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/service/mendeley/5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2/group/cd514115-949a-34c4-b220-b5e9bedcc2c8?jsonp=1&jsonp=1\"></script>\n \n
\n\n PHP\n
\n \n <?php\n $contents = file_get_contents(\"https://bibbase.org/service/mendeley/5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2/group/cd514115-949a-34c4-b220-b5e9bedcc2c8?jsonp=1\");\n print_r($contents);\n ?>\n \n
\n\n iFrame\n (not recommended)\n
\n \n <iframe src=\"https://bibbase.org/service/mendeley/5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2/group/cd514115-949a-34c4-b220-b5e9bedcc2c8?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 2021\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Evaluation of Fast Mobility Particle Sizer (FMPS) for Ambient Aerosol Measurement.\n \n \n \n \n\n\n \n Lin, Y.; Pham, L.; Wang, X.; Bahreini, R.; and Jung, H., S.\n\n\n \n\n\n\n Aerosol and Air Quality Research, 21. 2021.\n \n\n\n\n
\n\n\n\n \n \n \"EvaluationPaper\n  \n \n \n \"EvaluationWebsite\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 14 downloads\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
@article{\n title = {Evaluation of Fast Mobility Particle Sizer (FMPS) for Ambient Aerosol Measurement},\n type = {article},\n year = {2021},\n keywords = {Ambient particles,Fast Mobility Particle Sizer,Fractal dimension,Inversion matrix,Particle size distribution},\n volume = {21},\n websites = {https://aaqr.org/articles/aaqr-20-08-tn-0525},\n publisher = {Taiwan Association for Aerosol Research},\n id = {0e9503c4-235a-3500-b35a-1683b31245ba},\n created = {2021-02-04T16:00:30.771Z},\n accessed = {2021-02-04},\n file_attached = {true},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2021-02-04T16:00:30.865Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {13 Fast Mobility Particle Sizer (FMPS) size distribution measurements with different inversion matrices 14 were compared with the Scanning Mobility Particle Sizer (SMPS) for ambient aerosols sampled from a 15 background location in Riverside, CA in this study. The FMPS-compact matrix showed the best agreement},\n bibtype = {article},\n author = {Lin, Yue and Pham, Liem and Wang, Xiaoliang and Bahreini, Roya and Jung, Heejung S.},\n doi = {10.4209/aaqr.200525},\n journal = {Aerosol and Air Quality Research}\n}
\n
\n\n\n
\n 13 Fast Mobility Particle Sizer (FMPS) size distribution measurements with different inversion matrices 14 were compared with the Scanning Mobility Particle Sizer (SMPS) for ambient aerosols sampled from a 15 background location in Riverside, CA in this study. The FMPS-compact matrix showed the best agreement\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Variability in the mass absorption cross section of black carbon (BC) aerosols is driven by BC internal mixing state at a central European background site (Melpitz, Germany) in winter.\n \n \n \n\n\n \n Yuan, J.; Lewis Modini, R.; Zanatta, M.; Herber, A., B.; Müller, T.; Wehner, B.; Poulain, L.; Tuch, T.; Baltensperger, U.; and Gysel-Beer, M.\n\n\n \n\n\n\n Atmospheric Chemistry and Physics, 21(2): 635-655. 1 2021.\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 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{\n title = {Variability in the mass absorption cross section of black carbon (BC) aerosols is driven by BC internal mixing state at a central European background site (Melpitz, Germany) in winter},\n type = {article},\n year = {2021},\n pages = {635-655},\n volume = {21},\n month = {1},\n publisher = {Copernicus GmbH},\n day = {18},\n id = {74986f51-d3e9-312d-a473-66a5cf220c5f},\n created = {2021-02-08T17:23:49.762Z},\n accessed = {2021-02-08},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2021-02-08T17:23:49.871Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Properties of atmospheric black carbon (BC) particles were characterized during a field experiment at a rural background site (Melpitz, Germany) in February 2017. BC absorption at a wavelength of 870 nm was measured by a photoacoustic extinctiometer, and BC physical properties (BC mass concentration, core size distribution and coating thickness) were measured by a single-particle soot photometer (SP2). Additionally, a catalytic stripper was used to intermittently remove BC coatings by alternating between ambient and thermo-denuded conditions. From these data the mass absorption cross section of BC (MACBC) and its enhancement factor (EMAC) were inferred for essentially waterfree aerosol as present after drying to low relative humidity (RH). Two methods were applied independently to investigate the coating effect on EMAC: A correlation method (MACBC; ambient vs. BC coating thickness) and a denuding method (MACBC; ambient vs. MACBC; denuded). Observed EMAC values varied from 1.0 to 1.6 (lower limit from denuding method) or 1:2 to 1.9 (higher limit from correlation method), with the mean coating volume fraction ranging from 54% to 78% in the dominating mass equivalent BC core diameter range of 200?220 nm.MACBC and EMAC were strongly correlated with coating thickness of BC. By contrast, other potential drivers of EMAC variability, such as different BC sources (air mass origin and absorption Angström exponent), coating composition (ratio of inorganics to organics) and BC core size distribution, had only minor effects. These results for ambient BC measured at Melpitz during winter show that the lensing effect caused by coatings on BC is the main driver of the variations in MACBC and EMAC, while changes in other BC particle properties such as source, BC core size or coating composition play only minor roles at this rural background site with a large fraction of aged particles. Indirect evidence suggests that potential dampening of the lensing effect due to unfavorable morphology was most likely small or even negligible.},\n bibtype = {article},\n author = {Yuan, Jinfeng and Lewis Modini, Robin and Zanatta, Marco and Herber, Andreas B. and Müller, Thomas and Wehner, Birgit and Poulain, Laurent and Tuch, Thomas and Baltensperger, Urs and Gysel-Beer, Martin},\n doi = {10.5194/acp-21-635-2021},\n journal = {Atmospheric Chemistry and Physics},\n number = {2}\n}
\n
\n\n\n
\n Properties of atmospheric black carbon (BC) particles were characterized during a field experiment at a rural background site (Melpitz, Germany) in February 2017. BC absorption at a wavelength of 870 nm was measured by a photoacoustic extinctiometer, and BC physical properties (BC mass concentration, core size distribution and coating thickness) were measured by a single-particle soot photometer (SP2). Additionally, a catalytic stripper was used to intermittently remove BC coatings by alternating between ambient and thermo-denuded conditions. From these data the mass absorption cross section of BC (MACBC) and its enhancement factor (EMAC) were inferred for essentially waterfree aerosol as present after drying to low relative humidity (RH). Two methods were applied independently to investigate the coating effect on EMAC: A correlation method (MACBC; ambient vs. BC coating thickness) and a denuding method (MACBC; ambient vs. MACBC; denuded). Observed EMAC values varied from 1.0 to 1.6 (lower limit from denuding method) or 1:2 to 1.9 (higher limit from correlation method), with the mean coating volume fraction ranging from 54% to 78% in the dominating mass equivalent BC core diameter range of 200?220 nm.MACBC and EMAC were strongly correlated with coating thickness of BC. By contrast, other potential drivers of EMAC variability, such as different BC sources (air mass origin and absorption Angström exponent), coating composition (ratio of inorganics to organics) and BC core size distribution, had only minor effects. These results for ambient BC measured at Melpitz during winter show that the lensing effect caused by coatings on BC is the main driver of the variations in MACBC and EMAC, while changes in other BC particle properties such as source, BC core size or coating composition play only minor roles at this rural background site with a large fraction of aged particles. Indirect evidence suggests that potential dampening of the lensing effect due to unfavorable morphology was most likely small or even negligible.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2020\n \n \n (9)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Non-Volatile Particle Number Emission Measurements with Catalytic Strippers: A Review.\n \n \n \n \n\n\n \n Giechaskiel, B.; Melas, A., D.; Lähde, T.; and Martini, G.\n\n\n \n\n\n\n Vehicles, 2(2): 342-364. 6 2020.\n \n\n\n\n
\n\n\n\n \n \n \"Non-VolatileWebsite\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 7 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{\n title = {Non-Volatile Particle Number Emission Measurements with Catalytic Strippers: A Review},\n type = {article},\n year = {2020},\n pages = {342-364},\n volume = {2},\n websites = {https://www.mdpi.com/2624-8921/2/2/19},\n month = {6},\n day = {24},\n id = {e6a1fb67-00d1-3216-832b-537c69806310},\n created = {2020-07-07T16:17:20.597Z},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-05T10:30:50.312Z},\n read = {true},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {Vehicle regulations include limits for non-volatile particle number emissions with sizes larger than 23 nm. The measurements are conducted with systems that remove the volatile particles by means of dilution and heating. Recently, the option of measuring from 10 nm was included in the Global Technical Regulation (GTR 15) as an additional option to the current >23 nm methodology. In order to avoid artefacts, i.e., measuring volatile particles that have nucleated downstream of the evaporation tube, a heated oxidation catalyst (i.e., catalytic stripper) is required. This review summarizes the studies with laboratory aerosols that assessed the volatile removal efficiency of evaporation tube and catalytic stripper-based systems using hydrocarbons, sulfuric acid, mixture of them, and ammonium sulfate. Special emphasis was given to distinguish between artefacts that happened in the 10–23 nm range or below. Furthermore, studies with vehicles’ aerosols that reported artefacts were collected to estimate critical concentration levels of volatiles. Maximum expected levels of volatiles for mopeds, motorcycles, light-duty and heavy-duty vehicles were also summarized. Both laboratory and vehicle studies confirmed the superiority of catalytic strippers in avoiding artefacts. Open issues that need attention are the sulfur storage capacity and the standardization of technical requirements for catalytic strippers.},\n bibtype = {article},\n author = {Giechaskiel, Barouch and Melas, Anastasios D. and Lähde, Tero and Martini, Giorgio},\n doi = {10.3390/vehicles2020019},\n journal = {Vehicles},\n number = {2}\n}
\n
\n\n\n
\n Vehicle regulations include limits for non-volatile particle number emissions with sizes larger than 23 nm. The measurements are conducted with systems that remove the volatile particles by means of dilution and heating. Recently, the option of measuring from 10 nm was included in the Global Technical Regulation (GTR 15) as an additional option to the current >23 nm methodology. In order to avoid artefacts, i.e., measuring volatile particles that have nucleated downstream of the evaporation tube, a heated oxidation catalyst (i.e., catalytic stripper) is required. This review summarizes the studies with laboratory aerosols that assessed the volatile removal efficiency of evaporation tube and catalytic stripper-based systems using hydrocarbons, sulfuric acid, mixture of them, and ammonium sulfate. Special emphasis was given to distinguish between artefacts that happened in the 10–23 nm range or below. Furthermore, studies with vehicles’ aerosols that reported artefacts were collected to estimate critical concentration levels of volatiles. Maximum expected levels of volatiles for mopeds, motorcycles, light-duty and heavy-duty vehicles were also summarized. Both laboratory and vehicle studies confirmed the superiority of catalytic strippers in avoiding artefacts. Open issues that need attention are the sulfur storage capacity and the standardization of technical requirements for catalytic strippers.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Multi-technique physico-chemical characterization of particles generated by a gasoline engine: Towards measuring tailpipe emissions below 23 nm.\n \n \n \n \n\n\n \n Focsa, C.; Duca, D.; Noble, J.; Vojkovic, M.; Carpentier, Y.; Pirim, C.; Betrancourt, C.; Desgroux, P.; Tritscher, T.; Spielvogel, J.; Rahman, M.; Boies, A.; Lee, K.; Bhave, A.; Legendre, S.; Lancry, O.; Kreutziger, P.; and Rieker, M.\n\n\n \n\n\n\n Atmospheric Environment, 235: 117642. 8 2020.\n \n\n\n\n
\n\n\n\n \n \n \"Multi-techniquePaper\n  \n \n \n \"Multi-techniqueWebsite\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 11 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{\n title = {Multi-technique physico-chemical characterization of particles generated by a gasoline engine: Towards measuring tailpipe emissions below 23 nm},\n type = {article},\n year = {2020},\n pages = {117642},\n volume = {235},\n websites = {https://linkinghub.elsevier.com/retrieve/pii/S1352231020303733},\n month = {8},\n publisher = {Pergamon},\n day = {15},\n id = {060d3544-2810-38e2-bbfc-a4fb4b26a244},\n created = {2020-07-21T15:51:50.541Z},\n accessed = {2020-06-08},\n file_attached = {true},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-21T15:51:50.653Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n bibtype = {article},\n author = {Focsa, C. and Duca, D. and Noble, J.A. and Vojkovic, M. and Carpentier, Y. and Pirim, C. and Betrancourt, C. and Desgroux, P. and Tritscher, T. and Spielvogel, J. and Rahman, M. and Boies, A. and Lee, K.F. and Bhave, A.N. and Legendre, S. and Lancry, O. and Kreutziger, P. and Rieker, M.},\n doi = {10.1016/j.atmosenv.2020.117642},\n journal = {Atmospheric Environment}\n}
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Comprehensive characterization of mainstream marijuana and tobacco smoke.\n \n \n \n\n\n \n Graves, B., M.; Johnson, T., J.; Nishida, R., T.; Dias, R., P.; Harynuk, J., J.; Kazemimanesh, M.; Olfert, J., S.; and Boies, A., M.\n\n\n \n\n\n\n Scientific Reports, 10(1): 1-15. 4 2020.\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\n\n\n
\n
@article{\n title = {Comprehensive characterization of mainstream marijuana and tobacco smoke},\n type = {article},\n year = {2020},\n keywords = {Mass spectrometry,Nanoparticles},\n pages = {1-15},\n volume = {10},\n month = {4},\n publisher = {Nature Publishing Group},\n day = {28},\n id = {71e8de1b-cc73-3f39-b7a1-dc65009a5db2},\n created = {2020-07-21T15:55:40.784Z},\n accessed = {2020-04-28},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-06T21:55:26.816Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {Recent increases in marijuana use and legalization without adequate knowledge of the risks necessitate the characterization of the billions of nanoparticles contained in each puff of smoke. Tobacco smoke offers a benchmark given that it has been extensively studied. Tobacco and marijuana smoke particles are quantitatively similar in volatility, shape, density and number concentration, albeit with differences in size, total mass and chemical composition. Particles from marijuana smoke are on average 29% larger in mobility diameter than particles from tobacco smoke and contain 3.4× more total mass. New measurements of semi-volatile fractions determine over 97% of the mass and volume of the particles from either smoke source are comprised of semi-volatile compounds. For tobacco and marijuana smoke, respectively, 4350 and 2575 different compounds are detected, of which, 670 and 536 (231 in common) are tentatively identified, and of these, 173 and 110 different compounds (69 in common) are known to cause negative health effects through carcinogenic, mutagenic, teratogenic, or other toxic mechanisms. This study demonstrates striking similarities between marijuana and tobacco smoke in terms of their physical and chemical properties.},\n bibtype = {article},\n author = {Graves, Brian M and Johnson, Tyler J and Nishida, Robert T and Dias, Ryan P and Harynuk, James J and Kazemimanesh, Mohsen and Olfert, Jason S and Boies, Adam M},\n doi = {10.1038/s41598-020-63120-6},\n journal = {Scientific Reports},\n number = {1}\n}
\n
\n\n\n
\n Recent increases in marijuana use and legalization without adequate knowledge of the risks necessitate the characterization of the billions of nanoparticles contained in each puff of smoke. Tobacco smoke offers a benchmark given that it has been extensively studied. Tobacco and marijuana smoke particles are quantitatively similar in volatility, shape, density and number concentration, albeit with differences in size, total mass and chemical composition. Particles from marijuana smoke are on average 29% larger in mobility diameter than particles from tobacco smoke and contain 3.4× more total mass. New measurements of semi-volatile fractions determine over 97% of the mass and volume of the particles from either smoke source are comprised of semi-volatile compounds. For tobacco and marijuana smoke, respectively, 4350 and 2575 different compounds are detected, of which, 670 and 536 (231 in common) are tentatively identified, and of these, 173 and 110 different compounds (69 in common) are known to cause negative health effects through carcinogenic, mutagenic, teratogenic, or other toxic mechanisms. This study demonstrates striking similarities between marijuana and tobacco smoke in terms of their physical and chemical properties.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Characterization of particulate matter emitted by a marine engine operated with liquefied natural gas and diesel fuels.\n \n \n \n \n\n\n \n Corbin, J., C.; Peng, W.; Yang, J.; Sommer, D., E.; Trivanovic, U.; Kirchen, P.; Miller, J., W.; Rogak, S.; Cocker, D., R.; Smallwood, G., J.; Lobo, P.; and Gagné, S.\n\n\n \n\n\n\n Atmospheric Environment, 220: 117030. 1 2020.\n \n\n\n\n
\n\n\n\n \n \n \"CharacterizationWebsite\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 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{\n title = {Characterization of particulate matter emitted by a marine engine operated with liquefied natural gas and diesel fuels},\n type = {article},\n year = {2020},\n pages = {117030},\n volume = {220},\n websites = {https://www.sciencedirect.com/science/article/pii/S1352231019306697?via%3Dihub},\n month = {1},\n publisher = {Pergamon},\n day = {1},\n id = {283a1af7-b3c2-30af-a530-385fed0002ec},\n created = {2020-08-19T14:30:43.527Z},\n accessed = {2020-08-19},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-19T14:30:43.527Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Liquefied natural gas (LNG) is becoming increasingly popular as a marine fuel as emission regulations become more stringent. However, very little data are available on the particulate matter (PM) emissions of modern marine natural gas engines. In this study, we present a first detailed characterization of the composition of the PM emitted by a modern, in-use, natural-gas-powered vessel. The vessel engines use compression-ignition and only a small amount of diesel fuel as pilot. These engines drive electrical generators, providing propulsion as well as auxiliary power for the vessel. Our emissions characterization includes six different techniques to measure black carbon (BC), including all methods determined as appropriate for measuring BC emissions from ships by the International Maritime Organization, as well as particle size distributions, metal concentrations, and organic particulate emissions. PM emissions differed significantly between idle and at-sea operating conditions. At idle, PM emission factors were primarily organic (approximately 1500mg/kWh), with BC emission factors over two orders of magnitude lower (5.6±0.4mg/kWh). At engine loads above 25%, all emissions were independent of load and substantially lower than at idle, at 4.4±1.7mg/kWh for organics and 0.8±0.2mg/kWh for black carbon. When operated only on diesel fuel, this engine emitted 8-fold more organic PM (38±15mg/kWh) and 37-fold more BC (30±11mg/kWh) at loads above 25%. At idle loads, the diesel-fuel emissions were comparable to the natural-gas emissions. In addition to organics and BC, a third category of non-volatile sub-10-nm particles was identified. A detailed consideration of our measurements indicated that the sources of the organic, BC, and sub-10-nm particles were lubrication oil, diesel pilot fuel, and lubrication-oil metals, respectively. Future studies should seek to quantify the emissions of other dual-fuel engines that will be entering the market.},\n bibtype = {article},\n author = {Corbin, Joel C. and Peng, Weihan and Yang, Jiacheng and Sommer, David E. and Trivanovic, Una and Kirchen, Patrick and Miller, J. Wayne and Rogak, Steven and Cocker, David R. and Smallwood, Gregory J. and Lobo, Prem and Gagné, Stéphanie},\n doi = {10.1016/J.ATMOSENV.2019.117030},\n journal = {Atmospheric Environment}\n}
