situations, only small differences in the features associated with the particle orientation statistics of the simulation occur between the Rayleigh and the Rayleigh-Gans calculations over a wide range in particle geometry. In particular, the calculation of the average particle orientation angle is found to be insensitive to the choice of light scattering treatment. Generally, we infer from these simulations that the additional complexity of the Rayleigh-Gans treatment is usually not warranted for the analysis of typical rheo-optical data from dilute suspensions. textcopyright 1987.},\n\tAuthor = {Frattini, P L and Fuller, G.G.},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {2},\n\tPages = {335--351},\n\tTitle = {Conservative dichroism of a sheared suspension in the {Rayleigh}-{Gans} light scattering approximation},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0023437432&partnerID=40&md5=ca97301fd07acde13a313e0d64de41c8},\n\tVolume = {119},\n\tYear = {1987},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0023437432&partnerID=40&md5=ca97301fd07acde13a313e0d64de41c8}}\n\n
\n
\n\n\n
\n A polarization modulation linear dichroism technique has recently been developed and applied to studies of the particle microdynamics associated with subjecting dilute colloidal suspensions to time-dependent flows. In this paper, we compare the abilities of the Rayleigh and Rayleigh-Gans light scattering approximations to represent experimental dichroism data in terms of calculations based on known hydrodynamic models. In particular, the importance of considering finite particle size in the light scattering theory is investigated. Calculations of the conservative linear dichroism and of the average particle orientation angle are carried out for dilute suspensions subjected to transient simple shear flow in the absence of Brownian motion and in the presence of weak Brownian motion. For both situations, only small differences in the features associated with the particle orientation statistics of the simulation occur between the Rayleigh and the Rayleigh-Gans calculations over a wide range in particle geometry. In particular, the calculation of the average particle orientation angle is found to be insensitive to the choice of light scattering treatment. Generally, we infer from these simulations that the additional complexity of the Rayleigh-Gans treatment is usually not warranted for the analysis of typical rheo-optical data from dilute suspensions. textcopyright 1987.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1986\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Optical Rheometry of Dispersions.\n \n \n \n\n\n \n Fuller, G.; Johnson, S J; and Salem, A J\n\n\n \n\n\n\n In Proceedings of the Tenth US National Congress of Applied Mechanics, pages 149–153, 1986. The American Society of Mechanical Engineers, New York\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@inproceedings{fuller_optical_1986,\n\tAuthor = {Fuller, G.G. and Johnson, S J and Salem, A J},\n\tBooktitle = {Proceedings of the {Tenth} {US} {National} {Congress} of {Applied} {Mechanics}},\n\tPages = {149--153},\n\tPublisher = {The American Society of Mechanical Engineers, New York},\n\tTitle = {Optical {Rheometry} of {Dispersions}},\n\tYear = {1986}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n RHEO-OPTICAL STUDIES OF THE EFFECT OF WEAK BROWNIAN ROTATIONS IN SHEARED SUSPENSIONS.\n \n \n \n\n\n \n Frattini, P. L; and Fuller, G. G\n\n\n \n\n\n\n Journal of Fluid Mechanics, 168: 119–150. 1986.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{frattini_rheo-optical_1986,\n\tAbstract = {The orientation state of a dilute suspension of rigid, axisymmetric particles subjected to the sudden inception of simple shearing flow in the presence of small Brownian couples is investigated experimentally. In contrast to mechanical rheometric methods, the experimental technique, a recently developed optical method based on conservative linear dichroism, provides a direct probe of the constituent particle dynamics. The average particle orientation direction projected into the shear plane and the average degree of alignment about that direction are measured as functions of time for Peclet numbers ranging from 20 to 200. The data provide, for the first time, an experimental basis with which the closure approximation required for solution of the evolution equation for (u, u), the second moment of the particle orientation distribution, may be critically evaluated in shearing flows. Two closures are examined: the ad hoc but often-used pre-averaging closure, and the first-order closure suggested by E. J. Hinch \\{\\vphantom{\\}}\\&},\n\tAuthor = {Frattini, Paul L and Fuller, Gerald G},\n\tJournal = {Journal of Fluid Mechanics},\n\tPages = {119--150},\n\tTitle = {{RHEO}-{OPTICAL} {STUDIES} {OF} {THE} {EFFECT} {OF} {WEAK} {BROWNIAN} {ROTATIONS} {IN} {SHEARED} {SUSPENSIONS}.},\n\tVolume = {168},\n\tYear = {1986}}\n\n
\n
\n\n\n
\n The orientation state of a dilute suspension of rigid, axisymmetric particles subjected to the sudden inception of simple shearing flow in the presence of small Brownian couples is investigated experimentally. In contrast to mechanical rheometric methods, the experimental technique, a recently developed optical method based on conservative linear dichroism, provides a direct probe of the constituent particle dynamics. The average particle orientation direction projected into the shear plane and the average degree of alignment about that direction are measured as functions of time for Peclet numbers ranging from 20 to 200. The data provide, for the first time, an experimental basis with which the closure approximation required for solution of the evolution equation for (u, u), the second moment of the particle orientation distribution, may be critically evaluated in shearing flows. Two closures are examined: the ad hoc but often-used pre-averaging closure, and the first-order closure suggested by E. J. Hinch \\p̌hantom\\&\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Flowing colloidal suspensions in non-Newtonian suspending fluids: Decoupling the composite birefringence.\n \n \n \n \n\n\n \n Johnson, S J; and Fuller, G.\n\n\n \n\n\n\n Rheologica Acta, 25(4): 405–417. 1986.\n \n\n\n\n
\n\n\n\n \n \n \"FlowingPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{johnson_flowing_1986,\n\tAbstract = {The rheology of dilute, colloidal suspensions in polymeric suspending fluids can be studied with simultaneous dichroism and birefringence measurements. The dichroism provides a direct measure of the particle dynamics, but the birefringence is a composite property with independent contributions from the suspended particles and the polymer molecules. For suspensions where the contribution from the particles is significant, the composite birefringence must be decoupled in order to analyze the dynamics of the polymeric suspending fluid. A method to perform the decoupling is derived and then demonstrated through transient shear flow experiments with dilute suspensions of βFeOOH particles in semi-dilute, xanthan gum suspending fluids. The birefringence of the xanthan gum suspending fluid is calculated from experimental measurements of the composite birefringence and the dichroism of the suspension. To gather information on particle/polymer interactions, the calculated birefringence is compared to the birefringence of xanthan gum solutions containing no suspended particles and the dirchoism is compared to that of a suspension in a Newtonian fluid. textcopyright 1986 Steinkopff.},\n\tAuthor = {Johnson, S J and Fuller, G.G.},\n\tJournal = {Rheologica Acta},\n\tNumber = {4},\n\tPages = {405--417},\n\tTitle = {Flowing colloidal suspensions in non-{Newtonian} suspending fluids: {Decoupling} the composite birefringence},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0006654099&partnerID=40&md5=1beb2710a50a79820207fb2ee33dac38},\n\tVolume = {25},\n\tYear = {1986},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0006654099&partnerID=40&md5=1beb2710a50a79820207fb2ee33dac38}}\n\n
\n
\n\n\n
\n The rheology of dilute, colloidal suspensions in polymeric suspending fluids can be studied with simultaneous dichroism and birefringence measurements. The dichroism provides a direct measure of the particle dynamics, but the birefringence is a composite property with independent contributions from the suspended particles and the polymer molecules. For suspensions where the contribution from the particles is significant, the composite birefringence must be decoupled in order to analyze the dynamics of the polymeric suspending fluid. A method to perform the decoupling is derived and then demonstrated through transient shear flow experiments with dilute suspensions of βFeOOH particles in semi-dilute, xanthan gum suspending fluids. The birefringence of the xanthan gum suspending fluid is calculated from experimental measurements of the composite birefringence and the dichroism of the suspension. To gather information on particle/polymer interactions, the calculated birefringence is compared to the birefringence of xanthan gum solutions containing no suspended particles and the dirchoism is compared to that of a suspension in a Newtonian fluid. textcopyright 1986 Steinkopff.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1985\n \n \n (11)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Dynamics of rigid dumbbells in confined geometries Part II. Time-dependent shear flow.\n \n \n \n \n\n\n \n Park, O O.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 18(2): 111–122. January 1985.\n \n\n\n\n
