• R/O
  • SSH

Commit

Tags
Keine Tags

Frequently used words (click to add to your profile)

javac++androidlinuxc#windowsobjective-ccocoa誰得qtpythonphprubygameguibathyscaphec計画中(planning stage)翻訳omegatframeworktwitterdomtestvb.netdirectxゲームエンジンbtronarduinopreviewer

Commit MetaInfo

Revisione97c3b402ad223c16266d30c315a2ab67e7d2578 (tree)
Zeit2007-10-01 17:17:24
Autoriselllo
Commiteriselllo

Log Message

Yannis added a new intro to the paper + some minor modifications.

Ändern Zusammenfassung

Diff

diff -r ef89800fd9ba -r e97c3b402ad2 latex-documents/jas1/mydraft.tex
--- a/latex-documents/jas1/mydraft.tex Mon Oct 01 08:14:36 2007 +0000
+++ b/latex-documents/jas1/mydraft.tex Mon Oct 01 08:17:24 2007 +0000
@@ -1,11 +1,10 @@
11 % Template article for preprint document class `elsart'
22 % SP 2006/04/26
33
4-% %\documentclass{elsart}
4+ \documentclass{elsart}
55
66 % Use the option doublespacing or reviewcopy to obtain double line spacing
7-
8- \documentclass[]{elsart}
7+% \documentclass[doublespacing]{elsart}
98
109 % if you use PostScript figures in your article
1110 % use the graphics package for simple commands
@@ -16,7 +15,7 @@
1615 % \usepackage{epsfig}
1716
1817 \usepackage{graphicx}
19-
18+\usepackage{url}
2019 \usepackage{natbib}
2120 %\newcommand{\ol}{\overline}
2221 \renewcommand{\i}{\int}
@@ -122,7 +121,7 @@
122121 % \linenumbers
123122 \begin{document}
124123
125-%
124+%
126125 % \begin{linenumbers}
127126 \begin{frontmatter}
128127
@@ -142,247 +141,214 @@
142141 % \address{Address\thanksref{label3}}
143142 % \thanks[label3]{}
144143
145-\title{One-dimensional Dynamics of Diesel Exhaust Particles}
144+\title{One-dimensional dynamics of diesel exhaust particles}
146145
147146 % use optional labels to link authors explicitly to addresses:
148147
149148
150- \author{{L. Isella\corauthref{ise}}$^{\rm a}$}
149+ \author{{L. Isella\corauthref{ise}}$^{\rm a}$},
151150 \ead{lorenzo.isella@jrc.it}
152-\author{, B. Giechaskiel$^{\rm a}$}
153-\author{, Y. Drossinos$^{\rm a,b}$}
151+\author{B. Giechaskiel$^{\rm a}$},
152+\author{Y. Drossinos$^{\rm a,b}$}
154153 \address{$^{\rm a}$European Commission, Joint Research Centre, I-21020 Ispra (VA), Italy}
155154 \corauth[ise]{Corresponding author. Tel.: +39 0332786499}
156- \address{$^{\rm b}$School of Mechanical and Systems Engineering, Newcastle University, Newcastle upon Tyne NE1 7RU, United Kingdom}
155+ \address{$^{\rm b}$School of Mechanical and Systems Engineering, Newcastle University,
156+ Newcastle upon Tyne NE1 7RU, United Kingdom}
157157
158-
159-
158+%\date{today, \textrm{OneDSep20\_yd1.tex}}
160159
161160 % \author{}
162161
163162 % \address{}
164163
165164 \begin{abstract}
166-% [few lines by Makis sketching the experiment].
167-We study the one-dimensional dynamics of a diesel Euro-3 car exhaust solid particles along the transfer line (a cylindrical duct) leading the exhaust fumes from the tailpipe to the dilution tunnel.
168-For moderate Reynolds numbers and under steady-state conditions, we map the system dynamics into the dynamics of aerosol particles in an ageing chamber.
169-We account for agglomeration, diffusional and thermophoretic losses but we do not include particle fragmentation in this study. Inertial deposition is neglected because the observed sub-micron aggregates have a low Stokes number.
170-The time scales of the aerosol processes are estimated and thermophoretic deposition is found to be non-negligible at high engine loads, despite the dynamics being generally dominated by agglomeration.
171- The fractal dimension of the aggregates, determined self-consistently in this study, exerts a significant influence on the evolution of the number-size distribution. When the distribution remains ``frozen" inside the dilution tunnel, e.g. at high dilution ratios, the exhaust-particle residence time in the transfer line affects the measurement at the standard sampling point (dilution tunnel outlet).
172-We apply the methodology in this study to the exhaust solid particle measurements conducted at VELA-2 laboratories at the JRC Italy.
173-
174-
175-
176-
177-
165+We study, experimentally and theoretically, the dynamics of solid
166+particles emitted from a diesel Euro-3 light-duty vehicle along the
167+transfer line that conducts the exhaust fumes from the tailpipe to
168+the dilution tunnel. Particle agglomeration, and diffusional and
169+thermophoretic transport are modelled, whereas aggregate
170+fragmentation and inertial effects are neglected. For turbulent, but
171+moderate, Reynolds numbers and under steady-state conditions we map
172+the average, one-dimensional dynamics into the dynamics of aerosol
173+particles in an ageing chamber. The aggregate fractal dimension,
174+determined self-consistently to vary from 2 to 2.2 depending on
175+engine load, influences significantly the evolution of the
176+number distribution along the transfer line. The relative
177+importance of aerosol processes is estimated by defining appropriate