\n
\n\n\n
\n Liquefied natural gas (LNG) is becoming increasingly popular as a marine fuel as emission regulations become more stringent. However, very little data are available on the particulate matter (PM) emissions of modern marine natural gas engines. In this study, we present a first detailed characterization of the composition of the PM emitted by a modern, in-use, natural-gas-powered vessel. The vessel engines use compression-ignition and only a small amount of diesel fuel as pilot. These engines drive electrical generators, providing propulsion as well as auxiliary power for the vessel. Our emissions characterization includes six different techniques to measure black carbon (BC), including all methods determined as appropriate for measuring BC emissions from ships by the International Maritime Organization, as well as particle size distributions, metal concentrations, and organic particulate emissions. PM emissions differed significantly between idle and at-sea operating conditions. At idle, PM emission factors were primarily organic (approximately 1500mg/kWh), with BC emission factors over two orders of magnitude lower (5.6±0.4mg/kWh). At engine loads above 25%, all emissions were independent of load and substantially lower than at idle, at 4.4±1.7mg/kWh for organics and 0.8±0.2mg/kWh for black carbon. When operated only on diesel fuel, this engine emitted 8-fold more organic PM (38±15mg/kWh) and 37-fold more BC (30±11mg/kWh) at loads above 25%. At idle loads, the diesel-fuel emissions were comparable to the natural-gas emissions. In addition to organics and BC, a third category of non-volatile sub-10-nm particles was identified. A detailed consideration of our measurements indicated that the sources of the organic, BC, and sub-10-nm particles were lubrication oil, diesel pilot fuel, and lubrication-oil metals, respectively. Future studies should seek to quantify the emissions of other dual-fuel engines that will be entering the market.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Comparison of standardized sampling and measurement reference systems for aircraft engine non-volatile particulate matter emissions.\n \n \n \n \n\n\n \n Lobo, P.; Durdina, L.; Brem, B., T.; Crayford, A., P.; Johnson, M., P.; Smallwood, G., J.; Siegerist, F.; Williams, P., I.; Black, E., A.; Llamedo, A.; Thomson, K., A.; Trueblood, M., B.; Yu, Z.; Hagen, D., E.; Whitefield, P., D.; Miake-Lye, R., C.; and Rindlisbacher, T.\n\n\n \n\n\n\n Journal of Aerosol Science, 145: 105557. 7 2020.\n \n\n\n\n
\n\n\n\n \n \n \"ComparisonWebsite\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 2 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{\n title = {Comparison of standardized sampling and measurement reference systems for aircraft engine non-volatile particulate matter emissions},\n type = {article},\n year = {2020},\n pages = {105557},\n volume = {145},\n websites = {https://www.sciencedirect.com/science/article/abs/pii/S002185022030046X?via%3Dihub},\n month = {7},\n publisher = {Pergamon},\n day = {1},\n id = {0185f142-2548-380f-9fa4-00deb112641d},\n created = {2020-08-19T14:32:08.940Z},\n accessed = {2020-08-19},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-19T14:32:08.940Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {The International Civil Aviation Organization has established new regulatory standards for emissions certification of non-volatile particulate matter (nvPM) from aircraft turbine engines. The adoption of the nvPM emissions regulatory standards required development of a standardized sampling and measurement methodology, and rigorous testing. Three reference systems for aircraft engine nvPM emissions measurement, compliant with the specifications for the standardized methodology, were independently developed. This paper reports the results of the first inter-comparison of these three reference systems using a CFM56-7B26/3 aircraft engine to establish repeatability and intermediate precision of the sampling and measurement systems as part of the multi-agency international collaborative projects: Aviation-Particle Regulatory Instrumentation Demonstration Experiment (A-PRIDE) 5/Studying, sAmpling and Measuring of aircraft ParticuLate Emissions (SAMPLE) III - SC03. The instruments used in the three reference systems recorded nvPM mass and number concentration, which were converted to their respective emission indices for comparison. The reference systems generally agreed to within 15% of the average nvPM number emission index and 30% of the average nvPM mass emission index. The only exception was for the nvPM mass instruments, which exhibited a higher variation as the concentration levels approached the limit of detection. The additional measured particle size distributions could be approximated to lognormal distributions with the geometric mean diameter ranging from 15 nm to 38 nm, and the geometric standard deviation varying between 1.53 and 1.92. The results from this study are a benchmark for the variability in standardized sampling and measurement systems for measuring aircraft engine nvPM emissions.},\n bibtype = {article},\n author = {Lobo, Prem and Durdina, Lukas and Brem, Benjamin T. and Crayford, Andrew P. and Johnson, Mark P. and Smallwood, Greg J. and Siegerist, Frithjof and Williams, Paul I. and Black, Elizabeth A. and Llamedo, Andrea and Thomson, Kevin A. and Trueblood, Max B. and Yu, Zhenhong and Hagen, Donald E. and Whitefield, Philip D. and Miake-Lye, Richard C. and Rindlisbacher, Theo},\n doi = {10.1016/J.JAEROSCI.2020.105557},\n journal = {Journal of Aerosol Science}\n}
\n
\n\n\n
\n The International Civil Aviation Organization has established new regulatory standards for emissions certification of non-volatile particulate matter (nvPM) from aircraft turbine engines. The adoption of the nvPM emissions regulatory standards required development of a standardized sampling and measurement methodology, and rigorous testing. Three reference systems for aircraft engine nvPM emissions measurement, compliant with the specifications for the standardized methodology, were independently developed. This paper reports the results of the first inter-comparison of these three reference systems using a CFM56-7B26/3 aircraft engine to establish repeatability and intermediate precision of the sampling and measurement systems as part of the multi-agency international collaborative projects: Aviation-Particle Regulatory Instrumentation Demonstration Experiment (A-PRIDE) 5/Studying, sAmpling and Measuring of aircraft ParticuLate Emissions (SAMPLE) III - SC03. The instruments used in the three reference systems recorded nvPM mass and number concentration, which were converted to their respective emission indices for comparison. The reference systems generally agreed to within 15% of the average nvPM number emission index and 30% of the average nvPM mass emission index. The only exception was for the nvPM mass instruments, which exhibited a higher variation as the concentration levels approached the limit of detection. The additional measured particle size distributions could be approximated to lognormal distributions with the geometric mean diameter ranging from 15 nm to 38 nm, and the geometric standard deviation varying between 1.53 and 1.92. The results from this study are a benchmark for the variability in standardized sampling and measurement systems for measuring aircraft engine nvPM emissions.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Closure between particulate matter concentrations measured ex situ by thermal–optical analysis and in situ by the CPMA–electrometer reference mass system.\n \n \n \n \n\n\n \n Corbin, J., C.; Moallemi, A.; Liu, F.; Gagné, S.; Olfert, J., S.; Smallwood, G., J.; and Lobo, P.\n\n\n \n\n\n\n Aerosol Science and Technology,1-17. 7 2020.\n \n\n\n\n
\n\n\n\n \n \n \"ClosureWebsite\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\n
\n
@article{\n title = {Closure between particulate matter concentrations measured ex situ by thermal–optical analysis and in situ by the CPMA–electrometer reference mass system},\n type = {article},\n year = {2020},\n keywords = {Jingkun Jiang},\n pages = {1-17},\n websites = {https://www.tandfonline.com/doi/full/10.1080/02786826.2020.1788710},\n month = {7},\n publisher = { Taylor & Francis },\n day = {20},\n id = {73434c56-e48e-31b1-bd3c-f6a132ea16b0},\n created = {2020-08-19T14:32:37.010Z},\n accessed = {2020-08-19},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-19T14:32:37.010Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {The thermal-optical analysis (TOA) of black carbon in particulate matter (PM) collected on filters has been recommended and used for the calibration of mass-concentration instruments. However, filt...},\n bibtype = {article},\n author = {Corbin, Joel C. and Moallemi, Alireza and Liu, Fengshan and Gagné, Stéphanie and Olfert, Jason S. and Smallwood, Greg J. and Lobo, Prem},\n doi = {10.1080/02786826.2020.1788710},\n journal = {Aerosol Science and Technology}\n}
\n
\n\n\n
\n The thermal-optical analysis (TOA) of black carbon in particulate matter (PM) collected on filters has been recommended and used for the calibration of mass-concentration instruments. However, filt...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Characterization of few-layer graphene aerosols by laser-induced incandescence.\n \n \n \n \n\n\n \n Musikhin, S.; Fortugno, P.; Corbin, J., C.; Smallwood, G., J.; Dreier, T.; Daun, K., J.; and Schulz, C.\n\n\n \n\n\n\n Carbon, 167: 870-880. 10 2020.\n \n\n\n\n
\n\n\n\n \n \n \"CharacterizationPaper\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 2 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{\n title = {Characterization of few-layer graphene aerosols by laser-induced incandescence},\n type = {article},\n year = {2020},\n pages = {870-880},\n volume = {167},\n month = {10},\n publisher = {Elsevier Ltd},\n day = {15},\n id = {1d2685e9-054c-3df0-94eb-89d82a0ab53d},\n created = {2020-09-02T14:26:22.198Z},\n accessed = {2020-09-02},\n file_attached = {true},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-09-02T14:26:27.173Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Gas-phase synthesis is a promising route for producing large amounts of high quality few-layer graphene (FLG) nanoparticles economically, but optimizing these processes requires a detailed understanding of the formation kinetics, which in turn demands diagnostics for characterizing this material in situ. This work reports the first laser-induced incandescence measurements on FLG aerosols. Temporally- and spectrally-resolved incandescence signals from FLG particles are measured and used to calculate pyrometric temperatures. Differences between incandescence signals and pyrometric temperatures obtained from FLG and aerosolized soot nanoaggregates are attributed to the larger absorption cross-section and specific surface area of FLG compared to soot. LII signal intensity is found to vary linearly with particle number concentration measured independently by a condensation particle counter. Overall, these results demonstrate the potential for laser-induced incandescence to measure FLG nanoparticle mass (volume) fraction and active surface area in situ, as well as to differentiate graphene from other types of carbonaceous nanomaterials online.},\n bibtype = {article},\n author = {Musikhin, Stanislav and Fortugno, Paolo and Corbin, Joel C. and Smallwood, Greg J. and Dreier, Thomas and Daun, Kyle J. and Schulz, Christof},\n doi = {10.1016/j.carbon.2020.05.052},\n journal = {Carbon}\n}
\n
\n\n\n
\n Gas-phase synthesis is a promising route for producing large amounts of high quality few-layer graphene (FLG) nanoparticles economically, but optimizing these processes requires a detailed understanding of the formation kinetics, which in turn demands diagnostics for characterizing this material in situ. This work reports the first laser-induced incandescence measurements on FLG aerosols. Temporally- and spectrally-resolved incandescence signals from FLG particles are measured and used to calculate pyrometric temperatures. Differences between incandescence signals and pyrometric temperatures obtained from FLG and aerosolized soot nanoaggregates are attributed to the larger absorption cross-section and specific surface area of FLG compared to soot. LII signal intensity is found to vary linearly with particle number concentration measured independently by a condensation particle counter. Overall, these results demonstrate the potential for laser-induced incandescence to measure FLG nanoparticle mass (volume) fraction and active surface area in situ, as well as to differentiate graphene from other types of carbonaceous nanomaterials online.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Non-Volatile Particle Number Emission Measurements with Catalytic Strippers: A Review.\n \n \n \n \n\n\n \n Giechaskiel, B.; Melas, A., D.; Lähde, T.; and Martini, G.\n\n\n \n\n\n\n Vehicles, 2(2): 342-364. 6 2020.\n \n\n\n\n
\n\n\n\n \n \n \"Non-VolatilePaper\n  \n \n \n \"Non-VolatileWebsite\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 7 downloads\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
@article{\n title = {Non-Volatile Particle Number Emission Measurements with Catalytic Strippers: A Review},\n type = {article},\n year = {2020},\n keywords = {artefact,catalytic stripper,evaporation tube,non-volatiles,particle measurement programme (PMP),portable emissions measurement systems (PEMS),solid particle number,vehicle emissions,volatile removal efficiency},\n pages = {342-364},\n volume = {2},\n websites = {https://www.mdpi.com/2624-8921/2/2/19},\n month = {6},\n publisher = {MDPI AG},\n day = {24},\n id = {95072e66-cfcf-3ade-90d5-ee21ec0a93f0},\n created = {2021-02-01T17:12:23.674Z},\n accessed = {2021-02-01},\n file_attached = {true},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2021-02-01T17:12:35.135Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {<p>Vehicle regulations include limits for non-volatile particle number emissions with sizes larger than 23 nm. The measurements are conducted with systems that remove the volatile particles by means of dilution and heating. Recently, the option of measuring from 10 nm was included in the Global Technical Regulation (GTR 15) as an additional option to the current &gt;23 nm methodology. In order to avoid artefacts, i.e., measuring volatile particles that have nucleated downstream of the evaporation tube, a heated oxidation catalyst (i.e., catalytic stripper) is required. This review summarizes the studies with laboratory aerosols that assessed the volatile removal efficiency of evaporation tube and catalytic stripper-based systems using hydrocarbons, sulfuric acid, mixture of them, and ammonium sulfate. Special emphasis was given to distinguish between artefacts that happened in the 10–23 nm range or below. Furthermore, studies with vehicles’ aerosols that reported artefacts were collected to estimate critical concentration levels of volatiles. Maximum expected levels of volatiles for mopeds, motorcycles, light-duty and heavy-duty vehicles were also summarized. Both laboratory and vehicle studies confirmed the superiority of catalytic strippers in avoiding artefacts. Open issues that need attention are the sulfur storage capacity and the standardization of technical requirements for catalytic strippers.</p>},\n bibtype = {article},\n author = {Giechaskiel, Barouch and Melas, Anastasios D. and Lähde, Tero and Martini, Giorgio},\n doi = {10.3390/vehicles2020019},\n journal = {Vehicles},\n number = {2}\n}
\n
\n\n\n
\n

Vehicle regulations include limits for non-volatile particle number emissions with sizes larger than 23 nm. The measurements are conducted with systems that remove the volatile particles by means of dilution and heating. Recently, the option of measuring from 10 nm was included in the Global Technical Regulation (GTR 15) as an additional option to the current >23 nm methodology. In order to avoid artefacts, i.e., measuring volatile particles that have nucleated downstream of the evaporation tube, a heated oxidation catalyst (i.e., catalytic stripper) is required. This review summarizes the studies with laboratory aerosols that assessed the volatile removal efficiency of evaporation tube and catalytic stripper-based systems using hydrocarbons, sulfuric acid, mixture of them, and ammonium sulfate. Special emphasis was given to distinguish between artefacts that happened in the 10–23 nm range or below. Furthermore, studies with vehicles’ aerosols that reported artefacts were collected to estimate critical concentration levels of volatiles. Maximum expected levels of volatiles for mopeds, motorcycles, light-duty and heavy-duty vehicles were also summarized. Both laboratory and vehicle studies confirmed the superiority of catalytic strippers in avoiding artefacts. Open issues that need attention are the sulfur storage capacity and the standardization of technical requirements for catalytic strippers.

\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Assessment of 10-nm Particle Number (PN) Portable Emissions Measurement Systems (PEMS) for Future Regulations.\n \n \n \n \n\n\n \n Giechaskiel, B.; Lähde, T.; Gandi, S.; Keller, S.; Kreutziger, P.; and Mamakos, A.\n\n\n \n\n\n\n International Journal of Environmental Research and Public Health, 17(11): 3878. 5 2020.\n \n\n\n\n
\n\n\n\n \n \n \"AssessmentPaper\n  \n \n \n \"AssessmentWebsite\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 3 downloads\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
@article{\n title = {Assessment of 10-nm Particle Number (PN) Portable Emissions Measurement Systems (PEMS) for Future Regulations},\n type = {article},\n year = {2020},\n keywords = {Air pollution,Catalytic stripper,Particle measurement programme (PMP),Portable emission measurement systems (PEMS),Real-driving emissions (RDE),Regeneration,Solid particle number,Sub-23 nm,Vehicle emissions},\n pages = {3878},\n volume = {17},\n websites = {https://www.mdpi.com/1660-4601/17/11/3878},\n month = {5},\n publisher = {MDPI AG},\n day = {30},\n id = {6b2b3286-9e40-346c-8824-6cc9b489288d},\n created = {2021-02-09T10:25:20.699Z},\n accessed = {2021-02-09},\n file_attached = {true},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2021-02-09T10:25:20.799Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {<p>The particle number (PN) emissions of vehicles equipped with particulate filters are low. However, there are technologies that can have high PN levels, especially below the currently lower regulated particle size of 23 nm. Sub-23-nm particles are also considered at least as dangerous as the larger ultrafine particles. For this reason, the European Union (EU) is planning to regulate particles down to 10 nm. In this study we compared prototype portable emission measurement systems (PEMS) and reference laboratory systems measuring from 10 nm. The tests included cycles and constant speeds, using vehicles fuelled with diesel, gasoline or liquefied petroleum gas (LPG). The results showed that the PEMS were within ±40% of the reference systems connected to the tailpipe and the dilution tunnel. Based on the positive findings and the detection efficiencies of the prototype instruments, a proposal for the technical specifications for the future regulation was drafted.</p>},\n bibtype = {article},\n author = {Giechaskiel, Barouch and Lähde, Tero and Gandi, Sawan and Keller, Stefan and Kreutziger, Philipp and Mamakos, Athanasios},\n doi = {10.3390/ijerph17113878},\n journal = {International Journal of Environmental Research and Public Health},\n number = {11}\n}
\n
\n\n\n
\n

The particle number (PN) emissions of vehicles equipped with particulate filters are low. However, there are technologies that can have high PN levels, especially below the currently lower regulated particle size of 23 nm. Sub-23-nm particles are also considered at least as dangerous as the larger ultrafine particles. For this reason, the European Union (EU) is planning to regulate particles down to 10 nm. In this study we compared prototype portable emission measurement systems (PEMS) and reference laboratory systems measuring from 10 nm. The tests included cycles and constant speeds, using vehicles fuelled with diesel, gasoline or liquefied petroleum gas (LPG). The results showed that the PEMS were within ±40% of the reference systems connected to the tailpipe and the dilution tunnel. Based on the positive findings and the detection efficiencies of the prototype instruments, a proposal for the technical specifications for the future regulation was drafted.