\n\n\n\n \n \n \"DynamicsPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{park_dynamics_1985,\n\tAbstract = {This paper describes analytical solutions to the response of a rigid dumbbell to a time-dependent simple shear flow in a channel of arbitrary size. The results were applied for specific flows, such as the sudden inception and cessation of simple shear and oscillatory shear flow. The introduction of confining walls leads to several qualitative differences compared with an unbounded flow such as initial negative jumps in the contribution of the dumbbell to the viscosity and the first normal stress coefficient. In addition, wall effects lead to a breakdown of well-known relationships between the complex viscosity and the complex normal stress coefficient during oscillatory flow.},\n\tAuthor = {Park, O Ok and Fuller, Gerald G},\n\tDoi = {10.1016/0377-0257(85)85016-3},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tLanguage = {English},\n\tMonth = jan,\n\tNumber = {2},\n\tPages = {111--122},\n\tTitle = {Dynamics of rigid dumbbells in confined geometries {Part} {II}. {Time}-dependent shear flow},\n\tUrl = {http://linkinghub.elsevier.com/retrieve/pii/0377025785850163},\n\tVolume = {18},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://linkinghub.elsevier.com/retrieve/pii/0377025785850163},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1016/0377-0257(85)85016-3}}\n\n
\n
\n\n\n
\n This paper describes analytical solutions to the response of a rigid dumbbell to a time-dependent simple shear flow in a channel of arbitrary size. The results were applied for specific flows, such as the sudden inception and cessation of simple shear and oscillatory shear flow. The introduction of confining walls leads to several qualitative differences compared with an unbounded flow such as initial negative jumps in the contribution of the dumbbell to the viscosity and the first normal stress coefficient. In addition, wall effects lead to a breakdown of well-known relationships between the complex viscosity and the complex normal stress coefficient during oscillatory flow.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Small angle light scattering as a probe of flow-induced particle orientation.\n \n \n \n \n\n\n \n Salem, A J; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 108(1): 149–157. 1985.\n \n\n\n\n
\n\n\n\n \n \n \"SmallPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{salem_small_1985,\n\tAbstract = {A new optical technique which digitally records the evolution of the two-dimensional intensity distribution of the small angle light scattering of suspended particles subjected to transient shear flows is described. For dilute suspensions of absorbing particles which are larger than the wavelength of the incident light, scalar diffraction theory is applied to relate the moments of the measured pattern to the degree of alignment and average orientation angle of the particles. The diffraction pattern for non-Brownian, near-sphere particles in Stokes flow is solved and shows an oscillation between isotropic contours, when the particles are randomly oriented, and elliptical contours oriented perpendicular to the average orientation of the particles. Experimental results for the inception and reversal of simple shear flow of a dilute suspension of prolate ellipsoids in a parallel plate device are presented and exhibit this behavior. These data agree well with previous theoretical and experimental findings but bring with them new information that will prove valuable in future rheological studies. textcopyright 1985.},\n\tAuthor = {Salem, A J and Fuller, G.G.},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {1},\n\tPages = {149--157},\n\tTitle = {Small angle light scattering as a probe of flow-induced particle orientation},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0022162231&partnerID=40&md5=92f0f5716ba242c38d1c8243e3265dac},\n\tVolume = {108},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0022162231&partnerID=40&md5=92f0f5716ba242c38d1c8243e3265dac}}\n\n
\n
\n\n\n
\n A new optical technique which digitally records the evolution of the two-dimensional intensity distribution of the small angle light scattering of suspended particles subjected to transient shear flows is described. For dilute suspensions of absorbing particles which are larger than the wavelength of the incident light, scalar diffraction theory is applied to relate the moments of the measured pattern to the degree of alignment and average orientation angle of the particles. The diffraction pattern for non-Brownian, near-sphere particles in Stokes flow is solved and shows an oscillation between isotropic contours, when the particles are randomly oriented, and elliptical contours oriented perpendicular to the average orientation of the particles. Experimental results for the inception and reversal of simple shear flow of a dilute suspension of prolate ellipsoids in a parallel plate device are presented and exhibit this behavior. These data agree well with previous theoretical and experimental findings but bring with them new information that will prove valuable in future rheological studies. textcopyright 1985.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n RHEO-OPTICAL STUDIES OF CONCENTRATED POLYSTYRENE SOLUTIONS SUBJECTED TO TRANSIENT SIMPLE SHEAR FLOW.\n \n \n \n \n\n\n \n Zebrowski, B. E; and Fuller, G. G\n\n\n \n\n\n\n Journal of polymer science. Part A-2, Polymer physics, 23(3): 575–589. 1985.\n \n\n\n\n
\n\n\n\n \n \n \"RHEO-OPTICALPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{zebrowski_rheo-optical_1985,\n\tAbstract = {Concentrated polystyrene solutions were investigated on the inception and cessation of simple shear flow by means of the technique of two-color flow birefringence. Both monodisperse solutions of various molecular weights and biomodal mixtures were studied. The molecular weight affected both the amount of overshoot in the birefringence and the response time on the inception of shear flow. Large overshoots in birefringence, up to 250\\%, and undershoot in the orientation angle were observed. The shear stress and the first normal stress difference were calculated by using the stress-optical rule. The amount of strain at the peaks in the stress growth curves are presented along with the steady-state viscosity and primary normal stress coefficient. The experimental results are compared qualitatively with theoretical predictions of various molecular models.},\n\tAuthor = {Zebrowski, Beth E and Fuller, Gerald G},\n\tJournal = {Journal of polymer science. Part A-2, Polymer physics},\n\tNumber = {3},\n\tPages = {575--589},\n\tTitle = {{RHEO}-{OPTICAL} {STUDIES} {OF} {CONCENTRATED} {POLYSTYRENE} {SOLUTIONS} {SUBJECTED} {TO} {TRANSIENT} {SIMPLE} {SHEAR} {FLOW}.},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0343368609&partnerID=40&md5=f42a7ea8b7243b2e7527b06cf7745dee},\n\tVolume = {23},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0343368609&partnerID=40&md5=f42a7ea8b7243b2e7527b06cf7745dee}}\n\n
\n
\n\n\n
\n Concentrated polystyrene solutions were investigated on the inception and cessation of simple shear flow by means of the technique of two-color flow birefringence. Both monodisperse solutions of various molecular weights and biomodal mixtures were studied. The molecular weight affected both the amount of overshoot in the birefringence and the response time on the inception of shear flow. Large overshoots in birefringence, up to 250%, and undershoot in the orientation angle were observed. The shear stress and the first normal stress difference were calculated by using the stress-optical rule. The amount of strain at the peaks in the stress growth curves are presented along with the steady-state viscosity and primary normal stress coefficient. The experimental results are compared qualitatively with theoretical predictions of various molecular models.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The effect of segment/boundary hydrodynamic interactions on the dynamics of adsorbed polymer chains subjected to flow.\n \n \n \n \n\n\n \n Lee, J J; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 107(2): 308–313. 1985.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lee_effect_1985,\n\tAbstract = {In this paper the bead and spring model of a polymer chain adsorbed onto a solid boundary in the presence of flow is extended to include the effects of hydrodynamic interactions between the polymer segments and the solid surface. In addition, the consequence of the finite dimension of the polymer chain is considered as well as the finite extensibility of the chain. The addition of hydrodynamic interactions is shown to increase the rate at which the imposed flow can compress the adsorbed layer. As would be expected, however, allowing the chain to be of finite dimension decreases the compressibility of the attached chains. textcopyright 1985.},\n\tAuthor = {Lee, J J and Fuller, G.G.},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {2},\n\tPages = {308--313},\n\tTitle = {The effect of segment/boundary hydrodynamic interactions on the dynamics of adsorbed polymer chains subjected to flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0022145195&partnerID=40&md5=d3d46750d619c99bc11414ff0555a75f},\n\tVolume = {107},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0022145195&partnerID=40&md5=d3d46750d619c99bc11414ff0555a75f}}\n\n