178+characteristic time scales. For the experiments analyzed
179+agglomeration is the dominant process, suggesting that the particle
180+residence time in the transfer line might affect solid-particle
181+measurements at the dilution tunnel exit. Thermophoretic losses are
182+calculated non-negligible only at high engine loads.
178183 \end{abstract}
179184
180185 \begin{keyword}
181186 % keywords here, in the form: keyword \sep keyword
182-agglomeration, deposition, fractal, Smoluchowski, diesel particles
187+Agglomeration, Fractal dimension, Smoluchowski equation, Diesel particles
183188 % PACS codes here, in the form: \PACS code \sep code
184-\PACS
189+\PACS
185190 \end{keyword}
186191 \end{frontmatter}
187192
188193 % main text
189194
190195 \section{Introduction}
191-Exhaust fine particles raise growing environmental concerns ranging from adverse health effect to climate change.
192-The current European regulations for light-duty diesel engine particle emissions focus on particulate mass, but research calls for new metrics, such as e.g. the number-size distribution or the active surface of the emitted particles \citep{burtscher review}.
193-Car emissions (both in mass and number) are often measured using a constant volume sampler (CVS), which takes the exhaust fumes from the tailpipe to a dilution tunnel, where turbulent mixing with clean dilution air takes place.
194-The exhaust dynamics along the experimental manifold is in general sensitive to the sampling conditions and care is needed to obtain repeatable and reproducible results \citep{burtscher review}.
195-Understanding the diesel particle dynamics along the experimental manifold has a fundamental importance both for the planning and interpretation of the experiments and for the careful design of new instrumentation.
196-% However, the study of the particle dynamics along the experimental manyfold is a very demanding task. The exhaust soot particle processes occur within a turbulent flow.
197-In this study we investigate the dynamics of solid (soot) particles. The measurement of solid particles, through a procedure of hot sampling of the aerosol removing the volatiles which may coat particles, is recommended by the particle measurement protocol since it generally leads to more repeatable and reproducible results.
198-One distinctive feature of diesel exhaust aggregates is that they collide giving rise to complex, fractal-like structures.
199-In this work, we do not include the possibility of aggregate breakage.
200-The fractal-like nature of the aggregates has profound implications on the dynamics, but
201-the experimental mesurement of the fractal dimension usually requires sophisticated techniques (see e.g. \cite{fractal experimental} and references therein) which may not be available to every laboratory.
202- In the present study we devise a method for the algorithmic determination of the fractal dimension which relies both on the statistical properties and on the evolution of the particle-size distribution.
203-We limit the investigation of the dynamics of a Euro-3 diesel exhaust particles to the trasfer line connecting the car tailpipe to the dilution tunnel inlet.
204-The rationale behind this choice is that, at least for high dilution ratios in the dilution tunnel, we found that most of the evolution of the number-size distribution occurs before the aggregates enter the tunnel. In this situation the aerosol processes taking place in the transfer line heavily affect the outcome of a measurement conducted at the standard sampling point, typically located at the outlet of the dilution tunnel.
205-Due to the simplicity of the transfer line geometry (a pipe), we formulate an approximated one-dimensional description of the aerosol dynamics, whose limits of applicability are discussed.
206-The methodology is applied to interpret a series of measurements conducted on a Euro-3 vehicle in the VELA-2 laboratories at JRC, Italy, using a standard CVS system.
207-\begin{table}
208-% \begin{center}
209-\begin{tabular}{|c|r@{}lr@{}lr@{}lr@{}l||r|}
210- \hline
211-\multicolumn{10}{|c|}
212- {\rule[-3mm]{0mm}{8mm} \textbf{Statistics of the number-size distribution}} \\
213- Speed
214- & \multicolumn{2}{|c|}{Position}
215- & \multicolumn{2}{|c|}{$N$ [cm$^{-3}$] (fit)}
216- & \multicolumn{2}{|c|}{$N$ [cm$^{-3}$] (reconstructed)}
217- & \multicolumn{2}{|c|}{$\mu$ [nm]}
218- & $\sigma$ [nm] \\ \hline \hline
219-50 Km/h
220- & \multicolumn{2}{|c|}{\rm Inlet}
221- & \multicolumn{2}{|c|}{$7.6\cdot 10^7$}
222- &\multicolumn{2}{|c|}{$7.9\cdot 10^7$}
223- &\multicolumn{2}{|c|}{$62$}
224-& $1.75$
225-\\
226-50 Km/h
227- & \multicolumn{2}{|c|}{\rm Outlet}
228- & \multicolumn{2}{|c|}{$4.5\cdot 10^{7}$}
229- &\multicolumn{2}{|c|}{}
230- &\multicolumn{2}{|c|}{$ 87.3$}
231-& $1.63$
232-\\