\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2019\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Repeatability and intermediate precision of a mass concentration calibration system.\n \n \n \n \n\n\n \n Titosky, J.; Momenimovahed, A.; Corbin, J.; Thomson, K.; Smallwood, G.; and Olfert, J., S.\n\n\n \n\n\n\n Aerosol Science and Technology, 53(6): 701-711. 6 2019.\n \n\n\n\n
\n\n\n\n \n \n \"RepeatabilityWebsite\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 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n\n\n\n
\n
@article{\n title = {Repeatability and intermediate precision of a mass concentration calibration system},\n type = {article},\n year = {2019},\n keywords = {Jingkun Jiang},\n pages = {701-711},\n volume = {53},\n websites = {https://www.tandfonline.com/doi/full/10.1080/02786826.2019.1592103},\n month = {6},\n publisher = {Taylor & Francis},\n day = {3},\n id = {73aad99b-cd1d-3b6f-a5fa-ed5e7d12c2e0},\n created = {2020-08-18T13:15:40.030Z},\n accessed = {2020-08-18},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-18T13:15:40.030Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Many aerosol instruments require calibration to make accurate measurements. A centrifugal particle mass analyzer (CPMA) and aerosol electrometer can be used to calibrate aerosol instruments that me...},\n bibtype = {article},\n author = {Titosky, Jordan and Momenimovahed, Ali and Corbin, Joel and Thomson, Kevin and Smallwood, Greg and Olfert, Jason S.},\n doi = {10.1080/02786826.2019.1592103},\n journal = {Aerosol Science and Technology},\n number = {6}\n}
\n
\n\n\n
\n Many aerosol instruments require calibration to make accurate measurements. A centrifugal particle mass analyzer (CPMA) and aerosol electrometer can be used to calibrate aerosol instruments that me...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Properties of carbon black produced by the thermal decomposition of methane in the products of premixed flames.\n \n \n \n \n\n\n \n Afroughi, M., J.; Falahati, F.; Kostiuk, L., W.; and Olfert, J., S.\n\n\n \n\n\n\n Journal of Aerosol Science, 131: 13-27. 5 2019.\n \n\n\n\n
\n\n\n\n \n \n \"PropertiesWebsite\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{\n title = {Properties of carbon black produced by the thermal decomposition of methane in the products of premixed flames},\n type = {article},\n year = {2019},\n pages = {13-27},\n volume = {131},\n websites = {https://www.sciencedirect.com/science/article/abs/pii/S0021850218302921?via%3Dihub},\n month = {5},\n publisher = {Pergamon},\n day = {1},\n id = {4e9c449a-0671-371b-9fa7-8e1abb499f02},\n created = {2020-08-18T13:16:01.163Z},\n accessed = {2020-08-18},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-18T13:16:01.163Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {The thermal decomposition of methane is a technique used to manufacture hydrogen gas and carbon black. The physical properties of carbon black produced by the thermal decomposition of methane (TDM) in the O2-deficient gas products of two premixed flames (propane- or methane-air) were investigated under different flow rates of decomposing methane injection (0.5–5 SLPM). An inverted burner was designed to provide a fuel-rich, laminar premixed flame, to produce hot gas into which methane was injected to thermally decompose inside a reactor. Particles from TDM were extracted by a nitrogen dilution system at the immediate exit of the reaction chamber, where another branch of the exhaust was dried and directed to a gas chromatograph. The carbon black particles were characterized by size resolved number concentration, mass concentration, effective density, volatility, and internal mixing state using different arrangements of a differential mobility analyzer, catalytic denuder, centrifugal particle mass analyzer, and condensation particle counter, as well as by morphology and primary particle size using transmission electron microscopy. A bimodal number-size distribution was observed at all conditions with count median diameters (CMDs) less than 58  nm and 21  nm when using propane- or methane-air premixed flames as the heat source, respectively. Higher number concentrations and mass concentrations with larger CMDs were achieved under lower flow rates of decomposing methane injection. For a given flow rate of decomposition methane, mass concentration and CMD increased significantly when using propane as the fuel, compared to the methane fuel. The size segregated mass fraction of internally mixed volatile content in particles was similar for both heat sources, showing a roughly constant fraction of volatile material in particles produced by the decomposition of 0.5 and 5 SLPM of methane (10%–30%) but a measurably larger fraction (55%–30%) with a decreasing trend as a function of particle size from the decomposition of 1 SLPM of methane. The effective density of denuded particles was similar, but slightly higher, than the effective density of soot from a wide range of internal combustion engines. A higher denuded effective density was observed in the particles with higher volatile contents (particles from decomposition of 1 SLPM of methane), suggesting the restructuring of carbon black into more compact clusters due to excessive volatile condensation. TEM analysis revealed some similarity between the produced carbon black and engine soot in terms of morphology and primary particle diameter (both below 40  nm). Beside the carbon black properties, the efficiency of TDM in this configuration was also investigated by evaluation of methane destruction efficiency, as well as carbon black and hydrogen production efficiencies, based on the product gas composition and the particle mass concentration. It was found that longer residence times corresponding to lower flow rates resulted in higher conversion efficiencies in terms of methane destruction efficiency (maximum of ∼95%), hydrogen production efficiency (maximum of ∼80%), and carbon black production efficiency (maximum of ∼1%). It was concluded that the significantly low efficiency of carbon black production is due to the high amount of CO formation, likely through the gasification process.},\n bibtype = {article},\n author = {Afroughi, Mohammad Javad and Falahati, Farjad and Kostiuk, Larry W. and Olfert, Jason S.},\n doi = {10.1016/J.JAEROSCI.2019.02.002},\n journal = {Journal of Aerosol Science}\n}
\n
\n\n\n
\n The thermal decomposition of methane is a technique used to manufacture hydrogen gas and carbon black. The physical properties of carbon black produced by the thermal decomposition of methane (TDM) in the O2-deficient gas products of two premixed flames (propane- or methane-air) were investigated under different flow rates of decomposing methane injection (0.5–5 SLPM). An inverted burner was designed to provide a fuel-rich, laminar premixed flame, to produce hot gas into which methane was injected to thermally decompose inside a reactor. Particles from TDM were extracted by a nitrogen dilution system at the immediate exit of the reaction chamber, where another branch of the exhaust was dried and directed to a gas chromatograph. The carbon black particles were characterized by size resolved number concentration, mass concentration, effective density, volatility, and internal mixing state using different arrangements of a differential mobility analyzer, catalytic denuder, centrifugal particle mass analyzer, and condensation particle counter, as well as by morphology and primary particle size using transmission electron microscopy. A bimodal number-size distribution was observed at all conditions with count median diameters (CMDs) less than 58  nm and 21  nm when using propane- or methane-air premixed flames as the heat source, respectively. Higher number concentrations and mass concentrations with larger CMDs were achieved under lower flow rates of decomposing methane injection. For a given flow rate of decomposition methane, mass concentration and CMD increased significantly when using propane as the fuel, compared to the methane fuel. The size segregated mass fraction of internally mixed volatile content in particles was similar for both heat sources, showing a roughly constant fraction of volatile material in particles produced by the decomposition of 0.5 and 5 SLPM of methane (10%–30%) but a measurably larger fraction (55%–30%) with a decreasing trend as a function of particle size from the decomposition of 1 SLPM of methane. The effective density of denuded particles was similar, but slightly higher, than the effective density of soot from a wide range of internal combustion engines. A higher denuded effective density was observed in the particles with higher volatile contents (particles from decomposition of 1 SLPM of methane), suggesting the restructuring of carbon black into more compact clusters due to excessive volatile condensation. TEM analysis revealed some similarity between the produced carbon black and engine soot in terms of morphology and primary particle diameter (both below 40  nm). Beside the carbon black properties, the efficiency of TDM in this configuration was also investigated by evaluation of methane destruction efficiency, as well as carbon black and hydrogen production efficiencies, based on the product gas composition and the particle mass concentration. It was found that longer residence times corresponding to lower flow rates resulted in higher conversion efficiencies in terms of methane destruction efficiency (maximum of ∼95%), hydrogen production efficiency (maximum of ∼80%), and carbon black production efficiency (maximum of ∼1%). It was concluded that the significantly low efficiency of carbon black production is due to the high amount of CO formation, likely through the gasification process.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Size and morphology of soot produced by a dual-fuel marine engine.\n \n \n \n \n\n\n \n Trivanovic, U.; Corbin, J., C.; Baldelli, A.; Peng, W.; Yang, J.; Kirchen, P.; Miller, J., W.; Lobo, P.; Gagné, S.; and Rogak, S., N.\n\n\n \n\n\n\n Journal of Aerosol Science, 138: 105448. 12 2019.\n \n\n\n\n
\n\n\n\n \n \n \"SizeWebsite\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{\n title = {Size and morphology of soot produced by a dual-fuel marine engine},\n type = {article},\n year = {2019},\n pages = {105448},\n volume = {138},\n websites = {https://www.sciencedirect.com/science/article/abs/pii/S002185021930552X?via%3Dihub},\n month = {12},\n publisher = {Pergamon},\n day = {1},\n id = {1c23e151-155c-311a-a7de-0a422edb2308},\n created = {2020-08-19T14:29:15.698Z},\n accessed = {2020-08-19},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-19T15:43:08.733Z},\n read = {true},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {In recent years, liquefied natural gas (LNG) has become a popular alternative to heavy fuel oils for use in marine engines due to its low cost and increasingly stringent emissions regulations from the International Maritime Organization (IMO). While natural gas (NG) combustion produces substantially less soot than other fossil fuels, its soot emissions are not negligible and have not been morphologically characterized. In this study, transmission electron microscopy (TEM) was used to interpret scanning mobility particle sizer (SMPS) distributions of catalytically-denuded particulate matter (soot) from an NG-powered commercial vessel. The vessel studied was powered by a dual-fuel compression ignition engine which uses a diesel pilot only to ignite the NG. Less than 2% of the engine's power is provided by the diesel pilot for engine loads above 50%. For comparison, tests were also run in diesel-only mode. During NG operation, the SMPS size distributions did not follow a lognormal distribution as expected. Instead, the distribution remained roughly constant from 10 nm to 100 nm. High resolution (HR)-TEM revealed particles as small as 30 nm in diameter were soot. From the TEM images, soot primary particle size and projected-area-equivalent diameter were estimated for individual soot particles. These data were then used to infer the effective density of the particles collected. In general, the primary particle size and correspondingly the estimated effective density of soot from this engine were larger than those from the “universal fit” proposed for soot in previous literature. For both operating conditions the source of the soot was likely the diesel pilot however, there were statistically significant differences between particles produced by the two fuelling modes. With LNG fueling, there were far fewer particles, and mean soot aggregate size decreased, but the primary particle size (for a 100 nm aggregate) increased, relative to diesel. Engine load did not have a statistically significant effect on soot morphology.},\n bibtype = {article},\n author = {Trivanovic, Una and Corbin, Joel C. and Baldelli, Alberto and Peng, Weihan and Yang, Jiacheng and Kirchen, Patrick and Miller, J. Wayne and Lobo, Prem and Gagné, Stéphanie and Rogak, Steven N.},\n doi = {10.1016/J.JAEROSCI.2019.105448},\n journal = {Journal of Aerosol Science}\n}
\n
\n\n\n
\n In recent years, liquefied natural gas (LNG) has become a popular alternative to heavy fuel oils for use in marine engines due to its low cost and increasingly stringent emissions regulations from the International Maritime Organization (IMO). While natural gas (NG) combustion produces substantially less soot than other fossil fuels, its soot emissions are not negligible and have not been morphologically characterized. In this study, transmission electron microscopy (TEM) was used to interpret scanning mobility particle sizer (SMPS) distributions of catalytically-denuded particulate matter (soot) from an NG-powered commercial vessel. The vessel studied was powered by a dual-fuel compression ignition engine which uses a diesel pilot only to ignite the NG. Less than 2% of the engine's power is provided by the diesel pilot for engine loads above 50%. For comparison, tests were also run in diesel-only mode. During NG operation, the SMPS size distributions did not follow a lognormal distribution as expected. Instead, the distribution remained roughly constant from 10 nm to 100 nm. High resolution (HR)-TEM revealed particles as small as 30 nm in diameter were soot. From the TEM images, soot primary particle size and projected-area-equivalent diameter were estimated for individual soot particles. These data were then used to infer the effective density of the particles collected. In general, the primary particle size and correspondingly the estimated effective density of soot from this engine were larger than those from the “universal fit” proposed for soot in previous literature. For both operating conditions the source of the soot was likely the diesel pilot however, there were statistically significant differences between particles produced by the two fuelling modes. With LNG fueling, there were far fewer particles, and mean soot aggregate size decreased, but the primary particle size (for a 100 nm aggregate) increased, relative to diesel. Engine load did not have a statistically significant effect on soot morphology.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2018\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Comparative performance of a thermal denuder and a catalytic stripper in sampling laboratory and marine exhaust aerosols.\n \n \n \n \n\n\n \n Amanatidis, S.; Ntziachristos, L.; Karjalainen, P.; Saukko, E.; Simonen, P.; Kuittinen, N.; Aakko-Saksa, P.; Timonen, H.; Rönkkö, T.; and Keskinen, J.\n\n\n \n\n\n\n Aerosol Science and Technology, 52(4): 420-432. 4 2018.\n \n\n\n\n
\n\n\n\n \n \n \"ComparativeWebsite\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\n
\n
@article{\n title = {Comparative performance of a thermal denuder and a catalytic stripper in sampling laboratory and marine exhaust aerosols},\n type = {article},\n year = {2018},\n keywords = {Matti Maricq},\n pages = {420-432},\n volume = {52},\n websites = {https://www.tandfonline.com/doi/full/10.1080/02786826.2017.1422236},\n month = {4},\n publisher = {Taylor & Francis},\n day = {3},\n id = {5a983c80-057e-3362-9127-16b7c24299d5},\n created = {2020-08-13T13:04:08.826Z},\n accessed = {2020-08-13},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-13T13:04:08.826Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {The performance of a thermal denuder (thermodenuder—TD) and a fresh catalytic stripper (CS) was assessed by sampling laboratory aerosol, produced by different combinations of sulfuric acid, octacos...},\n bibtype = {article},\n author = {Amanatidis, Stavros and Ntziachristos, Leonidas and Karjalainen, Panu and Saukko, Erkka and Simonen, Pauli and Kuittinen, Niina and Aakko-Saksa, Päivi and Timonen, Hilkka and Rönkkö, Topi and Keskinen, Jorma},\n doi = {10.1080/02786826.2017.1422236},\n journal = {Aerosol Science and Technology},\n number = {4}\n}
\n
\n\n\n
\n The performance of a thermal denuder (thermodenuder—TD) and a fresh catalytic stripper (CS) was assessed by sampling laboratory aerosol, produced by different combinations of sulfuric acid, octacos...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Sources of variance in BC mass measurements from a small marine engine: Influence of the instruments, fuels and loads.\n \n \n \n \n\n\n \n Jiang, Y.; Yang, J.; Gagné, S.; Chan, T., W.; Thomson, K.; Fofie, E.; Cary, R., A.; Rutherford, D.; Comer, B.; Swanson, J.; Lin, Y.; Van Rooy, P.; Asa-Awuku, A.; Jung, H.; Barsanti, K.; Karavalakis, G.; Cocker, D.; Durbin, T., D.; Miller, J., W.; and Johnson, K., C.\n\n\n \n\n\n\n Atmospheric Environment, 182: 128-137. 6 2018.\n \n\n\n\n
\n\n\n\n \n \n \"SourcesWebsite\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{\n title = {Sources of variance in BC mass measurements from a small marine engine: Influence of the instruments, fuels and loads},\n type = {article},\n year = {2018},\n pages = {128-137},\n volume = {182},\n websites = {https://www.sciencedirect.com/science/article/abs/pii/S135223101830147X},\n month = {6},\n publisher = {Pergamon},\n day = {1},\n id = {fc8bafc9-56cf-3b23-8b15-0a9ec7c5ddc9},\n created = {2020-08-19T15:41:56.845Z},\n accessed = {2020-08-19},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-19T15:41:56.845Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Knowledge of black carbon (BC) emission factors from ships is important from human health and environmental perspectives. A study of instruments measuring BC and fuels typically used in marine operation was carried out on a small marine engine. Six analytical methods measured the BC emissions in the exhaust of the marine engine operated at two load points (25% and 75%) while burning one of three fuels: a distillate marine (DMA), a low sulfur, residual marine (RMB-30) and a high-sulfur residual marine (RMG-380). The average emission factors with all instruments increased from 0.08 to 1.88 gBC/kg fuel in going from 25 to 75% load. An analysis of variance (ANOVA) tested BC emissions against instrument, load, and combined fuel properties and showed that both engine load and fuels had a statistically significant impact on BC emission factors. While BC emissions were impacted by the fuels used, none of the fuel properties investigated (sulfur content, viscosity, carbon residue and CCAI) was a primary driver for BC emissions. Of the two residual fuels, RMB-30 with the lower sulfur content, lower viscosity and lower residual carbon, had the highest BC emission factors. BC emission factors determined with the different instruments showed a good correlation with the PAS values with correlation coefficients R2 >0.95. A key finding of this research is the relative BC measured values were mostly independent of load and fuel, except for some instruments in certain fuel and load combinations.},\n bibtype = {article},\n author = {Jiang, Yu and Yang, Jiacheng and Gagné, Stéphanie and Chan, Tak W. and Thomson, Kevin and Fofie, Emmanuel and Cary, Robert A. and Rutherford, Dan and Comer, Bryan and Swanson, Jacob and Lin, Yue and Van Rooy, Paul and Asa-Awuku, Akua and Jung, Heejung and Barsanti, Kelley and Karavalakis, Georgios and Cocker, David and Durbin, Thomas D. and Miller, J. Wayne and Johnson, Kent C.},\n doi = {10.1016/J.ATMOSENV.2018.03.008},\n journal = {Atmospheric Environment}\n}
\n
\n\n\n
\n Knowledge of black carbon (BC) emission factors from ships is important from human health and environmental perspectives. A study of instruments measuring BC and fuels typically used in marine operation was carried out on a small marine engine. Six analytical methods measured the BC emissions in the exhaust of the marine engine operated at two load points (25% and 75%) while burning one of three fuels: a distillate marine (DMA), a low sulfur, residual marine (RMB-30) and a high-sulfur residual marine (RMG-380). The average emission factors with all instruments increased from 0.08 to 1.88 gBC/kg fuel in going from 25 to 75% load. An analysis of variance (ANOVA) tested BC emissions against instrument, load, and combined fuel properties and showed that both engine load and fuels had a statistically significant impact on BC emission factors. While BC emissions were impacted by the fuels used, none of the fuel properties investigated (sulfur content, viscosity, carbon residue and CCAI) was a primary driver for BC emissions. Of the two residual fuels, RMB-30 with the lower sulfur content, lower viscosity and lower residual carbon, had the highest BC emission factors. BC emission factors determined with the different instruments showed a good correlation with the PAS values with correlation coefficients R2 >0.95. A key finding of this research is the relative BC measured values were mostly independent of load and fuel, except for some instruments in certain fuel and load combinations.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2017\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Effective Density and Volatility of Particles Sampled from a Helicopter Gas Turbine Engine.\n \n \n \n \n\n\n \n Olfert, J., S.; Dickau, M.; Momenimovahed, A.; Saffaripour, M.; Thompson, K.; Smallwood, G.; Stettler, M., E., J.; Boies, A.; Sevcenco, Y.; Crayford, A.; and Johnson, M.\n\n\n \n\n\n\n Aerosol Science and Technology,0-0. 2 2017.\n \n\n\n\n
\n\n\n\n \n \n \"EffectiveWebsite\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 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n\n\n\n
\n
@article{\n title = {Effective Density and Volatility of Particles Sampled from a Helicopter Gas Turbine Engine},\n type = {article},\n year = {2017},\n keywords = {Matti Maricq},\n pages = {0-0},\n websites = {https://www.tandfonline.com/doi/full/10.1080/02786826.2017.1292346},\n month = {2},\n publisher = {Taylor & Francis},\n day = {8},\n id = {58746ca4-4c9e-305e-be7e-fdb6ca614cbf},\n created = {2017-11-17T15:14:58.983Z},\n accessed = {2017-03-13},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-17T15:14:58.983Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {ABSTRACTThe effective density and size-resolved volatility of particles emitted from a Rolls-Royce Gnome helicopter turboshaft engine are measured at two engine speed settings (13,000 and 22,000 RPM). The effective density of denuded and undenuded particles was measured. The denuded effective densities are similar to the effective densities of particles from a gas turbine with a double annular combustor as well as a wide variety of internal combustion engines. The denuded effective density measurements were also used to estimate the size and number of primary particles in the soot aggregates. The primary particle size estimates show that the primary particle size was smaller at lower engine speed (in agreement with transmission electron microscopy analysis). As a demonstration, the size-resolved volatility of particles emitted from the engine is measured with a system consisting of a differential mobility analyzer, centrifugal particle mass analyzer, condensation particle counter, and catalytic stripper. ...},\n bibtype = {article},\n author = {Olfert, Jason S. and Dickau, Matthew and Momenimovahed, Ali and Saffaripour, Meghdad and Thompson, Kevin and Smallwood, Greg and Stettler, Marc E. J. and Boies, Adam and Sevcenco, Yura and Crayford, Andrew and Johnson, Mark},\n doi = {10.1080/02786826.2017.1292346},\n journal = {Aerosol Science and Technology}\n}