\n
\n\n\n
\n In this paper the bead and spring model of a polymer chain adsorbed onto a solid boundary in the presence of flow is extended to include the effects of hydrodynamic interactions between the polymer segments and the solid surface. In addition, the consequence of the finite dimension of the polymer chain is considered as well as the finite extensibility of the chain. The addition of hydrodynamic interactions is shown to increase the rate at which the imposed flow can compress the adsorbed layer. As would be expected, however, allowing the chain to be of finite dimension decreases the compressibility of the attached chains. textcopyright 1985.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Simultaneous dichroism and birefringence measurements of dilute colloidal suspensions in transient shear flow.\n \n \n \n \n\n\n \n Johnson, S J; Frattini, P L; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 104(2): 440–455. 1985.\n \n\n\n\n
\n\n\n\n \n \n \"SimultaneousPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{johnson_simultaneous_1985,\n\tAbstract = {A new experimental technique is presented for the study of the dynamics of dilute colloidal suspensions in transient shear flow. Based on the polarization modulation of light, the experiment permits simultaneous determinations of linear dichroism (Δntextacutedbl), linear birefringence (Δntextasciiacute), the orientation angle of the principal axis of Δntextacutedbl, and the orientation angle of the principal axis of Δntextasciiacute. All four calculated quantities are functions of time in transient flow situations. A detailed mathematical analysis of the experiment is presented, and six limiting cases of the general mathematical results are derived. To demonstrate the experimental technique, a 400 ppm suspension of ferric hydrous oxide (βFeOOH) particles in 97\\% glycerine and 3\\% water was subjected to transient, simple shear flow in a Couette flow cell. The βFeOOH suspension represents the limiting case of very small Δntextacutedbl and Δntextasciiacute which allows a clear demonstration of the experimental method. Magnitudes of 10 -7 for both Δntextacutedbl and Δntextasciiacute were measured experimentally at shear rates of 4, 8, and 24 sec -1 in flow reversal sequences. In addition, Δntextacutedbl and Δntextasciiacute were found to be coaxial. Simultaneous dichroism and birefringence measurements provide a means of studying the dynamics of suspensions in non-Newtonian suspending fluids. The application of this experimental method to the study of particles suspended in non-Newtonian, polymer fluids is discussed. textcopyright 1985.},\n\tAuthor = {Johnson, S J and Frattini, P L and Fuller, G.G.},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {2},\n\tPages = {440--455},\n\tTitle = {Simultaneous dichroism and birefringence measurements of dilute colloidal suspensions in transient shear flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0022042704&partnerID=40&md5=9280063200612b78c3a21e734222731f},\n\tVolume = {104},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0022042704&partnerID=40&md5=9280063200612b78c3a21e734222731f}}\n\n
\n
\n\n\n
\n A new experimental technique is presented for the study of the dynamics of dilute colloidal suspensions in transient shear flow. Based on the polarization modulation of light, the experiment permits simultaneous determinations of linear dichroism (Δntextacutedbl), linear birefringence (Δntextasciiacute), the orientation angle of the principal axis of Δntextacutedbl, and the orientation angle of the principal axis of Δntextasciiacute. All four calculated quantities are functions of time in transient flow situations. A detailed mathematical analysis of the experiment is presented, and six limiting cases of the general mathematical results are derived. To demonstrate the experimental technique, a 400 ppm suspension of ferric hydrous oxide (βFeOOH) particles in 97% glycerine and 3% water was subjected to transient, simple shear flow in a Couette flow cell. The βFeOOH suspension represents the limiting case of very small Δntextacutedbl and Δntextasciiacute which allows a clear demonstration of the experimental method. Magnitudes of 10 -7 for both Δntextacutedbl and Δntextasciiacute were measured experimentally at shear rates of 4, 8, and 24 sec -1 in flow reversal sequences. In addition, Δntextacutedbl and Δntextasciiacute were found to be coaxial. Simultaneous dichroism and birefringence measurements provide a means of studying the dynamics of suspensions in non-Newtonian suspending fluids. The application of this experimental method to the study of particles suspended in non-Newtonian, polymer fluids is discussed. textcopyright 1985.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rheooptical response of rodlike, shortened collagen protein to transient shear flow.\n \n \n \n \n\n\n \n Chow, A. W; Fuller, G. G; Wallace, D. G; and Madri, J. A\n\n\n \n\n\n\n Macromolecules, 18(4): 805–810. July 1985.\n \n\n\n\n
\n\n\n\n \n \n \"RheoopticalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chow_rheooptical_1985,\n\tAuthor = {Chow, Andrea W and Fuller, Gerald G and Wallace, Donald G and Madri, Joseph A},\n\tDoi = {10.1021/ma00146a035},\n\tJournal = {Macromolecules},\n\tMonth = jul,\n\tNumber = {4},\n\tPages = {805--810},\n\tTitle = {Rheooptical response of rodlike, shortened collagen protein to transient shear flow},\n\tUrl = {http://dx.doi.org/10.1021/ma00146a035},\n\tVolume = {18},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://dx.doi.org/10.1021/ma00146a035}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n ADSORPTION AND DESORPTION OF FLEXIBLE POLYMER CHAINS IN FLOWING SYSTEMS.\n \n \n \n \n\n\n \n Lee, J.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 103(2): 578–585. 1985.\n \n\n\n\n
\n\n\n\n \n \n \"ADSORPTIONPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lee_adsorption_1985,\n\tAbstract = {Hydrodynamic theory predicts that the retardation of the rate of site of a bubble as a function of surface active solute concentration reflects the transition from a liquid to a rigid interface. Ellipsometry has been applied to the study of adsorption and desorption of flexible polymer chains in flowing systems. This optical technique produces a measurement of both film thickness and adsorbance under velocity gradients up to 7800 sec** minus **1. Three molecular weight samples of monodisperse polystyrene ranging from 4 to 20 million were studied using cyclohexane at the theta condition as solvent. Bulk concentrations corresponded to saturated surface coverages. Chrome was used as the adsorbent. For preadsorbed polymer layers the results indicate flow-enhanced desorption whereas hydrodynamic forces were found to inhibit the adsorption of polymer chains from flowing solutions into onto initially clean surfaces.},\n\tAuthor = {Lee, Jen-Jiang and Fuller, Gerald G},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {2},\n\tPages = {578--585},\n\tTitle = {{ADSORPTION} {AND} {DESORPTION} {OF} {FLEXIBLE} {POLYMER} {CHAINS} {IN} {FLOWING} {SYSTEMS}.},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0022010172&partnerID=40&md5=e8bab4d9cc9b5c9b2286fd1bf9cce524},\n\tVolume = {103},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0022010172&partnerID=40&md5=e8bab4d9cc9b5c9b2286fd1bf9cce524}}\n\n
\n
\n\n\n
\n Hydrodynamic theory predicts that the retardation of the rate of site of a bubble as a function of surface active solute concentration reflects the transition from a liquid to a rigid interface. Ellipsometry has been applied to the study of adsorption and desorption of flexible polymer chains in flowing systems. This optical technique produces a measurement of both film thickness and adsorbance under velocity gradients up to 7800 sec** minus **1. Three molecular weight samples of monodisperse polystyrene ranging from 4 to 20 million were studied using cyclohexane at the theta condition as solvent. Bulk concentrations corresponded to saturated surface coverages. Chrome was used as the adsorbent. For preadsorbed polymer layers the results indicate flow-enhanced desorption whereas hydrodynamic forces were found to inhibit the adsorption of polymer chains from flowing solutions into onto initially clean surfaces.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Some experimental results on the development of Couette flow for non-Newtonian fluids.\n \n \n \n \n\n\n \n Chow, A W; and Fuller, G.\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 17(2): 233–243. 1985.\n \n\n\n\n
\n\n\n\n \n \n \"SomePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chow_experimental_1985,\n\tAbstract = {This paper reports some experimental results on the time development of a Couette flow following the start-up of shear flow using the technique of two-color flow birefringence. Measurements obtained on collagen solutions are consistent with two theoretical studies which predict that for some viscoelastic liquids, momentum is transferred from the moving Couette cell boundary to the interior of the fluid through a velocity wave propagating and reflecting between the cell boundaries. This non-Newtonian phenomenon, exhibited as an oscillatory response in the measured birefringence and orientation angle, is observed at shear rates above a critical value when the response time of the polymer solution approaches the flow development time in the Couette flow cell. textcopyright 1985.},\n\tAuthor = {Chow, A W and Fuller, G.G.},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {2},\n\tPages = {233--243},\n\tTitle = {Some experimental results on the development of {Couette} flow for non-{Newtonian} fluids},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0022029538&partnerID=40&md5=025539a051bed4e4c9fd96f687bd5a68},\n\tVolume = {17},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0022029538&partnerID=40&md5=025539a051bed4e4c9fd96f687bd5a68}}\n\n