\n
\n\n\n
\n ABSTRACTThe effective density and size-resolved volatility of particles emitted from a Rolls-Royce Gnome helicopter turboshaft engine are measured at two engine speed settings (13,000 and 22,000 RPM). The effective density of denuded and undenuded particles was measured. The denuded effective densities are similar to the effective densities of particles from a gas turbine with a double annular combustor as well as a wide variety of internal combustion engines. The denuded effective density measurements were also used to estimate the size and number of primary particles in the soot aggregates. The primary particle size estimates show that the primary particle size was smaller at lower engine speed (in agreement with transmission electron microscopy analysis). As a demonstration, the size-resolved volatility of particles emitted from the engine is measured with a system consisting of a differential mobility analyzer, centrifugal particle mass analyzer, condensation particle counter, and catalytic stripper. ...\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2016\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Methodology for Quantifying the Mixing State of an Aerosol.\n \n \n \n\n\n \n Dickau, M.; Olfert, J.; Stettler, M.; Boies, A., M.; Momenimovahed, A.; Thomson, K.; Smallwood, G.; and Johnson, M.\n\n\n \n\n\n\n Aerosol Science & Technology, 50(8): 759-772. 2016.\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 2 downloads\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
@article{\n title = {Methodology for Quantifying the Mixing State of an Aerosol},\n type = {article},\n year = {2016},\n keywords = {612-827-2421 http,ast,com,editorial office phone,manuscriptcentral,mc},\n pages = {759-772},\n volume = {50},\n id = {2be2be4a-9b16-3b0a-9014-8d1a23e58e58},\n created = {2017-11-17T15:14:58.957Z},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-14T12:52:28.733Z},\n read = {true},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {Mixing state refers to the relative proportions of chemical species in an aerosol, and the way these species are combined; either as a population where each particle consists of a single species (externally mixed) or where all particles individually consist of two or more species (internally mixed) or the case where some particles are pure and some particles consist of multiple species. The mixing state affects optical and hygroscopic properties, and quantifying it is therefore important for studying an aerosol's climate impact. In this article, we describe a method to quantify the volatile mixing state of an aerosol using a differential mobility analyzer, centrifugal particle mass analyzer, catalytic denuder, and condensation particle counter by measuring the mass distributions of the volatile and non-volatile components of an aerosol and determining how the material is mixed within and between particles as a function of mobility diameter. The method is demonstrated using two aerosol samples from a miniCAST soot generator, one with a high elemental carbon (EC) content, and one with a high organic carbon (OC) content. The measurements are presented in terms of the mass distribution of the volatile and non-volatile material, as well as measures of diversity and mixing state parameter. It was found that the high-EC soot nearly consisted of only pure particles where 86% of the total mass was non-volatile. The high-OC soot consisted of either pure volatile particles or particles that contained a mixture of volatile and non-volatile material where 8% of the total mass was pure volatile particles and 70% was non-volatile material (with the remaining 22% being volatile material condensed on non-volatile particles).},\n bibtype = {article},\n author = {Dickau, Matthew and Olfert, Jason and Stettler, Marc and Boies, Adam M and Momenimovahed, Ali and Thomson, Kevin and Smallwood, Greg and Johnson, Mark},\n doi = {10.1080/02786826.2016.1185509},\n journal = {Aerosol Science & Technology},\n number = {8}\n}
\n
\n\n\n
\n Mixing state refers to the relative proportions of chemical species in an aerosol, and the way these species are combined; either as a population where each particle consists of a single species (externally mixed) or where all particles individually consist of two or more species (internally mixed) or the case where some particles are pure and some particles consist of multiple species. The mixing state affects optical and hygroscopic properties, and quantifying it is therefore important for studying an aerosol's climate impact. In this article, we describe a method to quantify the volatile mixing state of an aerosol using a differential mobility analyzer, centrifugal particle mass analyzer, catalytic denuder, and condensation particle counter by measuring the mass distributions of the volatile and non-volatile components of an aerosol and determining how the material is mixed within and between particles as a function of mobility diameter. The method is demonstrated using two aerosol samples from a miniCAST soot generator, one with a high elemental carbon (EC) content, and one with a high organic carbon (OC) content. The measurements are presented in terms of the mass distribution of the volatile and non-volatile material, as well as measures of diversity and mixing state parameter. It was found that the high-EC soot nearly consisted of only pure particles where 86% of the total mass was non-volatile. The high-OC soot consisted of either pure volatile particles or particles that contained a mixture of volatile and non-volatile material where 8% of the total mass was pure volatile particles and 70% was non-volatile material (with the remaining 22% being volatile material condensed on non-volatile particles).\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Response of real-time black carbon mass instruments to mini-CAST soot.\n \n \n \n \n\n\n \n Durdina, L.; Lobo, P.; Trueblood, M., B.; Black, E., A.; Achterberg, S.; Hagen, D., E.; Brem, B., T.; and Wang, J.\n\n\n \n\n\n\n Aerosol Science and Technology, 50(9): 906-918. 9 2016.\n \n\n\n\n
\n\n\n\n \n \n \"ResponseWebsite\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\n
\n
@article{\n title = {Response of real-time black carbon mass instruments to mini-CAST soot},\n type = {article},\n year = {2016},\n keywords = {Matti Maricq},\n pages = {906-918},\n volume = {50},\n websites = {https://www.tandfonline.com/doi/full/10.1080/02786826.2016.1204423},\n month = {9},\n publisher = {Taylor & Francis},\n day = {1},\n id = {8acc6292-b1f4-368b-b694-5b74d726f4d6},\n created = {2020-08-19T14:28:35.050Z},\n accessed = {2020-08-19},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-19T14:28:35.050Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Soot is a climate forcer and a dangerous air pollutant that has been increasingly regulated. In aviation, regulatory measurements of soot mass concentration in the exhaust of aircraft turbine engin...},\n bibtype = {article},\n author = {Durdina, Lukas and Lobo, Prem and Trueblood, Max B. and Black, Elizabeth A. and Achterberg, Steven and Hagen, Donald E. and Brem, Benjamin T. and Wang, Jing},\n doi = {10.1080/02786826.2016.1204423},\n journal = {Aerosol Science and Technology},\n number = {9}\n}
\n
\n\n\n
\n Soot is a climate forcer and a dangerous air pollutant that has been increasingly regulated. In aviation, regulatory measurements of soot mass concentration in the exhaust of aircraft turbine engin...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Demonstration of a Regulatory Method for Aircraft Engine Nonvolatile PM Emissions Measurements with Conventional and Isoparaffinic Kerosene fuels.\n \n \n \n \n\n\n \n Lobo, P.; Condevaux, J.; Yu, Z.; Kuhlmann, J.; Hagen, D., E.; Miake-Lye, R., C.; Whitefield, P., D.; and Raper, D., W.\n\n\n \n\n\n\n Energy & Fuels, 30(9): 7770-7777. 9 2016.\n \n\n\n\n
\n\n\n\n \n \n \"DemonstrationWebsite\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{\n title = {Demonstration of a Regulatory Method for Aircraft Engine Nonvolatile PM Emissions Measurements with Conventional and Isoparaffinic Kerosene fuels},\n type = {article},\n year = {2016},\n pages = {7770-7777},\n volume = {30},\n websites = {https://pubs.acs.org/doi/10.1021/acs.energyfuels.6b01581},\n month = {9},\n publisher = {American Chemical Society},\n day = {15},\n id = {f29f7f3f-4857-30a7-907b-fa884746f992},\n created = {2020-08-19T14:28:49.805Z},\n accessed = {2020-08-19},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-19T14:28:49.805Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {The aviation industry is exploring the economic viability and environmental sustainability of the use of alternative fuels to power aircraft main engines and auxiliary power units. The Internationa...},\n bibtype = {article},\n author = {Lobo, Prem and Condevaux, Jamey and Yu, Zhenhong and Kuhlmann, Joshua and Hagen, Donald E. and Miake-Lye, Richard C. and Whitefield, Philip D. and Raper, David W.},\n doi = {10.1021/acs.energyfuels.6b01581},\n journal = {Energy & Fuels},\n number = {9}\n}
\n
\n\n\n
\n The aviation industry is exploring the economic viability and environmental sustainability of the use of alternative fuels to power aircraft main engines and auxiliary power units. The Internationa...\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2015\n \n \n (5)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Particle Emission Measurements from L-Category Vehicles.\n \n \n \n\n\n \n Giechaskiel, B.; Zardini, A.; and Martini, G.\n\n\n \n\n\n\n SAE International, 8(5). 2015.\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\n \n \n \n \n \n \n \n\n  \n \n \n \n \n\n\n\n
\n
@article{\n title = {Particle Emission Measurements from L-Category Vehicles},\n type = {article},\n year = {2015},\n volume = {8},\n id = {5bb36ae5-bee2-3e59-8ceb-fca39f0b7c12},\n created = {2017-11-17T15:14:58.999Z},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-17T15:14:58.999Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n bibtype = {article},\n author = {Giechaskiel, Barouch and Zardini, Alessandro and Martini, Giorgio},\n doi = {10.4271/2015-24-2512},\n journal = {SAE International},\n number = {5},\n keywords = {Catalytic Stripper}\n}
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Effective Density and Mass-Mobility Exponent of Aircraft Turbine Particulate Matter.\n \n \n \n \n\n\n \n Johnson, T., J.; Olfert, J., S.; Symonds, J., P., R.; Johnson, M.; Rindlisbacher, T.; Swanson, J., J.; Boies, A., M.; Thomson, K.; Smallwood, G.; Walters, D.; Sevcenco, Y.; Crayford, A.; Dastanpour, R.; Rogak, S., N.; Durdina, L.; Bahk, Y., K.; Brem, B.; and Wang, J.\n\n\n \n\n\n\n Journal of Propulsion and Power, 31(2): 573-582. 3 2015.\n \n\n\n\n
\n\n\n\n \n \n \"EffectiveWebsite\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{\n title = {Effective Density and Mass-Mobility Exponent of Aircraft Turbine Particulate Matter},\n type = {article},\n year = {2015},\n pages = {573-582},\n volume = {31},\n websites = {http://arc.aiaa.org/doi/10.2514/1.B35367},\n month = {3},\n id = {b294ece4-db8a-3b7d-a99d-9ff24db10ebe},\n created = {2017-11-17T15:14:59.000Z},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-14T12:52:28.709Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {Copyright ? 2014 by the American Institute of Aeronautics and Astronautics, Inc.A centrifugal particle mass analyzer and a modified differential mobility spectrometer were used to measure the mass and mobility of particulate matter emitted by CFM56-5B4/2P, CFM56-7B26/3, and PW4000-100 gas turbine engine sources. The mass-mobility exponent of the particulate matter from the CFM56-5B4/2P engine ranged from 2.68 to 2.82, whereas the effective particle densities varied from 600 to 1250 kg/m3, depending on the static engine thrust and sampling methodology used. The effective particle densities from the CFM56-7B26/3 and PW4000-100 engines also fell within this range. The sample was conditioned with or without a catalytic stripper and with or without dilution, which caused the effective density to change, indicating the presence of condensed semivolatile material on the particles. Variability of the determined effective densities across different engine thrusts, based on the scattering about the line of best fit, was lowest for the diluted samples and highest for the undiluted sample without a catalytic stripper. This variability indicates that the relative amount of semivolatile material produced was engine thrust dependent. It was found that the nonvolatile particulate matter, effective particle density (in kilograms per cubic meter) of the CFM56-5B4/2P engine at relative thrusts below 30% could be approximated using the particle mobility diameter (dme in meters) with 11.92d(2.76-3) me.},\n bibtype = {article},\n author = {Johnson, Tyler J. and Olfert, Jason S. and Symonds, Jonathan P. R. and Johnson, Mark and Rindlisbacher, Theo and Swanson, Jacob J. and Boies, Adam M. and Thomson, Kevin and Smallwood, Greg and Walters, David and Sevcenco, Yura and Crayford, Andrew and Dastanpour, Ramin and Rogak, Steven N. and Durdina, Lukas and Bahk, Yeon Kyoung and Brem, Benjamin and Wang, Jing},\n doi = {10.2514/1.B35367},\n journal = {Journal of Propulsion and Power},\n number = {2}\n}
\n
\n\n\n
\n Copyright ? 2014 by the American Institute of Aeronautics and Astronautics, Inc.A centrifugal particle mass analyzer and a modified differential mobility spectrometer were used to measure the mass and mobility of particulate matter emitted by CFM56-5B4/2P, CFM56-7B26/3, and PW4000-100 gas turbine engine sources. The mass-mobility exponent of the particulate matter from the CFM56-5B4/2P engine ranged from 2.68 to 2.82, whereas the effective particle densities varied from 600 to 1250 kg/m3, depending on the static engine thrust and sampling methodology used. The effective particle densities from the CFM56-7B26/3 and PW4000-100 engines also fell within this range. The sample was conditioned with or without a catalytic stripper and with or without dilution, which caused the effective density to change, indicating the presence of condensed semivolatile material on the particles. Variability of the determined effective densities across different engine thrusts, based on the scattering about the line of best fit, was lowest for the diluted samples and highest for the undiluted sample without a catalytic stripper. This variability indicates that the relative amount of semivolatile material produced was engine thrust dependent. It was found that the nonvolatile particulate matter, effective particle density (in kilograms per cubic meter) of the CFM56-5B4/2P engine at relative thrusts below 30% could be approximated using the particle mobility diameter (dme in meters) with 11.92d(2.76-3) me.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Particle Emission Characteristics of a Gas Turbine with a Double Annular Combustor.\n \n \n \n \n\n\n \n Boies, A., M.; Stettler, M., E., J.; Swanson, J., J.; Johnson, T., J.; Olfert, J., S.; Johnson, M.; Eggersdorfer, M., L.; Rindlisbacher, T.; Wang, J.; Thomson, K.; Smallwood, G.; Sevcenco, Y.; Walters, D.; Williams, P., I.; Corbin, J.; Mensah, A., A.; Symonds, J.; Dastanpour, R.; and Rogak, S., N.\n\n\n \n\n\n\n Aerosol Science and Technology, 49(9): 842-855. 9 2015.\n \n\n\n\n
\n\n\n\n \n \n \"ParticleWebsite\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{\n title = {Particle Emission Characteristics of a Gas Turbine with a Double Annular Combustor},\n type = {article},\n year = {2015},\n pages = {842-855},\n volume = {49},\n websites = {http://www.tandfonline.com/doi/full/10.1080/02786826.2015.1078452},\n month = {9},\n publisher = {Taylor & Francis},\n day = {2},\n id = {0c4427d4-2c4e-3dbd-8958-585c1a35ba26},\n created = {2017-11-20T14:54:13.096Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.096Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {The total climate, air quality, and health impact of aircraft black carbon (BC) emissions depend on quantity (mass and number concentration) as well as morphology (fractal dimension and surface area) of emitted BC aggregates. This study examines multiple BC emission metrics from a gas turbine with a double annular combustor, CFM56-5B4-2P. As a part of the SAMPLE III.2 campaign, concurrent measurements of particle mobility, particle mass, particle number concentration, and mass concentration, as well as collection of transmission electron microscopy (TEM) samples, allowed for characterization of the BC emissions. Mass- and number-based emission indices were strongly influenced by thrust setting during pilot combustion and ranged from <1 to 208 mg/kg-fuel and 3 ×× 1012 to 3 ×× 1016 particles/kg-fuel, respectively. Mobility measurements indicated that mean diameters ranged from 7 to 44 nm with a strong dependence on thrust during pilot-only combustion. Using aggregation and sintering theory with empirical ef...},\n bibtype = {article},\n author = {Boies, Adam M. and Stettler, Marc E. J. and Swanson, Jacob J. and Johnson, Tyler J. and Olfert, Jason S. and Johnson, Mark and Eggersdorfer, Max L. and Rindlisbacher, Theo and Wang, Jing and Thomson, Kevin and Smallwood, Greg and Sevcenco, Yura and Walters, David and Williams, Paul I. and Corbin, Joel and Mensah, Amewu A. and Symonds, Jonathan and Dastanpour, Ramin and Rogak, Steven N.},\n doi = {10.1080/02786826.2015.1078452},\n journal = {Aerosol Science and Technology},\n number = {9}\n}
\n
\n\n\n
\n The total climate, air quality, and health impact of aircraft black carbon (BC) emissions depend on quantity (mass and number concentration) as well as morphology (fractal dimension and surface area) of emitted BC aggregates. This study examines multiple BC emission metrics from a gas turbine with a double annular combustor, CFM56-5B4-2P. As a part of the SAMPLE III.2 campaign, concurrent measurements of particle mobility, particle mass, particle number concentration, and mass concentration, as well as collection of transmission electron microscopy (TEM) samples, allowed for characterization of the BC emissions. Mass- and number-based emission indices were strongly influenced by thrust setting during pilot combustion and ranged from <1 to 208 mg/kg-fuel and 3 ×× 1012 to 3 ×× 1016 particles/kg-fuel, respectively. Mobility measurements indicated that mean diameters ranged from 7 to 44 nm with a strong dependence on thrust during pilot-only combustion. Using aggregation and sintering theory with empirical ef...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Measurement of Aircraft Engine Non-Volatile PM Emissions: Results of the Aviation-Particle Regulatory Instrumentation Demonstration Experiment (A-PRIDE) 4 Campaign.\n \n \n \n \n\n\n \n Lobo, P.; Durdina, L.; Smallwood, G., J.; Rindlisbacher, T.; Siegerist, F.; Black, E., A.; Yu, Z.; Mensah, A., A.; Hagen, D., E.; Miake-Lye, R., C.; Thomson, K., A.; Brem, B., T.; Corbin, J., C.; Abegglen, M.; Sierau, B.; Whitefield, P., D.; and Wang, J.\n\n\n \n\n\n\n Aerosol Science and Technology, 49(7): 472-484. 7 2015.\n \n\n\n\n
\n\n\n\n \n \n \"MeasurementWebsite\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{\n title = {Measurement of Aircraft Engine Non-Volatile PM Emissions: Results of the Aviation-Particle Regulatory Instrumentation Demonstration Experiment (A-PRIDE) 4 Campaign},\n type = {article},\n year = {2015},\n pages = {472-484},\n volume = {49},\n websites = {http://www.tandfonline.com/doi/full/10.1080/02786826.2015.1047012},\n month = {7},\n publisher = {Taylor & Francis},\n day = {3},\n id = {edcca9cb-f2d3-31e1-883e-7c09c288098e},\n created = {2020-08-14T12:51:34.670Z},\n accessed = {2020-08-14},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-14T12:51:34.670Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {This study reports the first of a kind data on aircraft engine non-volatile particulate matter (nvPM) number- and mass-based emissions using standardized systems. Two compliant sampling and measure...},\n bibtype = {article},\n author = {Lobo, Prem and Durdina, Lukas and Smallwood, Gregory J. and Rindlisbacher, Theodor and Siegerist, Frithjof and Black, Elizabeth A. and Yu, Zhenhong and Mensah, Amewu A. and Hagen, Donald E. and Miake-Lye, Richard C. and Thomson, Kevin A. and Brem, Benjamin T. and Corbin, Joel C. and Abegglen, Manuel and Sierau, Berko and Whitefield, Philip D. and Wang, Jing},\n doi = {10.1080/02786826.2015.1047012},\n journal = {Aerosol Science and Technology},\n number = {7}\n}
\n
\n\n\n
\n This study reports the first of a kind data on aircraft engine non-volatile particulate matter (nvPM) number- and mass-based emissions using standardized systems. Two compliant sampling and measure...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n No 主観的健康感を中心とした在宅高齢者における 健康関連指標に関する共分散構造分析Title.\n \n \n \n \n\n\n \n Luz Yolanda Toro Suarez\n\n\n \n\n\n\n 2015.\n \n\n\n\n
\n\n\n\n \n \n \"NoWebsite\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{\n title = {No 主観的健康感を中心とした在宅高齢者における 健康関連指標に関する共分散構造分析Title},\n type = {misc},\n year = {2015},\n pages = {1-27},\n websites = {https://acp.copernicus.org/articles/21/635/2021/acp-21-635-2021.pdf},\n id = {74262075-ef96-317e-baf8-6c15e3346926},\n created = {2021-02-02T09:54:28.123Z},\n accessed = {2021-02-02},\n file_attached = {false},\n profile_id = {5d11b1c1-e46c-3dd5-b82f-eba7b0a695d2},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2021-02-04T16:00:43.279Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n bibtype = {misc},\n author = {Luz Yolanda Toro Suarez, undefined}\n}
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2014\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n A solid particle number measurement system including nanoparticles smaller than 23 nanometers.\n \n \n \n \n\n\n \n Otsuki, Y.; Takeda, K.; Haruta, K.; and Mori, N.\n\n\n \n\n\n\n In SAE Technical Papers, volume 1, 4 2014. \n \n\n\n\n
\n\n\n\n \n \n \"AWebsite\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
@inproceedings{\n title = {A solid particle number measurement system including nanoparticles smaller than 23 nanometers},\n type = {inproceedings},\n year = {2014},\n volume = {1},\n websites = {http://papers.sae.org/2014-01-1604/},\n month = {4},\n day = {1},\n id = {cedc6efe-2173-326c-ad0b-ee24172ae3d1},\n created = {2017-11-20T14:54:13.886Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-08-14T12:52:28.805Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n citation_key = {Otsuki2014},\n private_publication = {false},\n abstract = {The particle number (PN) emission regulation has been implemented since 2011 in Europe. PN measurement procedure defined in ECE regulation No. 83 requires detecting only solid particles by eliminating volatile particles, the concentrations of which are highly influenced by dilution conditions, using a volatile particle remover (VPR). To measure PN concentration after the VPR, a particle number counter (PNC) which has detection threshold at a particle size of 23 nm is used, because most solid particles generated by automotive engines are considered to be larger than 23 nm. On the other hand, several studies have reported the existence of solid and volatile particles smaller than 23 nm in engine exhaust. This paper describes investigation into a measurement method for ultrafine PNCs with thresholds of below 23 nm and evaluation of the VPR performance for the particles in this size range. The detection efficiency of an ultrafine PNC was verified by following the ECE regulation procedure. In addition, the possibility of re-nucleation of volatile particles at the VPR outlet was evaluated, because the re-nucleated volatile particles which are usually smaller than 23 nm will cause a high bias for solid particle measurements by the ultrafine PNC compared to the standard PNC. We applied a hot catalytic stripper (HCS) in order to effectively remove volatile fractions instead of an evaporation tube (ET). The modified VPR with the HCS showed higher removal efficiency for volatile particles than the ET. However, technical challenges like higher losses of the nanoparticles were revealed by this investigation. Copyright © 2014 SAE International.},\n bibtype = {inproceedings},\n author = {Otsuki, Yoshinori and Takeda, Kenji and Haruta, Kazuhiko and Mori, Nobuhisa},\n doi = {10.4271/2014-01-1604},\n booktitle = {SAE Technical Papers}\n}