\n
\n\n\n
\n This paper reports some experimental results on the time development of a Couette flow following the start-up of shear flow using the technique of two-color flow birefringence. Measurements obtained on collagen solutions are consistent with two theoretical studies which predict that for some viscoelastic liquids, momentum is transferred from the moving Couette cell boundary to the interior of the fluid through a velocity wave propagating and reflecting between the cell boundaries. This non-Newtonian phenomenon, exhibited as an oscillatory response in the measured birefringence and orientation angle, is observed at shear rates above a critical value when the response time of the polymer solution approaches the flow development time in the Couette flow cell. textcopyright 1985.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rheo-Optical Studies of Polyelectrolyte Solutions in Simple Shear Flow.\n \n \n \n \n\n\n \n Zebrowski, B E; and Fuller, G.\n\n\n \n\n\n\n Journal of Rheology (1978-present). 1985.\n \n\n\n\n
\n\n\n\n \n \n \"Rheo-OpticalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{zebrowski_rheo-optical_1985-1,\n\tAbstract = {... It was observed in these studies that the longest relaxation times increase with the coil expansion resulting from ... separate measurements to determine the two unknowns in a flowbirefringence experiment, the flow -induced optical anisotropy or ... RHEO - OPTICS IN SHEAR FLOW ...},\n\tAuthor = {Zebrowski, B E and Fuller, G.G.},\n\tJournal = {Journal of Rheology (1978-present)},\n\tTitle = {Rheo-{Optical} {Studies} of {Polyelectrolyte} {Solutions} in {Simple} {Shear} {Flow}},\n\tUrl = {http://scitation.aip.org/content/sor/journal/jor2/29/6/10.1122/1.549823},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://scitation.aip.org/content/sor/journal/jor2/29/6/10.1122/1.549823}}\n\n
\n
\n\n\n
\n ... It was observed in these studies that the longest relaxation times increase with the coil expansion resulting from ... separate measurements to determine the two unknowns in a flowbirefringence experiment, the flow -induced optical anisotropy or ... RHEO - OPTICS IN SHEAR FLOW ...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rheooptical response of rodlike chains subject to transient shear flow. 2. Two-color flow birefringence measurements on collagen protein.\n \n \n \n \n\n\n \n Chow, A. W; Fuller, G. G; Wallace, D. G; and Madri, J. A\n\n\n \n\n\n\n Macromolecules, 18(4): 793–804. July 1985.\n \n\n\n\n
\n\n\n\n \n \n \"RheoopticalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chow_rheooptical_1985-1,\n\tAuthor = {Chow, Andrea W and Fuller, Gerald G and Wallace, Donald G and Madri, Joseph A},\n\tDoi = {10.1021/ma00146a034},\n\tJournal = {Macromolecules},\n\tMonth = jul,\n\tNumber = {4},\n\tPages = {793--804},\n\tTitle = {Rheooptical response of rodlike chains subject to transient shear flow. 2. {Two}-color flow birefringence measurements on collagen protein},\n\tUrl = {http://dx.doi.org/10.1021/ma00146a034},\n\tVolume = {18},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://dx.doi.org/10.1021/ma00146a034}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rheooptical response of rodlike chains subject to transient shear flow. 1. Model calculations on the effects of polydispersity.\n \n \n \n \n\n\n \n Chow, A W; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 18(4): 786–793. 1985.\n \n\n\n\n
\n\n\n\n \n \n \"RheoopticalPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chow_rheooptical_1985-2,\n\tAbstract = {This paper is the first part of a two-part series investigating the flow dynamics of semidilute solutions of rodlike chains. Specifically, the effects of polydispersity are considered. In this paper, numerical solutions of the flow birefringence Δn and the average orientation angle χ for the Doi-Edwards model for monodisperse systems and the Doi-Edwards-Marrucci-Grizzuti model for polydisperse systems are studied. These models, to be tested in part 2 by experimental results on rodlike collagen proteins using the method of two-color flow birefringence, indicate that the rheooptical properties are very sensitive to the high molecular weight components. A polydisperse solution with a small quantity of high molecular weight chains exhibits substantially different behavior than a monodisperse solution, especially under transient flow conditions. In addition, the birefringence overshoot predicted for the rodlike system was found to be much smaller compared to that observed for flexible systems. textcopyright 1985 American Chemical Society.},\n\tAuthor = {Chow, A W and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {4},\n\tPages = {786--793},\n\tTitle = {Rheooptical response of rodlike chains subject to transient shear flow. 1. {Model} calculations on the effects of polydispersity},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0141603833&partnerID=40&md5=4d9a804da7f8e591ecdb5976c2017ab4},\n\tVolume = {18},\n\tYear = {1985},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0141603833&partnerID=40&md5=4d9a804da7f8e591ecdb5976c2017ab4}}\n\n
\n
\n\n\n
\n This paper is the first part of a two-part series investigating the flow dynamics of semidilute solutions of rodlike chains. Specifically, the effects of polydispersity are considered. In this paper, numerical solutions of the flow birefringence Δn and the average orientation angle χ for the Doi-Edwards model for monodisperse systems and the Doi-Edwards-Marrucci-Grizzuti model for polydisperse systems are studied. These models, to be tested in part 2 by experimental results on rodlike collagen proteins using the method of two-color flow birefringence, indicate that the rheooptical properties are very sensitive to the high molecular weight components. A polydisperse solution with a small quantity of high molecular weight chains exhibits substantially different behavior than a monodisperse solution, especially under transient flow conditions. In addition, the birefringence overshoot predicted for the rodlike system was found to be much smaller compared to that observed for flexible systems. textcopyright 1985 American Chemical Society.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1984\n \n \n (7)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n RESPONSE OF MODERA℡Y CONCENTRATED XANTHAN GUM SOLUTIONS TO TIME-DEPENDENT FLOWS USING TWO-COLOR FLOW BIREFRINGENCE.\n \n \n \n \n\n\n \n Chow, A. W; and Fuller, G. G\n\n\n \n\n\n\n Journal of Rheology, 28(1): 23–43. 1984.\n \n\n\n\n
\n\n\n\n \n \n \"RESPONSEPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{chow_response_1984,\n\tAbstract = {A new optical arrangement for the flow birefringence experiment, based on a two-color laser, is described. Using this system, flow birefringence measurements have been performed in transient flow conditions with the birefringence DELTA n and orientation angle chi of the sample being obtained simultaneously in time. Measurements using a quarter-wave retarder indicate that the experimental accuracy is 2. 2\\% for the birefringence and 0. 15 degree for the orientation angle. The minimum birefringence detectable by this system is estimated to be 2 multiplied by 10** minus **7. Using this technique, DELTA n and chi of moderately concentrated xanthan solutions were measured following the inception and cessation of simple shear. The effects of increased ionic strength by adding NaCl to the xanthan solution were also examined. A qualitative comparison of measurements between the no salt and 0. 1M NaCl solutions indicates that the macromolecules might be somewhat stiffened in the presence of salt, which agrees with results reported previously in the literature.},\n\tAuthor = {Chow, Andrea W and Fuller, Gerald G},\n\tJournal = {Journal of Rheology},\n\tNumber = {1},\n\tPages = {23--43},\n\tTitle = {{RESPONSE} {OF} {MODERA}℡{Y} {CONCENTRATED} {XANTHAN} {GUM} {SOLUTIONS} {TO} {TIME}-{DEPENDENT} {FLOWS} {USING} {TWO}-{COLOR} {FLOW} {BIREFRINGENCE}.},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0021376257&partnerID=40&md5=f5bc72a1466e0f3a1bc8e0015cb80a1a},\n\tVolume = {28},\n\tYear = {1984},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0021376257&partnerID=40&md5=f5bc72a1466e0f3a1bc8e0015cb80a1a}}\n\n
\n
\n\n\n