\n
\n\n\n
\n The particle number (PN) emission regulation has been implemented since 2011 in Europe. PN measurement procedure defined in ECE regulation No. 83 requires detecting only solid particles by eliminating volatile particles, the concentrations of which are highly influenced by dilution conditions, using a volatile particle remover (VPR). To measure PN concentration after the VPR, a particle number counter (PNC) which has detection threshold at a particle size of 23 nm is used, because most solid particles generated by automotive engines are considered to be larger than 23 nm. On the other hand, several studies have reported the existence of solid and volatile particles smaller than 23 nm in engine exhaust. This paper describes investigation into a measurement method for ultrafine PNCs with thresholds of below 23 nm and evaluation of the VPR performance for the particles in this size range. The detection efficiency of an ultrafine PNC was verified by following the ECE regulation procedure. In addition, the possibility of re-nucleation of volatile particles at the VPR outlet was evaluated, because the re-nucleated volatile particles which are usually smaller than 23 nm will cause a high bias for solid particle measurements by the ultrafine PNC compared to the standard PNC. We applied a hot catalytic stripper (HCS) in order to effectively remove volatile fractions instead of an evaporation tube (ET). The modified VPR with the HCS showed higher removal efficiency for volatile particles than the ET. However, technical challenges like higher losses of the nanoparticles were revealed by this investigation. Copyright © 2014 SAE International.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2013\n \n \n (9)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Simultaneous Reduction of Particulate Matter and NO x Emissions Using 4-Way Catalyzed Filtration Systems.\n \n \n \n \n\n\n \n Swanson, J., J.; Watts, W., F.; Newman, R., A.; Ziebarth, R., R.; and Kittelson, D., B.\n\n\n \n\n\n\n Environmental Science & Technology, 47(9): 4521-4527. 5 2013.\n \n\n\n\n
\n\n\n\n \n \n \"SimultaneousWebsite\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{\n title = {Simultaneous Reduction of Particulate Matter and NO <sub> <i>x</i> </sub> Emissions Using 4-Way Catalyzed Filtration Systems},\n type = {article},\n year = {2013},\n pages = {4521-4527},\n volume = {47},\n websites = {http://pubs.acs.org/doi/10.1021/es304971h},\n month = {5},\n publisher = {American Chemical Society},\n day = {7},\n id = {3f1b026f-f002-373f-b277-2bdec89859ea},\n created = {2017-11-20T14:54:13.355Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.355Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {The next generation of diesel emission control devices includes 4-way catalyzed filtration systems (4WCFS) consisting of both NOx and diesel particulate matter (DPM) control. A methodology was developed to simultaneously evaluate the NOx and DPM control performance of miniature 4WCFS made from acicular mullite, an advanced ceramic material (ACM), that were challenged with diesel exhaust. The impact of catalyst loading and substrate porosity on catalytic performance of the NOx trap was evaluated. Simultaneously with NOx measurements, the real-time solid particle filtration performance of catalyst-coated standard and high porosity filters was determined for steady-state and regenerative conditions. The use of high porosity ACM 4-way catalyzed filtration systems reduced NOx by 99% and solid and total particulate matter by 95% when averaged over 10 regeneration cycles. A “regeneration cycle” refers to an oxidizing (“lean”) exhaust condition followed by a reducing (“rich”) exhaust condition resulting in NOx st...},\n bibtype = {article},\n author = {Swanson, Jacob J. and Watts, Winthrop F. and Newman, Robert A. and Ziebarth, Robin R. and Kittelson, David B.},\n doi = {10.1021/es304971h},\n journal = {Environmental Science & Technology},\n number = {9}\n}
\n
\n\n\n
\n The next generation of diesel emission control devices includes 4-way catalyzed filtration systems (4WCFS) consisting of both NOx and diesel particulate matter (DPM) control. A methodology was developed to simultaneously evaluate the NOx and DPM control performance of miniature 4WCFS made from acicular mullite, an advanced ceramic material (ACM), that were challenged with diesel exhaust. The impact of catalyst loading and substrate porosity on catalytic performance of the NOx trap was evaluated. Simultaneously with NOx measurements, the real-time solid particle filtration performance of catalyst-coated standard and high porosity filters was determined for steady-state and regenerative conditions. The use of high porosity ACM 4-way catalyzed filtration systems reduced NOx by 99% and solid and total particulate matter by 95% when averaged over 10 regeneration cycles. A “regeneration cycle” refers to an oxidizing (“lean”) exhaust condition followed by a reducing (“rich”) exhaust condition resulting in NOx st...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Evaluation of an oxidation catalyst (\"catalytic stripper\") in eliminating volatile material from combustion aerosol.\n \n \n \n \n\n\n \n Amanatidis, S.; Ntziachristos, L.; Giechaskiel, B.; Katsaounis, D.; Samaras, Z.; and Bergmann, A.\n\n\n \n\n\n\n Journal of Aerosol Science, 57: 144-155. 3 2013.\n \n\n\n\n
\n\n\n\n \n \n \"EvaluationWebsite\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 \n \n \n \n\n\n\n
\n
@article{\n title = {Evaluation of an oxidation catalyst ("catalytic stripper") in eliminating volatile material from combustion aerosol},\n type = {article},\n year = {2013},\n keywords = {Exhaust aerosol,Exhaust sampling,Particle emissions,Sampling conditions},\n pages = {144-155},\n volume = {57},\n websites = {http://www.sciencedirect.com/science/article/pii/S0021850212001966},\n month = {3},\n publisher = {Pergamon},\n day = {1},\n id = {19a7a10a-0215-37b4-a4e7-dbd523d64fda},\n created = {2017-11-20T14:54:13.358Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.514Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {Combustion aerosol is a mixture of solid and volatile particulate matter. Separation of solid particles for research or regulatory purposes is often conducted with thermal treatment of the aerosol. For example, European automotive emission regulations address solid particles above 23. nm, which are separated by dilution and heating in a volatile particle remover (VPR). This study evaluated an oxidation catalyst - often referred to as a "catalytic stripper" (CS) - as an alternative technique to remove volatile components. A version of the CS was examined in this paper, where the oxidation catalyst was combined with a sulphur trap in order to oxidise hydrocarbon species and to bind sulphates on the CS surface. In order to characterise the performance of the CS, the position of the sulphur trap upstream or downstream of the oxidation catalyst was examined in relation to the light-off temperature, hydrocarbon oxidation efficiency, and sulphur storage capacity, defined as the point where sulphate particles start to form downstream of the CS. With the best performance achieved when the trap was positioned downstream of the oxidation catalyst, the CS was then characterised in terms of particle losses in the range 6-100. nm. Losses were found rather independent of particle size above 30. nm but significantly increased below 23. nm. The efficiency in removing volatile particles was characterised using tetracontane particles. Furthermore, the overall performance of the CS was compared against the VPR by using diesel nucleation mode particles as the challenge aerosol. Results showed that the CS could directly be used as an alternative to VPR for combustion aerosol measurements if only particles above 23. nm were considered. Extending the measurement below this range would also be possible. This would however require an evaporation tube to vaporise material before this reached the CS and attention in addressing the rapidly increasing losses with decreasing particle size in this range. © 2012 Elsevier Ltd.},\n bibtype = {article},\n author = {Amanatidis, Stavros and Ntziachristos, Leonidas and Giechaskiel, Barouch and Katsaounis, Dimitrios and Samaras, Zissis and Bergmann, Alexander},\n doi = {10.1016/j.jaerosci.2012.12.001},\n journal = {Journal of Aerosol Science}\n}
\n
\n\n\n
\n Combustion aerosol is a mixture of solid and volatile particulate matter. Separation of solid particles for research or regulatory purposes is often conducted with thermal treatment of the aerosol. For example, European automotive emission regulations address solid particles above 23. nm, which are separated by dilution and heating in a volatile particle remover (VPR). This study evaluated an oxidation catalyst - often referred to as a \"catalytic stripper\" (CS) - as an alternative technique to remove volatile components. A version of the CS was examined in this paper, where the oxidation catalyst was combined with a sulphur trap in order to oxidise hydrocarbon species and to bind sulphates on the CS surface. In order to characterise the performance of the CS, the position of the sulphur trap upstream or downstream of the oxidation catalyst was examined in relation to the light-off temperature, hydrocarbon oxidation efficiency, and sulphur storage capacity, defined as the point where sulphate particles start to form downstream of the CS. With the best performance achieved when the trap was positioned downstream of the oxidation catalyst, the CS was then characterised in terms of particle losses in the range 6-100. nm. Losses were found rather independent of particle size above 30. nm but significantly increased below 23. nm. The efficiency in removing volatile particles was characterised using tetracontane particles. Furthermore, the overall performance of the CS was compared against the VPR by using diesel nucleation mode particles as the challenge aerosol. Results showed that the CS could directly be used as an alternative to VPR for combustion aerosol measurements if only particles above 23. nm were considered. Extending the measurement below this range would also be possible. This would however require an evaporation tube to vaporise material before this reached the CS and attention in addressing the rapidly increasing losses with decreasing particle size in this range. © 2012 Elsevier Ltd.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Updated Correlation Between Aircraft Smoke Number and Black Carbon Concentration.\n \n \n \n \n\n\n \n Stettler, M., E., J.; Swanson, J., J.; Barrett, S., R., H.; and Boies, A., M.\n\n\n \n\n\n\n Aerosol Science and Technology, 47(11): 1205-1214. 11 2013.\n \n\n\n\n
\n\n\n\n \n \n \"UpdatedWebsite\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 2 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{\n title = {Updated Correlation Between Aircraft Smoke Number and Black Carbon Concentration},\n type = {article},\n year = {2013},\n pages = {1205-1214},\n volume = {47},\n websites = {http://www.tandfonline.com/doi/abs/10.1080/02786826.2013.829908},\n month = {11},\n publisher = { Taylor & Francis Group },\n id = {bd19232f-0647-3282-97d7-77c549460d3c},\n created = {2017-11-20T14:54:13.462Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.462Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Aircraft emissions of black carbon (BC) contribute to anthropogenic climate forcing and degrade air quality. The smoke number (SN) is the current regulatory measure of aircraft particulate matter emissions and quantifies exhaust plume visibility. Several correlations between SN and the exhaust mass concentration of BC (C BC) have been developed, based on measurements relevant to older aircraft engines. These form the basis of the current standard method used to estimate aircraft BC emissions (First Order Approximation version 3 [FOA3]) for the purposes of environmental impact analyses. In this study, BC with a geometric mean diameter (GMD) of 20, 30, and 60 nm and filter diameters of 19 and 35 mm are used to investigate the effect of particle size and sampling variability on SN measurements. For BC with 20 and 30 nm GMD, corresponding to BC emitted by modern aircraft engines, a smaller SN results from a given C BC than is the case for BC with 60 nm GMD, which is more typical of older engines. An updated c...},\n bibtype = {article},\n author = {Stettler, Marc E. J. and Swanson, Jacob J. and Barrett, Steven R. H. and Boies, Adam M.},\n doi = {10.1080/02786826.2013.829908},\n journal = {Aerosol Science and Technology},\n number = {11}\n}
\n
\n\n\n
\n Aircraft emissions of black carbon (BC) contribute to anthropogenic climate forcing and degrade air quality. The smoke number (SN) is the current regulatory measure of aircraft particulate matter emissions and quantifies exhaust plume visibility. Several correlations between SN and the exhaust mass concentration of BC (C BC) have been developed, based on measurements relevant to older aircraft engines. These form the basis of the current standard method used to estimate aircraft BC emissions (First Order Approximation version 3 [FOA3]) for the purposes of environmental impact analyses. In this study, BC with a geometric mean diameter (GMD) of 20, 30, and 60 nm and filter diameters of 19 and 35 mm are used to investigate the effect of particle size and sampling variability on SN measurements. For BC with 20 and 30 nm GMD, corresponding to BC emitted by modern aircraft engines, a smaller SN results from a given C BC than is the case for BC with 60 nm GMD, which is more typical of older engines. An updated c...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Characterization of Combustion Aerosol Produced by a Mini-CAST and Treated in a Catalytic Stripper.\n \n \n \n \n\n\n \n Mamakos, A.; Khalek, I.; Giannelli, R.; and Spears, M.\n\n\n \n\n\n\n Aerosol Science and Technology, 47(8): 927-936. 8 2013.\n \n\n\n\n
\n\n\n\n \n \n \"CharacterizationWebsite\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{\n title = {Characterization of Combustion Aerosol Produced by a Mini-CAST and Treated in a Catalytic Stripper},\n type = {article},\n year = {2013},\n pages = {927-936},\n volume = {47},\n websites = {http://www.tandfonline.com/doi/abs/10.1080/02786826.2013.802762},\n month = {8},\n publisher = { Taylor & Francis Group },\n id = {7067373b-6c6e-3167-b5f8-67c6afd1624c},\n created = {2017-11-20T14:54:13.538Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.511Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n citation_key = {Mamakos2013},\n private_publication = {false},\n abstract = {We characterized the properties of combustion aerosol produced at different operating conditions of a mini-CAST burner that was treated in a Catalytic Stripper (CS) operating at 300°C. The goal was to establish a methodology for the production of soot particles resembling those emitted from internal combustion engines. Thermo-optical analysis of samples collected on Quartz filters revealed that the particles contained semi-volatile material that survived the CS. The amount of semi-volatile species strongly depended on the operating conditions ranging from less than 10% to as high as 30% of the particle mass. The mini-CAST operating conditions were also found to have a strong effect on the effective particle density (ρe ). The ρe , for example, ranged from as low as 0.3 to 1.05 g/cm3 for mondisperse 80 nm particles, although the mass-mobility exponent remained relatively constant (2.1–2.25). These differences are indicative of differences in the primary particle diameter, which was estimated to range betwe...},\n bibtype = {article},\n author = {Mamakos, Athanasios and Khalek, Imad and Giannelli, Robert and Spears, Matthew},\n doi = {10.1080/02786826.2013.802762},\n journal = {Aerosol Science and Technology},\n number = {8}\n}
\n
\n\n\n
\n We characterized the properties of combustion aerosol produced at different operating conditions of a mini-CAST burner that was treated in a Catalytic Stripper (CS) operating at 300°C. The goal was to establish a methodology for the production of soot particles resembling those emitted from internal combustion engines. Thermo-optical analysis of samples collected on Quartz filters revealed that the particles contained semi-volatile material that survived the CS. The amount of semi-volatile species strongly depended on the operating conditions ranging from less than 10% to as high as 30% of the particle mass. The mini-CAST operating conditions were also found to have a strong effect on the effective particle density (ρe ). The ρe , for example, ranged from as low as 0.3 to 1.05 g/cm3 for mondisperse 80 nm particles, although the mass-mobility exponent remained relatively constant (2.1–2.25). These differences are indicative of differences in the primary particle diameter, which was estimated to range betwe...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n An Aerosolization Method for Characterizing Particle Contaminants in Diesel Fuel.\n \n \n \n \n\n\n \n Xiao, K.; Pei, C.; Swanson, J.; Kittelson, D.; and Pui, D.\n\n\n \n\n\n\n SAE International Journal of Fuels and Lubricants, 6(3): 2013-01-2668. 10 2013.\n \n\n\n\n
\n\n\n\n \n \n \"AnWebsite\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{\n title = {An Aerosolization Method for Characterizing Particle Contaminants in Diesel Fuel},\n type = {article},\n year = {2013},\n pages = {2013-01-2668},\n volume = {6},\n websites = {http://papers.sae.org/2013-01-2668/},\n month = {10},\n day = {14},\n id = {74587286-40b4-30cc-b66b-e6d5e7b6418e},\n created = {2017-11-20T14:54:13.562Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.562Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n bibtype = {article},\n author = {Xiao, Kai and Pei, Chenxing and Swanson, Jacob and Kittelson, David and Pui, David},\n doi = {10.4271/2013-01-2668},\n journal = {SAE International Journal of Fuels and Lubricants},\n number = {3}\n}
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Microstructural and loading characteristics of diesel aggregate cakes.\n \n \n \n \n\n\n \n Liu, J.; Swanson, J., J.; Kittelson, D., B.; Pui, D., Y.; and Wang, J.\n\n\n \n\n\n\n Powder Technology, 241: 244-251. 6 2013.\n \n\n\n\n
\n\n\n\n \n \n \"MicrostructuralWebsite\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 \n \n \n \n \n \n\n\n\n
\n
@article{\n title = {Microstructural and loading characteristics of diesel aggregate cakes},\n type = {article},\n year = {2013},\n keywords = {Catalytic stripper,Filter loading,PTFE filter,Porosity,Pressure drop},\n pages = {244-251},\n volume = {241},\n websites = {http://www.sciencedirect.com/science/article/pii/S0032591013002064?via%3Dihub},\n month = {6},\n publisher = {Elsevier},\n day = {1},\n id = {559ea540-6821-3583-9643-a9feda38dc83},\n created = {2017-11-20T14:54:13.673Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.676Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {Aggregates consisting of nanometer sized primary particles are ubiquitous in the atmosphere. Sources of these particles such as combustion and nanoparticle manufacturing require control by filtration. To further investigate how loading processes are impacted by aggregate characteristics, we evaluated the loading of diesel particulate matter (DPM) on membrane filters at ambient temperature. Experimental results indicated that the filtration face velocity, engine load, and the usage of a catalytic stripper (CS) impact the characteristics of the resulting aggregate deposit or "cake". Mode aggregate size and primary particle size were slightly higher at 75% load (63.8 and 44.9. nm, respectively) compared to 37.5% load (57.3 and 40.9. nm, respectively). In-situ direct measurements of cake thickness enabled calculations of the porosity of the aggregate cakes which indicated that porosity varied between 0.93 and 0.97. The CS removed condensed semi-volatile organic material from the aggregates, which in turn impacted the calculation of porosity. Modeling based on Endo, Chen and Pui (Powder Technology, 1998, 98, 241-249) [1] of the pressure drop was performed based on the observation that cakes formed by the aggregates could be regarded as formed by a uniform collection of primary particles and once deposited, the aggregates were indistinguishable. The sensitivity of this model to various analytical expressions for the void function depending on the porosity was investigated. Results showed that available expressions vary greatly, although some of them provided satisfactory fitting for the experimental data. © 2013 Elsevier B.V.},\n bibtype = {article},\n author = {Liu, Jingxian and Swanson, Jacob J. and Kittelson, David B. and Pui, David Y.H. and Wang, Jing},\n doi = {10.1016/j.powtec.2013.03.028},\n journal = {Powder Technology}\n}
\n
\n\n\n
\n Aggregates consisting of nanometer sized primary particles are ubiquitous in the atmosphere. Sources of these particles such as combustion and nanoparticle manufacturing require control by filtration. To further investigate how loading processes are impacted by aggregate characteristics, we evaluated the loading of diesel particulate matter (DPM) on membrane filters at ambient temperature. Experimental results indicated that the filtration face velocity, engine load, and the usage of a catalytic stripper (CS) impact the characteristics of the resulting aggregate deposit or \"cake\". Mode aggregate size and primary particle size were slightly higher at 75% load (63.8 and 44.9. nm, respectively) compared to 37.5% load (57.3 and 40.9. nm, respectively). In-situ direct measurements of cake thickness enabled calculations of the porosity of the aggregate cakes which indicated that porosity varied between 0.93 and 0.97. The CS removed condensed semi-volatile organic material from the aggregates, which in turn impacted the calculation of porosity. Modeling based on Endo, Chen and Pui (Powder Technology, 1998, 98, 241-249) [1] of the pressure drop was performed based on the observation that cakes formed by the aggregates could be regarded as formed by a uniform collection of primary particles and once deposited, the aggregates were indistinguishable. The sensitivity of this model to various analytical expressions for the void function depending on the porosity was investigated. Results showed that available expressions vary greatly, although some of them provided satisfactory fitting for the experimental data. © 2013 Elsevier B.V.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Filtration efficiency and pressure drop of miniature diesel particulate filters.\n \n \n \n \n\n\n \n Swanson, J.; Watts, W.; Kittelson, D.; Newman, R.; and Ziebarth, R.\n\n\n \n\n\n\n Aerosol Science and Technology, 47(4): 452-461. 4 2013.\n \n\n\n\n