\n A new optical arrangement for the flow birefringence experiment, based on a two-color laser, is described. Using this system, flow birefringence measurements have been performed in transient flow conditions with the birefringence DELTA n and orientation angle chi of the sample being obtained simultaneously in time. Measurements using a quarter-wave retarder indicate that the experimental accuracy is 2. 2% for the birefringence and 0. 15 degree for the orientation angle. The minimum birefringence detectable by this system is estimated to be 2 multiplied by 10** minus **7. Using this technique, DELTA n and chi of moderately concentrated xanthan solutions were measured following the inception and cessation of simple shear. The effects of increased ionic strength by adding NaCl to the xanthan solution were also examined. A qualitative comparison of measurements between the no salt and 0. 1M NaCl solutions indicates that the macromolecules might be somewhat stiffened in the presence of salt, which agrees with results reported previously in the literature.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Polymer Adsorption and Dispersion Stability.\n \n \n \n\n\n \n \n\n\n \n\n\n\n American Chemical Society, February 1984.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@book{_polymer_1984,\n\tIsbn = {0-8412-0820-4},\n\tMonth = feb,\n\tPublisher = {American Chemical Society},\n\tTitle = {Polymer {Adsorption} and {Dispersion} {Stability}},\n\tYear = {1984}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Note: A Note on Phase-Modulated Flow Birefringence: A Promising Rheo-Optical Method.\n \n \n \n\n\n \n Frattini, P. L; and Fuller, G. G\n\n\n \n\n\n\n Journal of Rheology (1978-present), 28(1): 61–70. 1984.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{frattini_note:_1984,\n\tAbstract = {Rheo - optical determination of flow birefringence and flow dichroism with the pulsed laser method Christian Clasen et al ... 2000)JCPSA6000113000022010246000001; Wetting-induced anisotropic structure at the interface of a glass and a sponge phase M. Magalh\\{\\vphantom{\\}}{\\textbackslash}textasciitilde},\n\tAuthor = {Frattini, Paul L and Fuller, Gerald G},\n\tDoi = {10.1122/1.549768},\n\tJournal = {Journal of Rheology (1978-present)},\n\tNumber = {1},\n\tPages = {61--70},\n\tTitle = {Note: {A} {Note} on {Phase}-{Modulated} {Flow} {Birefringence}: {A} {Promising} {Rheo}-{Optical} {Method}},\n\tVolume = {28},\n\tYear = {1984},\n\tBdsk-Url-1 = {http://dx.doi.org/10.1122/1.549768}}\n\n
\n
\n\n\n
\n Rheo - optical determination of flow birefringence and flow dichroism with the pulsed laser method Christian Clasen et al ... 2000)JCPSA6000113000022010246000001; Wetting-induced anisotropic structure at the interface of a glass and a sponge phase M. Magalh\\p̌hantom\\\\textasciitilde\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Dynamics of rigid and flexible polymer chains in confined geometries. part I: steady simple shear flow.\n \n \n \n \n\n\n \n Park, O O.; and Fuller, G. G\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 15(3): 309–329. January 1984.\n \n\n\n\n
\n\n\n\n \n \n \"DynamicsPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{park_dynamics_1984,\n\tAbstract = {This paper describes exact solutions to the response of both the elastic and rigid dumbbell models to a steady simple shear flow in a channel having a length scale comparable to the dumbbells themselves. Results are given for rheological properties over the entire range of the ratio of the channel width to the length of the dumbbell. It is found that both models lead to a decrease in viscosity as the channel is reduced in size with the elastic dumbbell predicting a stronger dependence on that parameter compared with the rigid dumbbell. The elastic dumbbell predicts shear independent rheological properties whereas the rigid dumbell predicts shear thinning as in the case of unbounded flows. The rate of shear thinning, however, decreases with decreasing channel width.},\n\tAuthor = {Park, O Ok and Fuller, Gerald G},\n\tDoi = {10.1016/0377-0257(84)80016-6},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tLanguage = {English},\n\tMonth = jan,\n\tNumber = {3},\n\tPages = {309--329},\n\tTitle = {Dynamics of rigid and flexible polymer chains in confined geometries. part {I}: steady simple shear flow},\n\tUrl = {http://linkinghub.elsevier.com/retrieve/pii/0377025784800166},\n\tVolume = {15},\n\tYear = {1984},\n\tBdsk-Url-1 = {http://linkinghub.elsevier.com/retrieve/pii/0377025784800166},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1016/0377-0257(84)80016-6}}\n\n
\n
\n\n\n
\n This paper describes exact solutions to the response of both the elastic and rigid dumbbell models to a steady simple shear flow in a channel having a length scale comparable to the dumbbells themselves. Results are given for rheological properties over the entire range of the ratio of the channel width to the length of the dumbbell. It is found that both models lead to a decrease in viscosity as the channel is reduced in size with the elastic dumbbell predicting a stronger dependence on that parameter compared with the rigid dumbbell. The elastic dumbbell predicts shear independent rheological properties whereas the rigid dumbell predicts shear thinning as in the case of unbounded flows. The rate of shear thinning, however, decreases with decreasing channel width.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The dynamics of dilute colloidal suspensions subject to time-dependent flow fields by conservative dichroism.\n \n \n \n \n\n\n \n Frattini, P L; and Fuller, G.\n\n\n \n\n\n\n Journal of Colloid and Interface Science, 100(2): 506–518. 1984.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{frattini_dynamics_1984,\n\tAbstract = {An apparatus is described which permits simultaneous determination of the conservative linear dichroism and the average orientation angle of a dilute suspension of optically anisotropic, nonabsorbing spheroids subjected to a transient shear field. The magnitude of the dichroism reflects the degree of alignment of the particles along their average orientation angle and measurement of this property in time directly obtains the rotational motion of the particles. This experiment makes use of a photoelastic modulator in a novel optical train; the first and second harmonics of the detector signal, which are independent functions of the dichroism and orientation angle, are measured simultaneously as functions of time. Preliminary data on a dilute suspension of bentonite in 90\\% glycerine and 10\\% water subjected to an instantaneous simple shear are presented to illustrate the technique. The qualitative features of these data are adequately described by the Rayleigh scattering theory coupled with Jeffery's solution for the dynamics of a rigid spheroid in a creeping simple shear. textcopyright 1984 Academic Press, Inc. All rights of reproduction in any form reserved.},\n\tAuthor = {Frattini, P L and Fuller, G.G.},\n\tJournal = {Journal of Colloid and Interface Science},\n\tNumber = {2},\n\tPages = {506--518},\n\tTitle = {The dynamics of dilute colloidal suspensions subject to time-dependent flow fields by conservative dichroism},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0021480161&partnerID=40&md5=6f90f1ae76d87dc7ebbab0d9afbb48f1},\n\tVolume = {100},\n\tYear = {1984},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0021480161&partnerID=40&md5=6f90f1ae76d87dc7ebbab0d9afbb48f1}}\n\n
\n
\n\n\n
\n An apparatus is described which permits simultaneous determination of the conservative linear dichroism and the average orientation angle of a dilute suspension of optically anisotropic, nonabsorbing spheroids subjected to a transient shear field. The magnitude of the dichroism reflects the degree of alignment of the particles along their average orientation angle and measurement of this property in time directly obtains the rotational motion of the particles. This experiment makes use of a photoelastic modulator in a novel optical train; the first and second harmonics of the detector signal, which are independent functions of the dichroism and orientation angle, are measured simultaneously as functions of time. Preliminary data on a dilute suspension of bentonite in 90% glycerine and 10% water subjected to an instantaneous simple shear are presented to illustrate the technique. The qualitative features of these data are adequately described by the Rayleigh scattering theory coupled with Jeffery's solution for the dynamics of a rigid spheroid in a creeping simple shear. textcopyright 1984 Academic Press, Inc. All rights of reproduction in any form reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Flow-Enhanced Desorption of Adsorbed Flexible Polymer Chains.\n \n \n \n \n\n\n \n G, F. G.; and JEN-JIANG, L E E\n\n\n \n\n\n\n In Polymer Adsorption and Dispersion Stability, pages 5–67. American Chemical Society, February 1984.\n \n\n\n\n
\n\n\n\n \n \n \"Flow-EnhancedPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@incollection{g_flow-enhanced_1984,\n\tAbstract = {We have applied ellipsometry to the study of polymer chain desorption in flowing systems. Using this technique both the polymer film thickness and adsorbance were measured under velocity gradients up to 7800 sec-1. Three molecular weight samples of polystyrene ranging from 4 to 20 million were studied using cyclohexane at the theta condition as the solvent. Chrome was used as the adsorbent. The results indicate very substantial flow enhancement of the desorption rates. Complete removal of the polymer at the highest velocity gradient was possible over a period of 2-3 hours. At the highest velocity gradient and for the higher molecular weights, the film thickness was observed to decrease. Several possible mechanisms for this decrease are proposed.},\n\tAuthor = {G, FULLER GERALD and JEN-JIANG, L E E},\n\tBooktitle = {Polymer {Adsorption} and {Dispersion} {Stability}},\n\tIsbn = {0-8412-0820-4},\n\tMonth = feb,\n\tPages = {5--67},\n\tPublisher = {American Chemical Society},\n\tTitle = {Flow-{Enhanced} {Desorption} of {Adsorbed} {Flexible} {Polymer} {Chains}},\n\tUrl = {http://dx.doi.org/10.1021/bk-1984-0240.ch005},\n\tYear = {1984},\n\tBdsk-Url-1 = {http://dx.doi.org/10.1021/bk-1984-0240.ch005}}\n\n
\n
\n\n\n
\n We have applied ellipsometry to the study of polymer chain desorption in flowing systems. Using this technique both the polymer film thickness and adsorbance were measured under velocity gradients up to 7800 sec-1. Three molecular weight samples of polystyrene ranging from 4 to 20 million were studied using cyclohexane at the theta condition as the solvent. Chrome was used as the adsorbent. The results indicate very substantial flow enhancement of the desorption rates. Complete removal of the polymer at the highest velocity gradient was possible over a period of 2-3 hours. At the highest velocity gradient and for the higher molecular weights, the film thickness was observed to decrease. Several possible mechanisms for this decrease are proposed.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Ellipsometry studies of adsorbed polymer chains subjected to flow.\n \n \n \n \n\n\n \n Lee, J J; and Fuller, G.\n\n\n \n\n\n\n Macromolecules, 17(3): 375–380. 1984.\n \n\n\n\n
\n\n\n\n \n \n \"EllipsometryPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{lee_ellipsometry_1984,\n\tAbstract = {Ellipsometry has been applied to study the average thickness of adsorbed polymer layers under flowing conditions. The measurement also produces a simultaneous determination of the surface concentration. Measurements have been conducted under flowing pure solvent at the \\$þeta\\$-temperature on four polystyrene samples of molecular weights ranging from 1.8 texttimes 106 to 20 texttimes 106 adsorbed previously onto chrome from quiescent cyclohexane solutions. No detectable changes in film thickness were observed for the three lower molecular weight samples over the whole range of velocity gradients applied. On the other hand, the film thickness of the 20 million molecular weight sample began to decrease as the apparent velocity gradient surpassed 2000 s-1 and decreased as much as 15\\% when the apparent velocity gradient reached 7800 s-1. The response of the adsorbed polymer layer to the imposition of the flow over short periods of time was reversible, and no change in adsorbance was observed between the inception and cessation of the flow. The measured conformational changes were also compared to the predictions of a simple bead-and-spring model of an attached polymer segment, and qualitative agreement was obtained when the effect of finite extensibility was included in the model. textcopyright 1984 American Chemical Society.},\n\tAuthor = {Lee, J J and Fuller, G.G.},\n\tJournal = {Macromolecules},\n\tNumber = {3},\n\tPages = {375--380},\n\tTitle = {Ellipsometry studies of adsorbed polymer chains subjected to flow},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000347373&partnerID=40&md5=36f58782fb47d07c8142b2aee525e2b8},\n\tVolume = {17},\n\tYear = {1984},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000347373&partnerID=40&md5=36f58782fb47d07c8142b2aee525e2b8}}\n\n
\n
\n\n\n
\n Ellipsometry has been applied to study the average thickness of adsorbed polymer layers under flowing conditions. The measurement also produces a simultaneous determination of the surface concentration. Measurements have been conducted under flowing pure solvent at the $þeta$-temperature on four polystyrene samples of molecular weights ranging from 1.8 texttimes 106 to 20 texttimes 106 adsorbed previously onto chrome from quiescent cyclohexane solutions. No detectable changes in film thickness were observed for the three lower molecular weight samples over the whole range of velocity gradients applied. On the other hand, the film thickness of the 20 million molecular weight sample began to decrease as the apparent velocity gradient surpassed 2000 s-1 and decreased as much as 15% when the apparent velocity gradient reached 7800 s-1. The response of the adsorbed polymer layer to the imposition of the flow over short periods of time was reversible, and no change in adsorbance was observed between the inception and cessation of the flow. The measured conformational changes were also compared to the predictions of a simple bead-and-spring model of an attached polymer segment, and qualitative agreement was obtained when the effect of finite extensibility was included in the model. textcopyright 1984 American Chemical Society.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1983\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Conformation and Dynamics of Adsorbed Polymer Molecules Subjected to Flow.\n \n \n \n \n\n\n \n Fuller, G. G\n\n\n \n\n\n\n Adhesion Aspects of Polymeric Coatings, (Chapter 13): 243–251. 1983.\n \n\n\n\n
\n\n\n\n \n \n \"ConformationPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_conformation_1983,\n\tAuthor = {Fuller, Gerald G},\n\tDoi = {10.1007/978-1-4613-3658-7_13},\n\tJournal = {Adhesion Aspects of Polymeric Coatings},\n\tNumber = {Chapter 13},\n\tPages = {243--251},\n\tTitle = {Conformation and {Dynamics} of {Adsorbed} {Polymer} {Molecules} {Subjected} to {Flow}},\n\tUrl = {http://link.springer.com/10.1007/978-1-4613-3658-7_13},\n\tYear = {1983},\n\tBdsk-Url-1 = {http://link.springer.com/10.1007/978-1-4613-3658-7_13},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1007/978-1-4613-3658-7_13}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Dynamics of adsorbed polymer chains subjected to flow: The dumbbell model.\n \n \n \n\n\n \n Fuller, G. G\n\n\n \n\n\n\n Journal of Polymer Science: Polymer Physics Edition, 21(1): 151–157. 1983.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_dynamics_1983,\n\tAuthor = {Fuller, Gerald G},\n\tJournal = {Journal of Polymer Science: Polymer Physics Edition},\n\tNumber = {1},\n\tPages = {151--157},\n\tTitle = {Dynamics of adsorbed polymer chains subjected to flow: {The} dumbbell model},\n\tVolume = {21},\n\tYear = {1983}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1981\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n The effects of conformation-dependent friction and internal viscosity on the dynamics of the nonlinear dumbbell model for a dilute polymer solution.\n \n \n \n \n\n\n \n Fuller, G.; and Leal, L G\n\n\n \n\n\n\n Journal of Non-Newtonian Fluid Mechanics, 8(3-4): 271–310. 1981.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_effects_1981,\n\tAbstract = {The effects of a nonlinear spring, variable hydrodynamic friction and internal viscosity are examined in detail for the dumbbell model of dilute polymer solutions in a general two-dimensional flow. Both steady and transient startup flows are investigated and the response of the bulk stress is calculated. The existence of a hysteresis in the steady end-to-end length of the dumbbell, which arises from the variable friction factor, is found to depend strongly on both the flow type (i.e. the ratio of vorticity to the rate of strain) and the moleculecular weight. The influence of internal viscosity is examined in detail using two methods of solution for the problems of transient and steady two-dimensional flows. The first method uses a perturbation expansion valid for small values of the internal viscosity and leads largely to analytical results. The second technique uses the pre-averaging approximation of Cerf and is capable of predicting the model response when the internal viscosity becomes large. The most striking result for large values of the internal viscosity and a linear spring dumbbell is the appearance of multiple steady states in the dumbbell length for flows between simple shear and pure straining, and the removal of the well known singularity in end-to-end length which otherwise appears in the linear dumbbell model at a critical value of the strain rate. textcopyright 1981.},\n\tAuthor = {Fuller, G.G. and Leal, L G},\n\tJournal = {Journal of Non-Newtonian Fluid Mechanics},\n\tNumber = {3-4},\n\tPages = {271--310},\n\tTitle = {The effects of conformation-dependent friction and internal viscosity on the dynamics of the nonlinear dumbbell model for a dilute polymer solution},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0019592088&partnerID=40&md5=613ff817e6ff1267f59fca9d70da6f13},\n\tVolume = {8},\n\tYear = {1981},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0019592088&partnerID=40&md5=613ff817e6ff1267f59fca9d70da6f13}}\n\n
\n
\n\n\n
\n The effects of a nonlinear spring, variable hydrodynamic friction and internal viscosity are examined in detail for the dumbbell model of dilute polymer solutions in a general two-dimensional flow. Both steady and transient startup flows are investigated and the response of the bulk stress is calculated. The existence of a hysteresis in the steady end-to-end length of the dumbbell, which arises from the variable friction factor, is found to depend strongly on both the flow type (i.e. the ratio of vorticity to the rate of strain) and the moleculecular weight. The influence of internal viscosity is examined in detail using two methods of solution for the problems of transient and steady two-dimensional flows. The first method uses a perturbation expansion valid for small values of the internal viscosity and leads largely to analytical results. The second technique uses the pre-averaging approximation of Cerf and is capable of predicting the model response when the internal viscosity becomes large. The most striking result for large values of the internal viscosity and a linear spring dumbbell is the appearance of multiple steady states in the dumbbell length for flows between simple shear and pure straining, and the removal of the well known singularity in end-to-end length which otherwise appears in the linear dumbbell model at a critical value of the strain rate. textcopyright 1981.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n NETWORK MODELS OF CONCENTRATED POLYMER SOLUTIONS DERIVED FROM THE YAMAMOTO NETWORK THEORY.\n \n \n \n \n\n\n \n Fuller, G.; and Leal, L G\n\n\n \n\n\n\n Journal of Polymer Science: Polymer Physics Edition, 19(4): 531–555. April 1981.\n \n\n\n\n