\n\n\n\n \n \n \"FiltrationWebsite\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{\n title = {Filtration efficiency and pressure drop of miniature diesel particulate filters},\n type = {article},\n year = {2013},\n pages = {452-461},\n volume = {47},\n websites = {http://www.tandfonline.com/doi/abs/10.1080/02786826.2012.763087},\n month = {4},\n publisher = { Taylor & Francis Group },\n id = {7cc14cc2-b355-3d41-afdc-364331c5dad1},\n created = {2017-11-20T14:54:13.679Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.960Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {A method was developed to evaluate miniature diesel particulate filters (DPFs). To validate the performance of the instrumentation and test apparatus, measurements were made using silicon carbide (SiC) and cordierite miniature filters with representative microstructures. Filtration efficiency (FE), the most penetrating particle size (MPPS), and pressure drop were measured for catalyzed and uncatalyzed advanced ceramic material (ACM) acicular mullite and representative commercial filters to determine the impact of substrate morphology, the formation of a soot cake, and the presence of a catalyst coating on filtration properties. FE measurements demonstrated that filter geometry and microstructure significantly influence initial filtration performance. ACM filters had high initial FE and the MPPS near ∼200 nm. Reduction of the ACM pore size in the absence of a reduction in porosity increased initial FE evenmore, but its influence on MPPS was not resolvable. The presence of a catalyst and washcoat on the ACM increased the pressure drop but increased initial FE and reduced MPPS to <100 nm. The addition of a washcoat allowed the rapid buildup of a soot cake, which resulted in a more rapid rate of increase in FE compared to uncatalyzed ACM. The similarity in the ACM and cordierite soot cakes after a long loading time is consistent with theory that suggests the formation of the soot cake depends primarily on the Péclet (Pe) number, which is influenced only by macroscopic filter geometry and prevailing test conditions. Copyright © American Association for Aerosol Research.},\n bibtype = {article},\n author = {Swanson, J. and Watts, W. and Kittelson, D. and Newman, R. and Ziebarth, R.},\n doi = {10.1080/02786826.2012.763087},\n journal = {Aerosol Science and Technology},\n number = {4}\n}
\n
\n\n\n
\n A method was developed to evaluate miniature diesel particulate filters (DPFs). To validate the performance of the instrumentation and test apparatus, measurements were made using silicon carbide (SiC) and cordierite miniature filters with representative microstructures. Filtration efficiency (FE), the most penetrating particle size (MPPS), and pressure drop were measured for catalyzed and uncatalyzed advanced ceramic material (ACM) acicular mullite and representative commercial filters to determine the impact of substrate morphology, the formation of a soot cake, and the presence of a catalyst coating on filtration properties. FE measurements demonstrated that filter geometry and microstructure significantly influence initial filtration performance. ACM filters had high initial FE and the MPPS near ∼200 nm. Reduction of the ACM pore size in the absence of a reduction in porosity increased initial FE evenmore, but its influence on MPPS was not resolvable. The presence of a catalyst and washcoat on the ACM increased the pressure drop but increased initial FE and reduced MPPS to <100 nm. The addition of a washcoat allowed the rapid buildup of a soot cake, which resulted in a more rapid rate of increase in FE compared to uncatalyzed ACM. The similarity in the ACM and cordierite soot cakes after a long loading time is consistent with theory that suggests the formation of the soot cake depends primarily on the Péclet (Pe) number, which is influenced only by macroscopic filter geometry and prevailing test conditions. Copyright © American Association for Aerosol Research.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Use of a Catalytic Stripper as an Alternative to the Original PMP Measurement Protocol.\n \n \n \n \n\n\n \n Ntziachristos, L.; Amanatidis, S.; Samaras, Z.; Giechaskiel, B.; and Bergmann, A.\n\n\n \n\n\n\n SAE International Journal of Fuels and Lubricants, 6(2): 2013-01-1563. 4 2013.\n \n\n\n\n
\n\n\n\n \n \n \"UseWebsite\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{\n title = {Use of a Catalytic Stripper as an Alternative to the Original PMP Measurement Protocol},\n type = {article},\n year = {2013},\n pages = {2013-01-1563},\n volume = {6},\n websites = {http://papers.sae.org/2013-01-1563/},\n month = {4},\n day = {8},\n id = {ca9a94d0-fd7c-3834-a41b-6589c8a8e103},\n created = {2017-11-20T14:54:13.711Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.711Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n bibtype = {article},\n author = {Ntziachristos, Leonidas and Amanatidis, Stavros and Samaras, Zissis and Giechaskiel, Barouch and Bergmann, Alexander},\n doi = {10.4271/2013-01-1563},\n journal = {SAE International Journal of Fuels and Lubricants},\n number = {2}\n}
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n A Miniature Catalytic Stripper for Particles Less Than 23 Nanometers.\n \n \n \n \n\n\n \n Swanson, J.; Kittelson, D.; Giechaskiel, B.; Bergmann, A.; and Twigg, M.\n\n\n \n\n\n\n SAE International Journal of Fuels and Lubricants, 6(2): 2013-01-1570. 4 2013.\n \n\n\n\n
\n\n\n\n \n \n \"AWebsite\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 2 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{\n title = {A Miniature Catalytic Stripper for Particles Less Than 23 Nanometers},\n type = {article},\n year = {2013},\n pages = {2013-01-1570},\n volume = {6},\n websites = {http://papers.sae.org/2013-01-1570/},\n month = {4},\n day = {8},\n id = {d36da476-4da7-30e1-a969-5e4d65befbca},\n created = {2017-11-20T14:54:13.994Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.994Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n bibtype = {article},\n author = {Swanson, Jacob and Kittelson, David and Giechaskiel, Barouch and Bergmann, Alexander and Twigg, Martyn},\n doi = {10.4271/2013-01-1570},\n journal = {SAE International Journal of Fuels and Lubricants},\n number = {2}\n}
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2012\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Comparison of Methods for Online Measurement of Diesel Particulate Matter.\n \n \n \n \n\n\n \n Liu, Z.; Swanson, J.; Kittelson, D., B.; and Pui, D., Y., H.\n\n\n \n\n\n\n Environmental Science & Technology, 46(11): 6127-6133. 6 2012.\n \n\n\n\n
\n\n\n\n \n \n \"ComparisonWebsite\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{\n title = {Comparison of Methods for Online Measurement of Diesel Particulate Matter},\n type = {article},\n year = {2012},\n pages = {6127-6133},\n volume = {46},\n websites = {http://pubs.acs.org/doi/10.1021/es3003537},\n month = {6},\n publisher = {American Chemical Society},\n day = {5},\n id = {bfc57203-f1e0-327a-a835-50aae341e51d},\n created = {2017-11-20T14:54:13.215Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.215Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Gravimetric analysis is the regulatory method for diesel particulate mass measurement. Because of issues such as adsorption/volatilization artifacts, it faces obstacles in measuring ultra low level emissions from modern diesel engines. Alternative methods of suspended particle mass measurement have been developed that show improvements in time resolution, sensitivity, and accuracy. Three size-resolved methods were considered here. Two methods rely on converting number size distributions obtained using a scanning mobility particle sizer (SMPS). Conversion techniques were based on effective density measurements and the Lall–Friedlander aggregate model. The third method employs the Universal Nanoparticle Analyzer (UNPA) to measure the aggregate size distribution from which mass is calculated. Results were compared with mass concentrations obtained using gravimetric analysis. The effective density conversion resulted in mass concentrations that were highly correlated (R2 >0.99) with filter mass. The ratios to...},\n bibtype = {article},\n author = {Liu, Zhun and Swanson, Jacob and Kittelson, David B. and Pui, David Y. H.},\n doi = {10.1021/es3003537},\n journal = {Environmental Science & Technology},\n number = {11}\n}
\n
\n\n\n
\n Gravimetric analysis is the regulatory method for diesel particulate mass measurement. Because of issues such as adsorption/volatilization artifacts, it faces obstacles in measuring ultra low level emissions from modern diesel engines. Alternative methods of suspended particle mass measurement have been developed that show improvements in time resolution, sensitivity, and accuracy. Three size-resolved methods were considered here. Two methods rely on converting number size distributions obtained using a scanning mobility particle sizer (SMPS). Conversion techniques were based on effective density measurements and the Lall–Friedlander aggregate model. The third method employs the Universal Nanoparticle Analyzer (UNPA) to measure the aggregate size distribution from which mass is calculated. Results were compared with mass concentrations obtained using gravimetric analysis. The effective density conversion resulted in mass concentrations that were highly correlated (R2 >0.99) with filter mass. The ratios to...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Crayford: SAMPLE III: Contribution to aircraft engine... - Google Scholar.\n \n \n \n \n\n\n \n Crayford, A.; and Johnson, M.\n\n\n \n\n\n\n . 2012.\n \n\n\n\n
\n\n\n\n \n \n \"Crayford:Website\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{\n title = {Crayford: SAMPLE III: Contribution to aircraft engine... - Google Scholar},\n type = {article},\n year = {2012},\n websites = {https://easa.europa.eu/system/files/dfu/SAMPLE III-SC02 Final report.pdf},\n id = {1605c396-aa6f-3fa9-bbc5-c4078e409ae0},\n created = {2017-11-20T14:54:13.890Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.951Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n bibtype = {article},\n author = {Crayford, A. and Johnson, M.}\n}
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2011\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Development of a Solid Exhaust Particle Number Measurement System Using a Catalytic Stripper Technology.\n \n \n \n \n\n\n \n Khalek, I., A.; and Bougher, T.\n\n\n \n\n\n\n SAE International Journal of Engines, 4(1): 2011-01-0635. 4 2011.\n \n\n\n\n
\n\n\n\n \n \n \"DevelopmentWebsite\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 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{\n title = {Development of a Solid Exhaust Particle Number Measurement System Using a Catalytic Stripper Technology},\n type = {article},\n year = {2011},\n pages = {2011-01-0635},\n volume = {4},\n websites = {http://papers.sae.org/2011-01-0635/},\n month = {4},\n day = {12},\n id = {0b2d47b3-2ca6-35ad-8638-037fef854682},\n created = {2017-11-20T14:54:14.105Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:14.105Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n bibtype = {article},\n author = {Khalek, Imad A. and Bougher, Thomas},\n doi = {10.4271/2011-01-0635},\n journal = {SAE International Journal of Engines},\n number = {1}\n}
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2010\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Evaluation of thermal denuder and catalytic stripper methods for solid particle measurements.\n \n \n \n \n\n\n \n Swanson, J.; and Kittelson, D.\n\n\n \n\n\n\n Journal of Aerosol Science, 41(12): 1113-1122. 12 2010.\n \n\n\n\n
\n\n\n\n \n \n \"EvaluationWebsite\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 1 download\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
@article{\n title = {Evaluation of thermal denuder and catalytic stripper methods for solid particle measurements},\n type = {article},\n year = {2010},\n keywords = {Diesel exhaust,Organic carbon,Real-time methods,Sulfuric acid},\n pages = {1113-1122},\n volume = {41},\n websites = {http://www.sciencedirect.com/science/article/pii/S0021850210002004?via%3Dihub},\n month = {12},\n publisher = {Pergamon},\n day = {1},\n id = {f7b6e5b3-c671-3048-b616-57b388203c9a},\n created = {2017-11-20T14:54:13.317Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:39.191Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {The objective of this study was to compare two methods that are used to separate the solid and volatile components of an aerosol: the thermal denuder (TD) and catalytic stripper (CS). We challenged the TD and CS with atmospheric and laboratory generated aerosols. Laboratory generated particles were composed of tetracosane, tetracosane and sulfuric acid, and dioctyl sebacate and sulfuric acid. These compositions were chosen because they roughly simulate the composition of nanoparticles found in Diesel exhaust. The TD method produced semi-volatile particle artifacts due to the incomplete removal of evaporated compounds that nucleated and formed particles and solid particle artifacts that formed during treatment of the aerosol by the TD. Our results suggest that the differences in these methods will lead to different conclusions regarding the presence or absence, size, and concentration of solid particles in Diesel exhaust. © 2010 Elsevier Ltd.},\n bibtype = {article},\n author = {Swanson, Jacob and Kittelson, David},\n doi = {10.1016/j.jaerosci.2010.09.003},\n journal = {Journal of Aerosol Science},\n number = {12}\n}
\n
\n\n\n
\n The objective of this study was to compare two methods that are used to separate the solid and volatile components of an aerosol: the thermal denuder (TD) and catalytic stripper (CS). We challenged the TD and CS with atmospheric and laboratory generated aerosols. Laboratory generated particles were composed of tetracosane, tetracosane and sulfuric acid, and dioctyl sebacate and sulfuric acid. These compositions were chosen because they roughly simulate the composition of nanoparticles found in Diesel exhaust. The TD method produced semi-volatile particle artifacts due to the incomplete removal of evaporated compounds that nucleated and formed particles and solid particle artifacts that formed during treatment of the aerosol by the TD. Our results suggest that the differences in these methods will lead to different conclusions regarding the presence or absence, size, and concentration of solid particles in Diesel exhaust. © 2010 Elsevier Ltd.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Effect of Organometallic Fuel Additives on Nanoparticle Emissions from a Gasoline Passenger Car.\n \n \n \n \n\n\n \n Gidney, J., T.; Twigg, M., V.; and Kittelson, D., B.\n\n\n \n\n\n\n Environmental Science & Technology, 44(7): 2562-2569. 4 2010.\n \n\n\n\n
\n\n\n\n \n \n \"EffectWebsite\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{\n title = {Effect of Organometallic Fuel Additives on Nanoparticle Emissions from a Gasoline Passenger Car},\n type = {article},\n year = {2010},\n pages = {2562-2569},\n volume = {44},\n websites = {http://pubs.acs.org/doi/abs/10.1021/es901868c},\n month = {4},\n publisher = { American Chemical Society},\n id = {67eee5da-3412-369f-b682-721fe700e249},\n created = {2017-11-20T14:54:13.792Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.792Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Particle size measurements were performed on the exhaust of a car operating on a chassis dynamometer fuelled with standard gasoline and gasoline containing low levels of Pb, Fe, and Mn organometallic additives. When additives were present there was a distinct nucleation mode consisting primarily of sub-10 nm nanoparticles. At equal molar dosing Mn and Fe gave similar nanoparticle concentrations at the tailpipe, whereas Pb gave a considerably lower concentration. A catalytic stripper was used to remove the organic component of these particles and revealed that they were mainly solid and, because of their association with inorganic additives, presumably inorganic. Solid nucleation mode nanoparticles of similar size and concentration to those observed here from a gasoline engine with Mn and Fe additives have also been observed from modern heavy-duty diesel engines without aftertreatment at idle, but these solid particles are a small fraction of the primarily volatile nucleation mode particles emitted. The so...},\n bibtype = {article},\n author = {Gidney, Jeremy T. and Twigg, Martyn V. and Kittelson, David B.},\n doi = {10.1021/es901868c},\n journal = {Environmental Science & Technology},\n number = {7}\n}
\n
\n\n\n
\n Particle size measurements were performed on the exhaust of a car operating on a chassis dynamometer fuelled with standard gasoline and gasoline containing low levels of Pb, Fe, and Mn organometallic additives. When additives were present there was a distinct nucleation mode consisting primarily of sub-10 nm nanoparticles. At equal molar dosing Mn and Fe gave similar nanoparticle concentrations at the tailpipe, whereas Pb gave a considerably lower concentration. A catalytic stripper was used to remove the organic component of these particles and revealed that they were mainly solid and, because of their association with inorganic additives, presumably inorganic. Solid nucleation mode nanoparticles of similar size and concentration to those observed here from a gasoline engine with Mn and Fe additives have also been observed from modern heavy-duty diesel engines without aftertreatment at idle, but these solid particles are a small fraction of the primarily volatile nucleation mode particles emitted. The so...\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Alternatives to the Gravimetric Method for Quantification of Diesel Particulate Matter near the Lower Level of Detection.\n \n \n \n \n\n\n \n Swanson, J.; Kittelson, D.; Pui, D.; and Watts, W.\n\n\n \n\n\n\n Journal of the Air & Waste Management Association, 60(10): 1177-1191. 10 2010.\n \n\n\n\n
\n\n\n\n \n \n \"AlternativesWebsite\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{\n title = {Alternatives to the Gravimetric Method for Quantification of Diesel Particulate Matter near the Lower Level of Detection},\n type = {article},\n year = {2010},\n pages = {1177-1191},\n volume = {60},\n websites = {https://www.tandfonline.com/doi/full/10.3155/1047-3289.60.10.1177},\n month = {10},\n publisher = { Taylor & Francis Group },\n day = {24},\n id = {743b6127-354c-325d-a8e6-85f293d53a51},\n created = {2017-11-20T14:54:13.947Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.947Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Abstract This paper is part of the Journal of the Air & Waste Management Association’s 2010 special issue on combustion aerosol measurements. The issue is a combination of papers that synthesize and evaluate ideas and perspectives that were presented by experts at a series of workshops sponsored by the Coordinating Research Council that aimed to evaluate the current and future status of diesel particulate matter (DPM) measurement. Measurement of DPM is a complex issue with many stakeholders, including air quality management and enforcement agencies, engine manufacturers, health experts, and climatologists. Adoption of the U.S. Environmental Protection Agency 2007 heavy-duty engine DPM standards posed a unique challenge to engine manufacturers. The new standards reduced DPM emissions to the point that improvements to the gravimetric method were required to increase the accuracy and the sensitivity of the measurement. Despite these improvements, the method still has shortcomings. The objectives of this pape...},\n bibtype = {article},\n author = {Swanson, Jacob and Kittelson, David and Pui, David and Watts, Winthrop},\n doi = {10.3155/1047-3289.60.10.1177},\n journal = {Journal of the Air & Waste Management Association},\n number = {10}\n}
\n
\n\n\n
\n Abstract This paper is part of the Journal of the Air & Waste Management Association’s 2010 special issue on combustion aerosol measurements. The issue is a combination of papers that synthesize and evaluate ideas and perspectives that were presented by experts at a series of workshops sponsored by the Coordinating Research Council that aimed to evaluate the current and future status of diesel particulate matter (DPM) measurement. Measurement of DPM is a complex issue with many stakeholders, including air quality management and enforcement agencies, engine manufacturers, health experts, and climatologists. Adoption of the U.S. Environmental Protection Agency 2007 heavy-duty engine DPM standards posed a unique challenge to engine manufacturers. The new standards reduced DPM emissions to the point that improvements to the gravimetric method were required to increase the accuracy and the sensitivity of the measurement. Despite these improvements, the method still has shortcomings. The objectives of this pape...\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2007\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Comparing measurements of carbon in diesel exhaust aerosols using the aethalometer, NIOSH method 5040, and SMPS.\n \n \n \n \n\n\n \n Ng, I., P.; Ma, H.; Kittelson, D.; and Miller, A.\n\n\n \n\n\n\n In SAE Technical Papers, 4 2007. \n \n\n\n\n
\n\n\n\n \n \n \"ComparingWebsite\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
@inproceedings{\n title = {Comparing measurements of carbon in diesel exhaust aerosols using the aethalometer, NIOSH method 5040, and SMPS},\n type = {inproceedings},\n year = {2007},\n websites = {http://papers.sae.org/2007-01-0334/},\n month = {4},\n day = {16},\n id = {8bc4a189-4a9e-3ff9-9fbd-e8407f04d358},\n created = {2017-11-20T14:54:13.316Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:39.148Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {Combustion aerosols consist mainly of elemental and organic carbon (EC and OC). Since EC strongly absorbs light and thus affects atmospheric visibility and radiation balance, there is great interest in its measurement. To this end, the National Institute for Occupational Safety and Health (NIOSH) published a standard method to determine the mass of EC and OC on filter samples. Another common method of measuring carbon in aerosols is the aethalometer, which uses light extinction to measure "black carbon" or BC, which is considered to approximate EC. A third method sometimes used for estimating carbon in submicron combustion aerosols, is to measure particle size distributions using a scanning mobility particle sizer (SMPS) and calculate mass using the assumptions that the particles are spherical, carbonaceous and of known density. With an eye toward evaluating the use of these methods for measuring carbon in freshly emitted diesel aerosols, the objective of this paper is to compare the results among these methods and to investigate the role of OC when measuring the mass concentration of elemental carbon (or black carbon) in aerosols emitted by a medium-duty Diesel engine. Results indicate that the aethalometer response correlates well with the data from filter samples (R2 = 0.99), with a very slight positive interference at high levels of OC. Copyright © 2007 SAE International.},\n bibtype = {inproceedings},\n author = {Ng, Iam Pou and Ma, Hongbin and Kittelson, David and Miller, Art},\n doi = {10.4271/2007-01-0334},\n booktitle = {SAE Technical Papers}\n}
\n
\n\n\n
\n Combustion aerosols consist mainly of elemental and organic carbon (EC and OC). Since EC strongly absorbs light and thus affects atmospheric visibility and radiation balance, there is great interest in its measurement. To this end, the National Institute for Occupational Safety and Health (NIOSH) published a standard method to determine the mass of EC and OC on filter samples. Another common method of measuring carbon in aerosols is the aethalometer, which uses light extinction to measure \"black carbon\" or BC, which is considered to approximate EC. A third method sometimes used for estimating carbon in submicron combustion aerosols, is to measure particle size distributions using a scanning mobility particle sizer (SMPS) and calculate mass using the assumptions that the particles are spherical, carbonaceous and of known density. With an eye toward evaluating the use of these methods for measuring carbon in freshly emitted diesel aerosols, the objective of this paper is to compare the results among these methods and to investigate the role of OC when measuring the mass concentration of elemental carbon (or black carbon) in aerosols emitted by a medium-duty Diesel engine. Results indicate that the aethalometer response correlates well with the data from filter samples (R2 = 0.99), with a very slight positive interference at high levels of OC. Copyright © 2007 SAE International.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Sampling System for Solid and Volatile Exhaust Particle Size, Number, and Mass Emissions.\n \n \n \n \n\n\n \n Khalek, I., A.\n\n\n \n\n\n\n In 4 2007. \n \n\n\n\n
\n\n\n\n \n \n \"SamplingWebsite\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
@inproceedings{\n title = {Sampling System for Solid and Volatile Exhaust Particle Size, Number, and Mass Emissions},\n type = {inproceedings},\n year = {2007},\n websites = {http://papers.sae.org/2007-01-0307/},\n month = {4},\n day = {16},\n id = {e045e0fb-2794-3bce-8d67-28b98c272fa2},\n created = {2017-11-20T14:54:13.757Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.757Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n bibtype = {inproceedings},\n author = {Khalek, Imad A.},\n doi = {10.4271/2007-01-0307}\n}
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2006\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Characteristics of SME Biodiesel-Fueled Diesel Particle Emissions and the Kinetics of Oxidation.\n \n \n \n \n\n\n \n Heejung Jung, †.; David B. Kittelson, *., a.; and Zachariah§, M., R.\n\n\n \n\n\n\n . 2006.\n \n\n\n\n
\n\n\n\n \n \n \"CharacteristicsWebsite\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{\n title = {Characteristics of SME Biodiesel-Fueled Diesel Particle Emissions and the Kinetics of Oxidation},\n type = {article},\n year = {2006},\n websites = {http://pubs.acs.org/doi/abs/10.1021/es0515452},\n publisher = { American Chemical Society },\n id = {e5464f4b-df84-3e88-bad0-294c51c81dcd},\n created = {2017-11-20T14:54:13.081Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.081Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n abstract = {Biodiesel is one of the most promising alternative diesel fuels. As diesel emission regulations have become more stringent, the diesel particulate filter (DPF) has become an essential part of the aftertreatment system. Knowledge of kinetics of exhaust particle oxidation for alternative diesel fuels is useful in estimating the change in regeneration behavior of a DPF with such fuels. This study examines the characteristics of diesel particulate emissions as well as kinetics of particle oxidation using a 1996 John Deere T04045TF250 off-highway engine and 100% soy methyl ester (SME) biodiesel (B100) as fuel. Compared to standard D2 fuel, this B100 reduced particle size, number, and volume in the accumulation mode where most of the particle mass is found. At 75% load, number decreased by 38%, DGN decreased from 80 to 62 nm, and volume decreased by 82%. Part of this decrease is likely associated with the fact that the particles were more easily oxidized. Arrhenius parameters for the biodiesel fuel showed a 2−3...},\n bibtype = {article},\n author = {Heejung Jung, † and David B. Kittelson, *,‡ and and Zachariah§, Michael R.},\n doi = {10.1021/ES0515452}\n}
\n
\n\n\n
\n Biodiesel is one of the most promising alternative diesel fuels. As diesel emission regulations have become more stringent, the diesel particulate filter (DPF) has become an essential part of the aftertreatment system. Knowledge of kinetics of exhaust particle oxidation for alternative diesel fuels is useful in estimating the change in regeneration behavior of a DPF with such fuels. This study examines the characteristics of diesel particulate emissions as well as kinetics of particle oxidation using a 1996 John Deere T04045TF250 off-highway engine and 100% soy methyl ester (SME) biodiesel (B100) as fuel. Compared to standard D2 fuel, this B100 reduced particle size, number, and volume in the accumulation mode where most of the particle mass is found. At 75% load, number decreased by 38%, DGN decreased from 80 to 62 nm, and volume decreased by 82%. Part of this decrease is likely associated with the fact that the particles were more easily oxidized. Arrhenius parameters for the biodiesel fuel showed a 2−3...\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2005\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Determination of the aerosol size distribution from the mobility distribution of the charged fraction of aerosols.\n \n \n \n \n\n\n \n Allen, M., D.; Raabe, O., G.; Chen, D., R.; Pui, D., Y., H.; Hummes, D.; Fissan, H.; Quant, F., R.; Sem, G., J.; Gramotnev, D., K.; Gramotnev, G.; Hering, S., V.; Stolzenburg, M., R.; Quant, F., R.; Oberreit, D., R.; Keady, P., B.; Hoppel, W., A.; Iida, K.; Stolzenburg, M., R.; McMurry, P., H.; Smith, J., N.; Quant, F., R.; Oberreit, D., R.; Keady, P., B.; Eiguren-Fernandez, A.; Lewis, G., S.; Kreisberg, N., M.; Hering, S., V.; Knutson, E., O.; Whitby, K., T.; Liu, B., Y., H.; Pui, D., Y., H.; McMurry, P., H.; Stolzenburg, M., R.; McMurry, P., H.; Swihart, M., T.; Wang, S., C.; Flagan, R., C.; Considerations, G.; Ii, P., H., M.; Hering, S., V.; Stolzenburg, M., R.; Jiang, J.; Attoui, M.; Heim, M.; Brunelli, N., a.; McMurry, P., H.; Kasper, G.; Flagan, R., C.; Giapis, K.; Mouret, G.; Scientific, E.; Company, P.; Kim, S., H.; Woo, K., S.; Liu, B., Y., H.; Zachariah, M., R.; Kim, J., H.; Mulholland, G., W.; Kukuck, S., R.; Kinney, P., D.; Pui, D., Y., H.; Mulholland, G., W.; Bryner, N., P.; Taylor, P.; Kousaka, Y.; Okuyama, K.; Li, W.; Li, L.; Chen, D., R.; Li, W.; Li, L.; Chen, D., R.; Rader, D., J.; McMurry, P., H.; Manual, S.; Vemury, S.; Pratsinis, S., E.; Feldpausch, M.; Helsper, C.; and Wiedensohler, A.\n\n\n \n\n\n\n Journal of Aerosol Science, 13(2): 659-672. 2005.\n \n\n\n\n
\n\n\n\n \n \n \"DeterminationWebsite\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 \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 \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
@article{\n title = {Determination of the aerosol size distribution from the mobility distribution of the charged fraction of aerosols},\n type = {article},\n year = {2005},\n keywords = {2005,Aerosol evolution Nanoparticle transformation Busy,CONDENSATION BUTANOL AEROSOLS PARTICLES LAMINAR fl,Fine particles Size distributions Aerosol sampling,Nanoparticle Nanocrystal Aerosol,PARTICLES (Nuclear physics) LAMINAR flow AEROSOLS,Slip Correction SMPS Spreedsheet,accepted,aerosol generator,atomiz-,boltzmann theory,charge population balance,condensation nuclei counters,diffusion flame,electrical mobility,electron,electrospray,equation,ers,january 5,knudsen,nanosized aerosols,negative ion,nmda,number,particle size,particles,polystyrene latex,polystyrene latex spheres,positive ion,slip correction factor,tandem differential mobility analyzer},\n pages = {659-672},\n volume = {13},\n websites = {http://www.sciencedirect.com/science/article/B6VRY-4846T0J-3/2/b1f3ba085442f7502062d37be4352423,http://www.sciencedirect.com/science/article/B6VH3-3YVD9S9-5/2/930e22b064f983e0db81a03ab6ae8851,http://www.sciencedirect.com/science/article/B6WHR-4CTN1VV-8T/2},\n id = {87c971cc-f6c8-3861-84a0-8816d25ff4a5},\n created = {2017-11-17T15:14:59.095Z},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-17T15:14:59.095Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n source_type = {JOUR},\n notes = {<b>From Duplicate 19 (<i>Vapor-phase synthesis of nanoparticles</i> - Swihart, Mark T)<br/></b><br/>doi: DOI: 10.1016/S1359-0294(03)00007-4<br/><br/><b>From Duplicate 24 (<i>An Ultrafine, Water-Based Condensation Particle Counter and its Evaluation under Field Conditions</i> - Iida, Kenjiro; Stolzenburg, Mark R; McMurry, Peter H; Smith, James N; Quant, Frederick R; Oberreit, Derek R; Keady, Patricia B; Eiguren-Fernandez, Arantza; Lewis, Gregory S; Kreisberg, Nathan M; Hering, Susanne V)<br/></b><br/>Article<br/>Accession Number: 34320035; Iida, Kenjiro 1 Stolzenburg, Mark R. 1 McMurry, Peter H. 1 Smith, James N. 2 Quant, Frederick R. 3 Oberreit, Derek R. 3 Keady, Patricia B. 3 Eiguren-Fernandez, Arantza 4 Lewis, Gregory S. 5 Kreisberg, Nathan M. 5 Hering, Susanne V. 5; Email Address: susanne@aerosol.us; Affiliation: 1: Mechanical Engineering, University of Minnesota, Minneapolis, Minnesota, USA 2: National Center for Atmospheric Research, Boulder, Colorado, USA 3: Quant Technologies LLC, Blaine, Minnesota, USA 4: University of California, Los Angeles, California, USA 5: Aerosol Dynamics Inc., Berkeley, California, USA; Source Info: Oct2008, Vol. 42 Issue 10, p862; Subject Term: CONDENSATION; Subject Term: BUTANOL; Subject Term: AEROSOLS; Subject Term: PARTICLES; Subject Term: LAMINAR flow; Subject Term: ATMOSPHERIC aerosols; Subject Term: FLUID dynamics; Subject Term: BOULDER (Colo.); Subject Term: COLORADO; Number of Pages: 10p; Illustrations: 2 diagrams, 4 graphs; Document Type: Article<br/><br/><b>From Duplicate 26 (<i>A Laminar-Flow, Water-Based Condensation Particle Counter (WCPC)</i> - Hering, Susanne V; Stolzenburg, Mark R; Quant, Frederick R; Oberreit, Derek R; Keady, Patricia B)<br/></b><br/>Article<br/>Accession Number: 21029298; Hering, Susanne V. 1 Stolzenburg, Mark R. 1 Quant, Frederick R. 2 Oberreit, Derek R. 2 Keady, Patricia B. 2; Affiliation: 1: Aerosol Dynamics Inc., Berkeley, California, USA 2: Quant Technologies, LLC., Blaine, Minnesota, USA; Source Info: Jul2005, Vol. 39 Issue 7, p659; Subject Term: PARTICLES (Nuclear physics); Subject Term: LAMINAR flow; Subject Term: AEROSOLS; Subject Term: FLUID dynamics; Subject Term: CONDENSATION; Number of Pages: 14p; Illustrations: 1 chart, 1 diagram, 12 graphs; Document Type: Article<br/><br/><b>From Duplicate 27 (<i>A new mechanism of aerosol evolution near a busy road: fragmentation of nanoparticles</i> - Gramotnev, D K; Gramotnev, G)<br/></b><br/>doi: 10.1016/j.jaerosci.2004.10.003<br/><br/><b>From Duplicate 28 (<i>Design and evaluation of a nanometer aerosol differential mobility analyzer (Nano-DMA)</i> - Chen, D R; Pui, D Y H; Hummes, D; Fissan, H; Quant, F R; Sem, G J)<br/></b><br/>doi: DOI: 10.1016/S0021-8502(97)10018-0},\n private_publication = {false},\n abstract = {Recent developments in atmospheric aerosol measurements are reviewed. The topics included complement those covered in the recent review by Chow (JAWMA 45: 320-382, 1995) which focuses on regulatory compliance measurements and filter measurements of particulate composition. This review focuses on measurements of aerosol integral properties (total number concentration, CCN concentration, optical coefficients, etc.), aerosol physical chemical properties (density, refractive index, equilibrium water content, etc.), measurements of aerosol size distributions, and measurements of size-resolved aerosol composition. Such measurements play an essential role in studies of secondary aerosol formation by atmospheric chemical transformations and enable one to quantify the contributions of various species to effects including light scattering/absorption, health effects, dry deposition, etc. Aerosol measurement evolved from an art to a science in the 1970s following the development of instrumentation to generate monodisperse calibration aerosols of known size, composition, and concentration. While such calibration tools permit precise assessments of instrument responses to known laboratory-generated aerosols, unquantifiable uncertainties remain even when carefully calibrated instruments are used for atmospheric measurements. This is because instrument responses typically depend on aerosol properties including composition, shape, density, etc., which, for atmospheric aerosols, may vary from particle-to-particle and are often unknown. More effort needs to be made to quantify measurement accuracies that can be achieved for realistic atmospheric sampling scenarios. The measurement of organic species in atmospheric particles requires substantial development. Atmospheric aerosols typically include hundreds of organic compounds, and only a small fraction (~10%) of these can be identified by state-of-the-art analytical methodologies. Even the measurement of the total particulate organic carbon mass concentration is beset by difficulties including the unknown extent of evaporative losses during sampling, adsorption of gas-phase organic compounds onto sampling substrates, and the unknown relationship between carbon mass and mass of the particulate organics. The development of improved methodologies for such measurements should be a high priority for the future. Mass spectrometers that measure the composition of individual particles have recently been developed. It is not clear that these instruments will provide quantitative information on species mass concentrations, and more work is needed to routinely interpret the vast quantities of data generated during field sampling. Nevertheless, these instruments substantially expand the range of atmospheric aerosol issues that can be explored experimentally. These instruments represent the most significant advance in aerosol instrumentation in recent years.},\n bibtype = {article},\n author = {Allen, M D and Raabe, O G and Chen, Da-ren R and Pui, David Y H and Hummes, D and Fissan, H and Quant, Frederick R and Sem, G J and Gramotnev, D K and Gramotnev, G and Hering, Susanne V. and Stolzenburg, Mark R. and Quant, Frederick R and Oberreit, Derek R and Keady, Patricia B and Hoppel, W A and Iida, Kenjiro and Stolzenburg, Mark R. and McMurry, Peter H. and Smith, James N and Quant, Frederick R and Oberreit, Derek R and Keady, Patricia B and Eiguren-Fernandez, Arantza and Lewis, Gregory S and Kreisberg, Nathan M and Hering, Susanne V. and Knutson, E O and Whitby, K T and Liu, Benjamin Y H and Pui, David Y H and McMurry, Peter H. and Stolzenburg, Mark R. and McMurry, Peter H. and Swihart, Mark T and Wang, Shih Chen and Flagan, Richard C. and Considerations, General and Ii, P H Mcmurry and Hering, Susanne V. and Stolzenburg, Mark R. and Jiang, Jingkun and Attoui, Michel and Heim, Michael and Brunelli, Nicholas a. and McMurry, Peter H. and Kasper, Gerhard and Flagan, Richard C. and Giapis, Konstantinos and Mouret, Guillaume and Scientific, Elsevier and Company, Publishing and Kim, S H and Woo, K S and Liu, Benjamin Y H and Zachariah, M R and Kim, Jung Hyeun and Mulholland, George W and Kukuck, Scott R and Kinney, Patrick D and Pui, David Y H and Mulholland, George W and Bryner, Nelson P and Taylor, Publisher and Kousaka, Y and Okuyama, K and Li, Weiling and Li, Lin and Chen, Da-ren R and Li, Weiling and Li, Lin and Chen, Da-ren R and Rader, D J and McMurry, Peter H. and Manual, Service and Vemury, Srinivas and Pratsinis, Sotiris E and Feldpausch, M and Helsper, C and Wiedensohler, A},\n doi = {10.1080/027868290953416},\n journal = {Journal of Aerosol Science},\n number = {2}\n}
\n
\n\n\n
\n Recent developments in atmospheric aerosol measurements are reviewed. The topics included complement those covered in the recent review by Chow (JAWMA 45: 320-382, 1995) which focuses on regulatory compliance measurements and filter measurements of particulate composition. This review focuses on measurements of aerosol integral properties (total number concentration, CCN concentration, optical coefficients, etc.), aerosol physical chemical properties (density, refractive index, equilibrium water content, etc.), measurements of aerosol size distributions, and measurements of size-resolved aerosol composition. Such measurements play an essential role in studies of secondary aerosol formation by atmospheric chemical transformations and enable one to quantify the contributions of various species to effects including light scattering/absorption, health effects, dry deposition, etc. Aerosol measurement evolved from an art to a science in the 1970s following the development of instrumentation to generate monodisperse calibration aerosols of known size, composition, and concentration. While such calibration tools permit precise assessments of instrument responses to known laboratory-generated aerosols, unquantifiable uncertainties remain even when carefully calibrated instruments are used for atmospheric measurements. This is because instrument responses typically depend on aerosol properties including composition, shape, density, etc., which, for atmospheric aerosols, may vary from particle-to-particle and are often unknown. More effort needs to be made to quantify measurement accuracies that can be achieved for realistic atmospheric sampling scenarios. The measurement of organic species in atmospheric particles requires substantial development. Atmospheric aerosols typically include hundreds of organic compounds, and only a small fraction (~10%) of these can be identified by state-of-the-art analytical methodologies. Even the measurement of the total particulate organic carbon mass concentration is beset by difficulties including the unknown extent of evaporative losses during sampling, adsorption of gas-phase organic compounds onto sampling substrates, and the unknown relationship between carbon mass and mass of the particulate organics. The development of improved methodologies for such measurements should be a high priority for the future. Mass spectrometers that measure the composition of individual particles have recently been developed. It is not clear that these instruments will provide quantitative information on species mass concentrations, and more work is needed to routinely interpret the vast quantities of data generated during field sampling. Nevertheless, these instruments substantially expand the range of atmospheric aerosol issues that can be explored experimentally. These instruments represent the most significant advance in aerosol instrumentation in recent years.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Influence of a catalytic stripper on the response of real time aerosol instruments to diesel exhaust aerosol.\n \n \n \n \n\n\n \n Kittelson, D., B.; Watts, W., F.; Savstrom, J., C.; and Johnson, J., P.\n\n\n \n\n\n\n Journal of Aerosol Science, 36(9): 1089-1107. 9 2005.\n \n\n\n\n
\n\n\n\n \n \n \"InfluenceWebsite\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 \n \n \n \n \n \n \n \n\n\n\n
\n
@article{\n title = {Influence of a catalytic stripper on the response of real time aerosol instruments to diesel exhaust aerosol},\n type = {article},\n year = {2005},\n keywords = {Aerosol,Catalytic stripper,Diesel,Nanoparticles,Nuclei,Surface area},\n pages = {1089-1107},\n volume = {36},\n websites = {http://www.sciencedirect.com/science/article/pii/S0021850204004148},\n month = {9},\n publisher = {Pergamon},\n day = {1},\n id = {ebfae4a0-b7bf-3b40-ab10-6ea6de96cc7f},\n created = {2017-11-20T14:54:13.139Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.504Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {The objective of this study was to evaluate the response of a suite of portable, real-time aerosol instruments to Diesel exhaust aerosol with and without a catalytic stripper (CS) to determine the change in response as a function of particle size and volatility. The response of the photoemission aerosol sensor (PAS) was strongly influenced by the physical and chemical nature of Diesel aerosol. The presence of a large, predominantly volatile nuclei mode, and/or the presence of volatile material on the surface of the solid carbonaceous agglomerates in the accumulation mode suppressed the PAS response. Removal of the volatile material by passage of the aerosol through the CS enhanced the response, and improved correlations between the PAS, the diffusion charger (DC) and the scanning mobility particle sizer (SMPS). Data on aerosol size distributions, number, volume, and surface area concentrations with and without the CS in the sample stream are reported. © 2005 Elsevier Ltd. All rights reserved.},\n bibtype = {article},\n author = {Kittelson, D. B. and Watts, W. F. and Savstrom, J. C. and Johnson, J. P.},\n doi = {10.1016/j.jaerosci.2004.11.021},\n journal = {Journal of Aerosol Science},\n number = {9}\n}
\n
\n\n\n
\n The objective of this study was to evaluate the response of a suite of portable, real-time aerosol instruments to Diesel exhaust aerosol with and without a catalytic stripper (CS) to determine the change in response as a function of particle size and volatility. The response of the photoemission aerosol sensor (PAS) was strongly influenced by the physical and chemical nature of Diesel aerosol. The presence of a large, predominantly volatile nuclei mode, and/or the presence of volatile material on the surface of the solid carbonaceous agglomerates in the accumulation mode suppressed the PAS response. Removal of the volatile material by passage of the aerosol through the CS enhanced the response, and improved correlations between the PAS, the diffusion charger (DC) and the scanning mobility particle sizer (SMPS). Data on aerosol size distributions, number, volume, and surface area concentrations with and without the CS in the sample stream are reported. © 2005 Elsevier Ltd. All rights reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The influence of a cerium additive on ultrafine diesel particle emissions and kinetics of oxidation.\n \n \n \n \n\n\n \n Jung, H.; Kittelson, D., B.; and Zachariah, M., R.\n\n\n \n\n\n\n Combustion and Flame, 142(3): 276-288. 8 2005.\n \n\n\n\n
\n\n\n\n \n \n \"TheWebsite\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 \n \n \n \n \n \n \n \n\n\n\n
\n
@article{\n title = {The influence of a cerium additive on ultrafine diesel particle emissions and kinetics of oxidation},\n type = {article},\n year = {2005},\n keywords = {Additive,Cerium,Diesel,Emissions,Kinetics,Oxidation},\n pages = {276-288},\n volume = {142},\n websites = {http://www.sciencedirect.com/science/article/pii/S0010218005000994},\n month = {8},\n publisher = {Elsevier},\n day = {1},\n id = {4a0d0644-57e1-39de-941b-d2468129f85d},\n created = {2017-11-20T14:54:13.410Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.643Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {The influence of a cerium additive on the kinetics of oxidation and size distribution of ultrafine diesel particles was studied using a high-temperature oxidation-tandem differential mobility analysis method over the temperature range 300-700°C. The addition of cerium to the diesel fuel was observed to cause significant changes in number-weighted size distributions, light-off temperature, and kinetics of oxidation. The peak number concentration in the accumulation mode decreased 50 and 65%, respectively, for 25 and 100 ppm dosing levels under 1400 rpm and 75% engine load. The light-off temperature was reduced by 250 and 300°C, respectively, for 25 and 100 ppm dosing levels. The oxidation rate increased significantly (×20) with the addition of cerium to the fuel; however, the rate was relatively insensitive to dosing level. The activation energy for cerium-dosed oxidation was, within experimental error, equivalent to that for undosed fuel (Ea=100-110 kJmol-1). From a phenomenological kinetic rate perspective, the increase in oxidation rate was attributed solely to an increase in the preexponential factor. These results suggested that diesel particles using regular, undosed diesel fuels were already metal-catalyzed to some extent, most likely from metals in the lube oil. The addition of cerium likely increased the number of catalytic sites but had no effect on the overall activation energy due to the presence of other metals in the diesel particulate matter coming from lube oil. The characteristics of cerium-laden diesel particles were also investigated. Two principal types of aggregates were found using transmission electron microscopy and energy-dispersive spectrometry analysis. The first was composed mainly of agglomerates of carbonaceous spherules and a few, considerably smaller cerium oxide nanoparticles. The second consisted of metallic aggregates composed mainly of cerium oxide nanoparticles and some carbon. © 2005 Published by Elsevier Inc. on behalf of The Combustion Institute.},\n bibtype = {article},\n author = {Jung, Heejung and Kittelson, David B. and Zachariah, Michael R.},\n doi = {10.1016/j.combustflame.2004.11.015},\n journal = {Combustion and Flame},\n number = {3}\n}
\n
\n\n\n