\n\n\n\n \n \n \"NETWORKPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_network_1981,\n\tAbstract = {In the reported study, several choices of the functions describing the creation and destruction processes of entanglement junctions in the Yamamoto network theory of concentrated polymer solutions have been examined. These choices are simple functions of the extension of the network segments bridging the entanglement points and it is demonstrated that the moments of the distribution function describing the network conformation can be solved for analytically. This has been done for a wide range of two-dimensional flows, both for the steady state and transient start-up and relaxation problems. The macroscopic stress tensor and flow birefringence are calculated and a variety of nonlinear effects are predicted and discussed.},\n\tAuthor = {Fuller, G.G. and Leal, L G},\n\tDoi = {10.1002/pol.1981.180190401},\n\tJournal = {Journal of Polymer Science: Polymer Physics Edition},\n\tMonth = apr,\n\tNumber = {4},\n\tPages = {531--555},\n\tTitle = {{NETWORK} {MODELS} {OF} {CONCENTRATED} {POLYMER} {SOLUTIONS} {DERIVED} {FROM} {THE} {YAMAMOTO} {NETWORK} {THEORY}.},\n\tUrl = {http://doi.wiley.com/10.1002/pol.1981.180190401 http://www.scopus.com/inward/record.url?eid=2-s2.0-0019558369&partnerID=40&md5=e8799f932ecbf01b12590b223665cd40 http://www.scopus.com/inward/record.url?eid=2-s2.0-0019557270&partnerID=40&md5=c2da5c5658cb9009025255a6935a716b http://doi.wiley.com/10.1002/pol.1981.180190402},\n\tVolume = {19},\n\tYear = {1981},\n\tBdsk-Url-1 = {http://doi.wiley.com/10.1002/pol.1981.180190401%20http://www.scopus.com/inward/record.url?eid=2-s2.0-0019558369&partnerID=40&md5=e8799f932ecbf01b12590b223665cd40%20http://www.scopus.com/inward/record.url?eid=2-s2.0-0019557270&partnerID=40&md5=c2da5c5658cb9009025255a6935a716b%20http://doi.wiley.com/10.1002/pol.1981.180190402},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1002/pol.1981.180190401}}\n\n
\n
\n\n\n
\n In the reported study, several choices of the functions describing the creation and destruction processes of entanglement junctions in the Yamamoto network theory of concentrated polymer solutions have been examined. These choices are simple functions of the extension of the network segments bridging the entanglement points and it is demonstrated that the moments of the distribution function describing the network conformation can be solved for analytically. This has been done for a wide range of two-dimensional flows, both for the steady state and transient start-up and relaxation problems. The macroscopic stress tensor and flow birefringence are calculated and a variety of nonlinear effects are predicted and discussed.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Flow birefringence of concentrated polymer solutions in two-dimensional flows.\n \n \n \n \n\n\n \n Fuller, G.; and Leal, L G\n\n\n \n\n\n\n Journal of Polymer Science: Polymer Physics Edition, 19(4): 557–587. April 1981.\n \n\n\n\n
\n\n\n\n \n \n \"FlowPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_flow_1981,\n\tAuthor = {Fuller, G.G. and Leal, L G},\n\tDoi = {10.1002/pol.1981.180190402},\n\tJournal = {Journal of Polymer Science: Polymer Physics Edition},\n\tMonth = apr,\n\tNumber = {4},\n\tPages = {557--587},\n\tTitle = {Flow birefringence of concentrated polymer solutions in two-dimensional flows},\n\tUrl = {http://doi.wiley.com/10.1002/pol.1981.180190402},\n\tVolume = {19},\n\tYear = {1981},\n\tBdsk-Url-1 = {http://doi.wiley.com/10.1002/pol.1981.180190402},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1002/pol.1981.180190402}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1980\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n The measurement of velocity gradients in laminar flow by homodyne light-scattering spectroscopy.\n \n \n \n \n\n\n \n Fuller, G.; Rallison, J M; Schmidt, R L; and Leal, L G\n\n\n \n\n\n\n Journal of Fluid Mechanics, 100(03): 555–575. 1980.\n \n\n\n\n
\n\n\n\n \n \n \"ThePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_measurement_1980,\n\tAbstract = {Abstract A technique for measuring velocity gradients in laminar flows by homodyne light scattering is presented. A theory which describes the light - scattering spectrum is derived that includes the effects of different types of linear flow fields, particle diffusion and the intensity ...},\n\tAuthor = {Fuller, G.G. and Rallison, J M and Schmidt, R L and Leal, L G},\n\tDoi = {10.1017/S0022112080001280},\n\tJournal = {Journal of Fluid Mechanics},\n\tNumber = {03},\n\tPages = {555--575},\n\tTitle = {The measurement of velocity gradients in laminar flow by homodyne light-scattering spectroscopy},\n\tUrl = {http://journals.cambridge.org/action/displayAbstract?aid=375517},\n\tVolume = {100},\n\tYear = {1980},\n\tBdsk-Url-1 = {http://journals.cambridge.org/action/displayAbstract?aid=375517},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1017/S0022112080001280}}\n\n
\n
\n\n\n
\n Abstract A technique for measuring velocity gradients in laminar flows by homodyne light scattering is presented. A theory which describes the light - scattering spectrum is derived that includes the effects of different types of linear flow fields, particle diffusion and the intensity ...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Flow birefringence of dilute polymer solutions in two-dimensional flows.\n \n \n \n \n\n\n \n Fuller, G.; and Leal, L G\n\n\n \n\n\n\n Rheologica Acta, 19(5): 580–600. 1980.\n \n\n\n\n
\n\n\n\n \n \n \"FlowPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_flow_1980,\n\tAbstract = {Flow birefringence measurements have been obtained on three molecular weight samples (2-8 texttimes 106MW, MW/MN = 1.14-1.3) of polystyrene in dilute solution (50-100 ppm) in a viscous polychorinated biphenyl solvent. The flows were generated using a four roll mill which could simulate a wide range of two dimensional flows in which the flow type (i.e. the ratio of the rate of rotation to the rate of strain) could be varied independently of the velocity gradient. The normalized birefringence, corrected for concentration, (Δn/nc), was observed to approach a saturation value at high velocity gradients in purely extensional flow. This saturation value was independent of both the molecular weight and the concentration c, in agreement with theory. In addition, the magnitude of the saturation value is consistent with nearly fully extended chains and suggests extensions in the range of 20-50 times the rest state size. The data for the birefringence over a wide range of flows was found to be well correlated against the eigenvalue of the velocity gradient tensor in agreement with the results of the "strong/weak" flow theories of Tanner (1976) and Olbricht et al. (1980). The experiments are compared with a simple dumbbell model which incorporates the effects of a nonlinear spring variable hydrodynamic friction, and internal viscosity. It is shown that this simple model can simulate the experimental results surprisingly well if the effects of molecular weight distribution and finite transit times in the flow device are taken into account. textcopyright 1980 Dr. Dietrich Steinkopff Verlag.},\n\tAuthor = {Fuller, G.G. and Leal, L G},\n\tJournal = {Rheologica Acta},\n\tNumber = {5},\n\tPages = {580--600},\n\tTitle = {Flow birefringence of dilute polymer solutions in two-dimensional flows},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000485304&partnerID=40&md5=0594fccdd3cb1415a4145f4a36c17f9a},\n\tVolume = {19},\n\tYear = {1980},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0000485304&partnerID=40&md5=0594fccdd3cb1415a4145f4a36c17f9a}}\n\n
\n
\n\n\n
\n Flow birefringence measurements have been obtained on three molecular weight samples (2-8 texttimes 106MW, MW/MN = 1.14-1.3) of polystyrene in dilute solution (50-100 ppm) in a viscous polychorinated biphenyl solvent. The flows were generated using a four roll mill which could simulate a wide range of two dimensional flows in which the flow type (i.e. the ratio of the rate of rotation to the rate of strain) could be varied independently of the velocity gradient. The normalized birefringence, corrected for concentration, (Δn/nc), was observed to approach a saturation value at high velocity gradients in purely extensional flow. This saturation value was independent of both the molecular weight and the concentration c, in agreement with theory. In addition, the magnitude of the saturation value is consistent with nearly fully extended chains and suggests extensions in the range of 20-50 times the rest state size. The data for the birefringence over a wide range of flows was found to be well correlated against the eigenvalue of the velocity gradient tensor in agreement with the results of the \"strong/weak\" flow theories of Tanner (1976) and Olbricht et al. (1980). The experiments are compared with a simple dumbbell model which incorporates the effects of a nonlinear spring variable hydrodynamic friction, and internal viscosity. It is shown that this simple model can simulate the experimental results surprisingly well if the effects of molecular weight distribution and finite transit times in the flow device are taken into account. textcopyright 1980 Dr. Dietrich Steinkopff Verlag.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Effect of Molecular Weight and Flow Type on Flow Birefringence of Dilute Polymer Solutions.\n \n \n \n \n\n\n \n Fuller, G. G; and Leal, L G.\n\n\n \n\n\n\n Rheology, (Chapter 74): 393–398. 1980.\n \n\n\n\n