\n The influence of a cerium additive on the kinetics of oxidation and size distribution of ultrafine diesel particles was studied using a high-temperature oxidation-tandem differential mobility analysis method over the temperature range 300-700°C. The addition of cerium to the diesel fuel was observed to cause significant changes in number-weighted size distributions, light-off temperature, and kinetics of oxidation. The peak number concentration in the accumulation mode decreased 50 and 65%, respectively, for 25 and 100 ppm dosing levels under 1400 rpm and 75% engine load. The light-off temperature was reduced by 250 and 300°C, respectively, for 25 and 100 ppm dosing levels. The oxidation rate increased significantly (×20) with the addition of cerium to the fuel; however, the rate was relatively insensitive to dosing level. The activation energy for cerium-dosed oxidation was, within experimental error, equivalent to that for undosed fuel (Ea=100-110 kJmol-1). From a phenomenological kinetic rate perspective, the increase in oxidation rate was attributed solely to an increase in the preexponential factor. These results suggested that diesel particles using regular, undosed diesel fuels were already metal-catalyzed to some extent, most likely from metals in the lube oil. The addition of cerium likely increased the number of catalytic sites but had no effect on the overall activation energy due to the presence of other metals in the diesel particulate matter coming from lube oil. The characteristics of cerium-laden diesel particles were also investigated. Two principal types of aggregates were found using transmission electron microscopy and energy-dispersive spectrometry analysis. The first was composed mainly of agglomerates of carbonaceous spherules and a few, considerably smaller cerium oxide nanoparticles. The second consisted of metallic aggregates composed mainly of cerium oxide nanoparticles and some carbon. © 2005 Published by Elsevier Inc. on behalf of The Combustion Institute.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2003\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n A closure study of aerosol mass concentration measurements: Comparison of values obtained with filters and by direct measurements of mass distributions.\n \n \n \n \n\n\n \n Park, K.; Kittelson, D., B.; and McMurry, P., H.\n\n\n \n\n\n\n Atmospheric Environment, 37(9-10): 1223-1230. 3 2003.\n \n\n\n\n
\n\n\n\n \n \n \"AWebsite\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{\n title = {A closure study of aerosol mass concentration measurements: Comparison of values obtained with filters and by direct measurements of mass distributions},\n type = {article},\n year = {2003},\n pages = {1223-1230},\n volume = {37},\n websites = {http://www.sciencedirect.com/science/article/pii/S1352231002010166},\n month = {3},\n publisher = {Pergamon},\n day = {1},\n id = {afaeb24a-d973-317f-a186-1e61e24ef79a},\n created = {2017-11-20T14:54:13.177Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.912Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {We compare measurements of aerosol mass concentrations obtained gravimetrically using Teflon coated glass fiber filters and by integrating mass distributions measured with the differential mobility analyzer-aerosol particle mass analyzer (DMA-APM) technique (Aerosol Sci. Technol. 36 (2002) 227). The DMA-APM technique measures the distribution of particle mass as a function of mobility size for particles of arbitrary shape and composition (Relationship between particle mass and mobility, and between aerodynamic and mobility size distributions for diesel exhaust particles, Environ. Sci. Technol., 2003). Because DMA-APM measurements are made on particles suspended in the air, data are not affected by volatilization or adsorption that can affect the accuracy of the filter measurements. We show that the average ratio of the filter to DMA-APM mass concentrations for laboratory-generated dioctyl sebacate (DOS) and sodium chloride (NaCl) aerosols is 1.14±0.28, and they are well correlated (R2>0.97). For diesel exhaust aerosols from an engine operating at 75% load, the two techniques agreed well with the average ratio of 0.98±0.20. When the engine was operated at a low (10%) load, mass concentrations measured with the filter were 2.13±0.54 times higher than values measured with the DMA-APM. We believe that the higher filter loading may be due to the adsorption of condensable vapors, which are emitted at higher rates under low engine load conditions. Measurements in which the condensable organics were removed with a catalytic stripper show much better agreement between the filter and DMA-APM, which support the hypothesis that vapor adsorption leads to artificially high filter data for low-load measurements. We conclude that the DMA-APM technique can be used to evaluate the accuracy of filter samples that may be affected by sampling artifacts, and to measure mass distributions with high time resolution for sub-0.5μm aerosols. © 2003 Elsevier Science Ltd. All rights reserved.},\n bibtype = {article},\n author = {Park, Kihong and Kittelson, David B. and McMurry, Peter H.},\n doi = {10.1016/S1352-2310(02)01016-6},\n journal = {Atmospheric Environment},\n number = {9-10}\n}
\n
\n\n\n
\n We compare measurements of aerosol mass concentrations obtained gravimetrically using Teflon coated glass fiber filters and by integrating mass distributions measured with the differential mobility analyzer-aerosol particle mass analyzer (DMA-APM) technique (Aerosol Sci. Technol. 36 (2002) 227). The DMA-APM technique measures the distribution of particle mass as a function of mobility size for particles of arbitrary shape and composition (Relationship between particle mass and mobility, and between aerodynamic and mobility size distributions for diesel exhaust particles, Environ. Sci. Technol., 2003). Because DMA-APM measurements are made on particles suspended in the air, data are not affected by volatilization or adsorption that can affect the accuracy of the filter measurements. We show that the average ratio of the filter to DMA-APM mass concentrations for laboratory-generated dioctyl sebacate (DOS) and sodium chloride (NaCl) aerosols is 1.14±0.28, and they are well correlated (R2>0.97). For diesel exhaust aerosols from an engine operating at 75% load, the two techniques agreed well with the average ratio of 0.98±0.20. When the engine was operated at a low (10%) load, mass concentrations measured with the filter were 2.13±0.54 times higher than values measured with the DMA-APM. We believe that the higher filter loading may be due to the adsorption of condensable vapors, which are emitted at higher rates under low engine load conditions. Measurements in which the condensable organics were removed with a catalytic stripper show much better agreement between the filter and DMA-APM, which support the hypothesis that vapor adsorption leads to artificially high filter data for low-load measurements. We conclude that the DMA-APM technique can be used to evaluate the accuracy of filter samples that may be affected by sampling artifacts, and to measure mass distributions with high time resolution for sub-0.5μm aerosols. © 2003 Elsevier Science Ltd. All rights reserved.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n The Influence of Engine Lubricating Oil on Diesel Nanoparticle Emissions and Kinetics of Oxidation.\n \n \n \n \n\n\n \n Jung, H.; Kittelson, D., B.; and Zachariah, M., R.\n\n\n \n\n\n\n In 10 2003. \n \n\n\n\n
\n\n\n\n \n \n \"TheWebsite\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
@inproceedings{\n title = {The Influence of Engine Lubricating Oil on Diesel Nanoparticle Emissions and Kinetics of Oxidation},\n type = {inproceedings},\n year = {2003},\n websites = {http://papers.sae.org/2003-01-3179/},\n month = {10},\n day = {27},\n id = {b9fb9726-7bad-3f78-82cf-9d86dc234137},\n created = {2017-11-20T14:54:13.529Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2017-11-20T14:54:13.529Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {false},\n hidden = {false},\n private_publication = {false},\n bibtype = {inproceedings},\n author = {Jung, Heejung and Kittelson, David B. and Zachariah, Michael R.},\n doi = {10.4271/2003-01-3179}\n}
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1998\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Characterization of exhaust particulate emissions from a spark ignition engine.\n \n \n \n \n\n\n \n Graskow, B., R.; Kittelson, D., B.; Abdul-Khalek, I., S.; Ahmadi, M., R.; and Morris, J., E.\n\n\n \n\n\n\n In SAE Technical Papers, 2 1998. \n \n\n\n\n
\n\n\n\n \n \n \"CharacterizationWebsite\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
@inproceedings{\n title = {Characterization of exhaust particulate emissions from a spark ignition engine},\n type = {inproceedings},\n year = {1998},\n websites = {http://papers.sae.org/980528/},\n month = {2},\n day = {23},\n id = {365d1ca1-2205-3e2b-910b-3acd38f4d732},\n created = {2017-11-20T14:54:13.635Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.814Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {Exhaust particulate emissions from a 4-cylinder, 2.25 liter spark ignition engine were measured and characterized. A single-stage ejector-diluter system was used to dilute and cool the exhaust sample for measurement. The particulate measurement equipment included a condensation nucleus counter and a scanning mobility particle sizer. Exhaust measurements were made both upstream and downstream of the catalytic converter using three different fuels. Unlike particulate emissions in diesel engines, spark ignition exhaust particle emissions were found to be highly unstable. Typically, a stable "baseline" concentration on the order of 105 particles/cm3 is emitted. Occasionally, however, a "spike" in the exhaust particle concentration is observed. The exhaust particle concentrations observed during these spikes can increase by as much as two orders of magnitude over the baseline concentration. The spikes were found to be composed of nearly 100% volatile particles which were below 30 nm in diameter. Particle number emissions at a 120 km/hr equivalent operating condition were found to be on the order of 1011-1012 particles per kilometer, more than two orders of magnitude below emissions previously reported for SI engines. The brake specific particle number emissions ranged from 1.6 × 1011 particles per kilowatt-hour at light load conditions to 7.3 × 1013 particles per kilowatt-hour at high load conditions. Brake specific particle emissions at 2000 and 2500 rpm were found to increase exponentially as a function of engine load, while the number-weighted geometric mean diameter was found to increase linearly as a function of engine load. The type of fuel used did not affect general trends, though more than a factor of two variation in the magnitude of the exhaust emissions was observed between the fuels. Copyright © 1998 Society of Automotive Engineers, Inc.},\n bibtype = {inproceedings},\n author = {Graskow, B. R. and Kittelson, D. B. and Abdul-Khalek, I. S. and Ahmadi, M. R. and Morris, J. E.},\n doi = {10.4271/980528},\n booktitle = {SAE Technical Papers}\n}
\n
\n\n\n
\n Exhaust particulate emissions from a 4-cylinder, 2.25 liter spark ignition engine were measured and characterized. A single-stage ejector-diluter system was used to dilute and cool the exhaust sample for measurement. The particulate measurement equipment included a condensation nucleus counter and a scanning mobility particle sizer. Exhaust measurements were made both upstream and downstream of the catalytic converter using three different fuels. Unlike particulate emissions in diesel engines, spark ignition exhaust particle emissions were found to be highly unstable. Typically, a stable \"baseline\" concentration on the order of 105 particles/cm3 is emitted. Occasionally, however, a \"spike\" in the exhaust particle concentration is observed. The exhaust particle concentrations observed during these spikes can increase by as much as two orders of magnitude over the baseline concentration. The spikes were found to be composed of nearly 100% volatile particles which were below 30 nm in diameter. Particle number emissions at a 120 km/hr equivalent operating condition were found to be on the order of 1011-1012 particles per kilometer, more than two orders of magnitude below emissions previously reported for SI engines. The brake specific particle number emissions ranged from 1.6 × 1011 particles per kilowatt-hour at light load conditions to 7.3 × 1013 particles per kilowatt-hour at high load conditions. Brake specific particle emissions at 2000 and 2500 rpm were found to increase exponentially as a function of engine load, while the number-weighted geometric mean diameter was found to increase linearly as a function of engine load. The type of fuel used did not affect general trends, though more than a factor of two variation in the magnitude of the exhaust emissions was observed between the fuels. Copyright © 1998 Society of Automotive Engineers, Inc.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Diesel trap performance: Particle size measurements and trends.\n \n \n \n \n\n\n \n Abdul-Khalek, I., S.; Kittelson, D., B.; and Brear, F.\n\n\n \n\n\n\n In SAE Technical Papers, 10 1998. \n \n\n\n\n
\n\n\n\n \n \n \"DieselWebsite\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
@inproceedings{\n title = {Diesel trap performance: Particle size measurements and trends},\n type = {inproceedings},\n year = {1998},\n websites = {http://papers.sae.org/982599/},\n month = {10},\n day = {19},\n id = {5bc42db5-8331-3d3b-b6a9-34df39a7bca5},\n created = {2017-11-20T14:54:13.923Z},\n accessed = {2017-11-20},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.737Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n private_publication = {false},\n abstract = {Particle concentrations and size distributions were measured in the exhaust of a turbocharged, aftercooled, direct-injection, Diesel engine equipped with a ceramic filter (trap). Measurements were performed both upstream and downstream of the filter using a two-stage, variable residence time, micro-dilution system, a condensation particle counter and a scanning mobility particle sizer set up to count and size particles in the 7-320 nm diameter range. Engine operating conditions of the ISO 11 Mode test were used. The engine out (upstream of filter) size distribution has a bimodal, log normal structure, consisting of a nuclei mode with a geometric number mean diameter, DGN, in the 10-30 nm range and an accumulation mode with DGN in the 50-80 nm range. The modal structure of the size distribution is less distinct downstream of the filter. Nearly all the particle number emissions come from the nuclei mode, are nanoparticles (Dp < 50nm), and are volatile. The emission of nanoparticles downstream of the filter is influenced strongly by the residence time; changing the residence time from 40 ms to 6000 ms increases the number concentrations of nanoparticles by up to three orders of magnitude. Residence time also considerably influences the engine out number concentrations and the shape of the size distributions, especially in the nuclei mode range. In ISO mode 8, total number concentrations vary from 47 to 600 to 311x106 part./cm3 as residence time increased from 40 to 1000 to 6000 ms. The number weighted filter penetration is highly dependent on residence time. At a residence time of 40 ms, penetrations are less than 1% for nearly all operating modes, but at a residence time of 6000 ms, modes 1,2,3, and 6 gave fairly high overall penetrations of 60 %, 73%, 130 %, and 47 %, respectively. These high penetrations result from particle nucleation downstream of the filter, not inherent filtration performance. The filter self-regenerates when the engine is operated at the rated power condition (ISO mode 1). During regeneration, emissions of particles below 13 nm in diameter increase by more than three orders of magnitude for a few minutes. Particle nucleation and growth during dilution are poorly understood and strongly influence measured number emissions and size distributions. More work is needed to establish consistent, well-defined sampling and dilution procedures for measurements of particle number emissions and size, especially in the nanoparticle size range. These procedures should, to the extent possible, mimic atmospheric dilution conditions so that measurements are representative of real world exposures rather than arbitrarily defined laboratory conditions. © 1998 Society of Automotive Engineers, Inc.},\n bibtype = {inproceedings},\n author = {Abdul-Khalek, Imad S. and Kittelson, David B. and Brear, Fred},\n doi = {10.4271/982599},\n booktitle = {SAE Technical Papers}\n}
\n
\n\n\n
\n Particle concentrations and size distributions were measured in the exhaust of a turbocharged, aftercooled, direct-injection, Diesel engine equipped with a ceramic filter (trap). Measurements were performed both upstream and downstream of the filter using a two-stage, variable residence time, micro-dilution system, a condensation particle counter and a scanning mobility particle sizer set up to count and size particles in the 7-320 nm diameter range. Engine operating conditions of the ISO 11 Mode test were used. The engine out (upstream of filter) size distribution has a bimodal, log normal structure, consisting of a nuclei mode with a geometric number mean diameter, DGN, in the 10-30 nm range and an accumulation mode with DGN in the 50-80 nm range. The modal structure of the size distribution is less distinct downstream of the filter. Nearly all the particle number emissions come from the nuclei mode, are nanoparticles (Dp < 50nm), and are volatile. The emission of nanoparticles downstream of the filter is influenced strongly by the residence time; changing the residence time from 40 ms to 6000 ms increases the number concentrations of nanoparticles by up to three orders of magnitude. Residence time also considerably influences the engine out number concentrations and the shape of the size distributions, especially in the nuclei mode range. In ISO mode 8, total number concentrations vary from 47 to 600 to 311x106 part./cm3 as residence time increased from 40 to 1000 to 6000 ms. The number weighted filter penetration is highly dependent on residence time. At a residence time of 40 ms, penetrations are less than 1% for nearly all operating modes, but at a residence time of 6000 ms, modes 1,2,3, and 6 gave fairly high overall penetrations of 60 %, 73%, 130 %, and 47 %, respectively. These high penetrations result from particle nucleation downstream of the filter, not inherent filtration performance. The filter self-regenerates when the engine is operated at the rated power condition (ISO mode 1). During regeneration, emissions of particles below 13 nm in diameter increase by more than three orders of magnitude for a few minutes. Particle nucleation and growth during dilution are poorly understood and strongly influence measured number emissions and size distributions. More work is needed to establish consistent, well-defined sampling and dilution procedures for measurements of particle number emissions and size, especially in the nanoparticle size range. These procedures should, to the extent possible, mimic atmospheric dilution conditions so that measurements are representative of real world exposures rather than arbitrarily defined laboratory conditions. © 1998 Society of Automotive Engineers, Inc.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 1995\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Real time measurement of volatile and solid exhaust particles using a catalytic stripper.\n \n \n \n\n\n \n Abdul-Khalek, I., S.; and Kittelson, D., B.\n\n\n \n\n\n\n SAE Technical Papers, 950236. 1995.\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{\n title = {Real time measurement of volatile and solid exhaust particles using a catalytic stripper},\n type = {article},\n year = {1995},\n volume = {950236},\n id = {3901c6f3-a030-38bb-b982-6d0dfb9af77f},\n created = {2017-11-17T15:14:59.085Z},\n file_attached = {false},\n profile_id = {317fdcd2-b041-3222-bca0-702f39879f87},\n group_id = {cd514115-949a-34c4-b220-b5e9bedcc2c8},\n last_modified = {2020-07-28T15:23:37.764Z},\n read = {false},\n starred = {false},\n authored = {false},\n confirmed = {true},\n hidden = {false},\n source_type = {JOUR},\n private_publication = {false},\n abstract = {A system has been developed that allows near real time measurements of total, volatile, and nonvolatile particle concentrations in engine exhaust. It consists of a short section of heated catalyst, a cooling coil, and an electrical aerosol analyzer. The performance of this catalytic stripper system has been characterized with nonvolatile (NaCl), volatile sulfate ((NH 4)2 SO4), and volatile hydrocarbon (engine oil) particles with diameters ranging from 0.05-0.5 μm. The operating temperature of 300°C gives essentially complete removal of volatile sulfate and hydrocarbon particles, but also leads to removal of 15-25% of solid particles. This system has been used to determine total, volatile, and nonvolatile particle concentrations in the exhaust of a Diesel engine and a spark ignition engine. Volatile volume fractions measured in Diesel exhaust with the catalytic stripper system increased from 19-65% as the equivalence ratio (load) decreased from 0.64-0.13. Volatile volume fractions agreed with volatile mass fractions measured by a filter vacuum sublimation method, except at low equivalence ratios (0.13-0.4) where the filter method gave somewhat higher values. Particles were measured in the exhaust of the spark ignition engine upstream of its catalytic converter. Concentrations are very low under normal operating conditions, in the same range as ambient air, but show a very strong increase with increasing equivalence ratio. This suggests that under rich conditions, e.g., cold starts, high altitudes, and high loads, particle emissions may be significant. Volume volatile fractions for the spark ignition exhaust particles are highest, about 60% at equivalence ratios near 1 and fall to less than 30% at an equivalence ratio of 1.35. © Copyright 1995 Society of Automotive Engineers, Inc.},\n bibtype = {article},\n author = {Abdul-Khalek, Imad S. and Kittelson, David B.},\n doi = {10.4271/950236},\n journal = {SAE Technical Papers}\n}
\n
\n\n\n
\n A system has been developed that allows near real time measurements of total, volatile, and nonvolatile particle concentrations in engine exhaust. It consists of a short section of heated catalyst, a cooling coil, and an electrical aerosol analyzer. The performance of this catalytic stripper system has been characterized with nonvolatile (NaCl), volatile sulfate ((NH 4)2 SO4), and volatile hydrocarbon (engine oil) particles with diameters ranging from 0.05-0.5 μm. The operating temperature of 300°C gives essentially complete removal of volatile sulfate and hydrocarbon particles, but also leads to removal of 15-25% of solid particles. This system has been used to determine total, volatile, and nonvolatile particle concentrations in the exhaust of a Diesel engine and a spark ignition engine. Volatile volume fractions measured in Diesel exhaust with the catalytic stripper system increased from 19-65% as the equivalence ratio (load) decreased from 0.64-0.13. Volatile volume fractions agreed with volatile mass fractions measured by a filter vacuum sublimation method, except at low equivalence ratios (0.13-0.4) where the filter method gave somewhat higher values. Particles were measured in the exhaust of the spark ignition engine upstream of its catalytic converter. Concentrations are very low under normal operating conditions, in the same range as ambient air, but show a very strong increase with increasing equivalence ratio. This suggests that under rich conditions, e.g., cold starts, high altitudes, and high loads, particle emissions may be significant. Volume volatile fractions for the spark ignition exhaust particles are highest, about 60% at equivalence ratios near 1 and fall to less than 30% at an equivalence ratio of 1.35. © Copyright 1995 Society of Automotive Engineers, Inc.\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);