\n\n\n\n \n \n \"EffectPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_effect_1980,\n\tAuthor = {Fuller, Gerald G and Leal, L Gary},\n\tDoi = {10.1007/978-1-4684-3743-0_74},\n\tJournal = {Rheology},\n\tNumber = {Chapter 74},\n\tPages = {393--398},\n\tTitle = {Effect of {Molecular} {Weight} and {Flow} {Type} on {Flow} {Birefringence} of {Dilute} {Polymer} {Solutions}},\n\tUrl = {http://link.springer.com/10.1007/978-1-4684-3743-0_74},\n\tYear = {1980},\n\tBdsk-Url-1 = {http://link.springer.com/10.1007/978-1-4684-3743-0_74},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1007/978-1-4684-3743-0_74}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Studies of the flow-induced stretching of a macromolecule in a dilute solution.\n \n \n \n\n\n \n Leal, L G; Fuller, G.; and Olbricht, W L\n\n\n \n\n\n\n Viscous Drag Reduction, 72. 1980.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{leal_studies_1980,\n\tAuthor = {Leal, L G and Fuller, G.G. and Olbricht, W L},\n\tJournal = {Viscous Drag Reduction},\n\tTitle = {Studies of the flow-induced stretching of a macromolecule in a dilute solution},\n\tVolume = {72},\n\tYear = {1980}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1976\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Applications of field theoretical methods to the calculation of infrared band shapes of molecules in strongly interacting solvents. I.\n \n \n \n \n\n\n \n Paul, R; and Fuller, G.\n\n\n \n\n\n\n Journal of Chemical Physics, 64(9): 3809–3825. 1976.\n \n\n\n\n
\n\n\n\n \n \n \"ApplicationsPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{paul_applications_1976,\n\tAbstract = {Starting from a finite temperature Green's function formalism we have developed a diagrammatic method for calculating the complex dielectric constant of a system consisting of diatomic molecules that are strongly interacting with the solvent medium. Perturbative treatments suitable for weak interactions have been replaced by a variational method, and diagrams picked up are summed up to all orders in the solvent-solute interaction potential. A method is presented whereby all the vectorial integrations can be carried out exactly leaving a final quadrature which must be done numerically. It is pointed out that the internal states of the diatomic molecule are coupled to each other via a very complicated cooperative effect induced by the solvent.},\n\tAuthor = {Paul, R and Fuller, G.G.},\n\tDoi = {10.1063/1.432699},\n\tJournal = {Journal of Chemical Physics},\n\tNumber = {9},\n\tPages = {3809--3825},\n\tTitle = {Applications of field theoretical methods to the calculation of infrared band shapes of molecules in strongly interacting solvents. {I}.},\n\tUrl = {http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=1976JChPh..64.3809P&link_type=EJOURNAL},\n\tVolume = {64},\n\tYear = {1976},\n\tBdsk-Url-1 = {http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=1976JChPh..64.3809P&link_type=EJOURNAL},\n\tBdsk-Url-2 = {http://dx.doi.org/10.1063/1.432699}}\n\n
\n
\n\n\n
\n Starting from a finite temperature Green's function formalism we have developed a diagrammatic method for calculating the complex dielectric constant of a system consisting of diatomic molecules that are strongly interacting with the solvent medium. Perturbative treatments suitable for weak interactions have been replaced by a variational method, and diagrams picked up are summed up to all orders in the solvent-solute interaction potential. A method is presented whereby all the vectorial integrations can be carried out exactly leaving a final quadrature which must be done numerically. It is pointed out that the internal states of the diatomic molecule are coupled to each other via a very complicated cooperative effect induced by the solvent.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n A modified Redlich-Kwong-Soave equation of state capable of representing the liquid state.\n \n \n \n \n\n\n \n Fuller, G.\n\n\n \n\n\n\n Ind. Eng. Chem. Fund., 15(4): 254–257. 1976.\n \n\n\n\n
\n\n\n\n \n \n \"APaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{fuller_modified_1976,\n\tAbstract = {The modified form of the Redlich-Kwong equation of statees with greater accuracy. The additional modifications contain two features; first, the equation of state now leads to a variable critical co proposed by Soave is further generalized in order to reproduce saturated liquid volumes and compressed liquid volumes of pure substancmpressibility factor, and second, a new universal temperature function is incorporated in the equation making both the a and b parameters functions of temperature. The added modifications preserve the ability of the equation to predict vapor pressures with only the critical volume and parachor being required as additional data input. The results of calculations indicate that the proposed equation is capable of describing even polar molecules such as water and ammonia with reasonable accuracy.},\n\tAuthor = {Fuller, G.G.},\n\tJournal = {Ind. Eng. Chem. Fund.},\n\tNumber = {4},\n\tPages = {254--257},\n\tTitle = {A modified {Redlich}-{Kwong}-{Soave} equation of state capable of representing the liquid state},\n\tUrl = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0017015508&partnerID=40&md5=6eb740246d02845b1b97fd48a49423b0},\n\tVolume = {15},\n\tYear = {1976},\n\tBdsk-Url-1 = {http://www.scopus.com/inward/record.url?eid=2-s2.0-0017015508&partnerID=40&md5=6eb740246d02845b1b97fd48a49423b0}}\n\n
\n
\n\n\n
\n The modified form of the Redlich-Kwong equation of statees with greater accuracy. The additional modifications contain two features; first, the equation of state now leads to a variable critical co proposed by Soave is further generalized in order to reproduce saturated liquid volumes and compressed liquid volumes of pure substancmpressibility factor, and second, a new universal temperature function is incorporated in the equation making both the a and b parameters functions of temperature. The added modifications preserve the ability of the equation to predict vapor pressures with only the critical volume and parachor being required as additional data input. The results of calculations indicate that the proposed equation is capable of describing even polar molecules such as water and ammonia with reasonable accuracy.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n undefined\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Physicochemical characteristics of droplet interface bilayers.\n \n \n \n\n\n \n \n\n\n \n\n\n\n . .\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Surface energy and separation mechanics of droplet interface phospholipid bilayers.\n \n \n \n\n\n \n Huang, Y.; Chandran Suja, V.; Tajuelo, J.; and Fuller, G. G\n\n\n \n\n\n\n Journal of the Royal Society Interface, 18(175): 20200860. .\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{huang2021surface,\n  title={Surface energy and separation mechanics of droplet interface phospholipid bilayers},\n  author={Huang, Yaoqi and Chandran Suja, Vineeth and Tajuelo, Javier and Fuller, Gerald G},\n  journal={Journal of the Royal Society Interface},\n  volume={18},\n  number={175},\n  pages={20200860},\n  publisher={The Royal Society}\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Interfacial rheology and structure of straight-chain and branched hexadecanol mixtures.\n \n \n \n\n\n \n Gavranovic, G. T; Kurtz, R. E; Golemanov, K.; Lange, A.; and Fuller, G. G\n\n\n \n\n\n\n Industrial \\p̌hantom\\&. .\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{gavranovic_interfacial_????,\n\tAuthor = {Gavranovic, Grant T and Kurtz, Rachel E and Golemanov, Konstantin and Lange, Arno and Fuller, Gerald G},\n\tJournal = {Industrial \\{\\vphantom{\\}}\\&},\n\tTitle = {Interfacial rheology and structure of straight-chain and branched hexadecanol mixtures}}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Structure and dynamics of liquid crystalline droplets suspended in polymer liquids.\n \n \n \n\n\n \n Meyer, E. L; Fuller, G. G; and Reamey, R. H\n\n\n \n\n\n\n In IS\\p̌hantom\\&, . \n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@inproceedings{meyer_structure_????,\n\tAuthor = {Meyer, Eleanor L and Fuller, Gerald G and Reamey, Robert H},\n\tBooktitle = {{IS}\\{\\vphantom{\\}}\\&},\n\tTitle = {Structure and dynamics of liquid crystalline droplets suspended in polymer liquids},\n\tBdsk-File-1 = {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}}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n\n\n\n
\n\n\n \n\n \n \n \n \n\n
\n"}; document.write(bibbase_data.data);