196+Diesel exhaust fine particles are extensively studied, both experimentally and theoretically, because they have been associated with various effects, ranging from adverse health effect to climate change. Current European regulations for light-duty diesel engine particle emissions are based on total emitted particulate mass. The exhaust gases of vehicles are diluted in a (full) dilution tunnel from where a sample is drawn and collected on a filter. Recent research results suggest, however, that different metrics, such as number distribution or active surface, may quantify particles effect more precisely, especially their possible health-related [Ober..].
197+Moreover concerns are raised whether the results from the dilution tunnel are representative of the reality. \cite{makis2004} and \cite{ronkko} showed that tailpipe measurements were representative of vehicle emissions by comparing the size distributions a few meters behind a Euro 3 vehicle (with chasing experiments) with those measured with the same sampling system in the laboratory. However, \cite{maricq tailpipe} showed that there are differences between tailpipe and dilution tunnel in mass and number. Although they emphasized mainly the differences in the nucleation mode of the size distribution, the figures in their study show that there are also differences in the accumulation mode (at figure 10 in their paper the size distribution at the tunnel is more narrow, at bigger diameters and with lower concentration). \cite{vogt} presented that the size distributions at the tailpipe and the dilution tunnel were different with similar tendency as the one described by \cite{maricq tailpipe}. They attributed the differences to agglomeration. More recently \cite{casati} found even bigger differences between the tailpipe and the dilution tunnel size distributions. The studies mentioned before measured total particles (volatile and non-volatile part) as their sampling systems used dilution air at ambient temperature, so many phenomena could explain the differences they observed (e.g. nucleation, condensation, agglomeration, deposition etc). There is a drawback with their approach: by favoring procedures like nucleation and condensation at the tailpipe measurements, the effect of the sampling system cannot be isolated from the procedures that take place actually in the transfer tube. Simple calculations show that no nucleation and condensation takes place in the transfer tube and in order to study the procedures that take place the sampling conditions must be such that only the accumulation (soot) mode is measured. In this direction the “PMP” protocol is more appropriate.
198+The Particulate Measurement Programme (PMP), an international collaborative programme, has been established to develop a measurement protocol to measure repeatably (intra-laboratory) and reproducibly (inter-laboratory) particle number emissions to replace, or complement, the existing mass-based system for regulatory purposes. According to PMP the particle number system should consist of a first stage hot dilution and then a heated section at 300-400°C for the removal of volatiles, avoiding thus the uncertainty of the volatiles which affects the repeatability (and reproducibility) of the measurements. However, in order to minimize required changes to the current type approval facilities, the system will be sampling from the full dilution tunnel. The results of the light duty phase of the PMP, which were recently published \citep{andersson1}, showed that the number method has better sensitivity than the mass method and can be used for legislative purposes. In fact a new number limit of $5\cdot10^{11}$cm$^{-3}$ is already proposed \citep{good}. Although there is general guidline for the sampling point at the dilution tunnel (10-20 diameters downstream the mixing point), there are no specifications for the transfer tube. Thus there are labs with different lengths. For example, \cite{leon1} and \cite{vogt} have $6$m insulated tubes, \cite{maricq coagulation} have $8$m ($10$cm inner diameter), the lab of the present study has $9$m heated line.
199+It is, thus, essential to analyze and model particle dynamics in the transfer tube from the tailpipe to the dilution tunnel, to ensure that measurements in different laboratories are comparable and to suggest the most appropriate experimental geometry for regulatory measurements. The transfer tube is considered the most critical part as, at least for high dilution ratios, in the dilution tunnel experimental measurements and theoretical calculation suggest that the size distribution freezes.
200+In this paper the particle number (and mass) distributions are studies both experimentally and theoretically. Procedures like agglomeration, thermophoresis, diffusion and deposition are studied and their relative importance is discusses for a Euro 3 vehicle at different steady state tests. The fractal structure of the particles is also taken into account.
233201
234-120 Km/h
235- & \multicolumn{2}{|c|}{\rm Inlet}
236- & \multicolumn{2}{|c|}{$1.4\cdot 10^8$}
237- &\multicolumn{2}{|c|}{$1.38\cdot10^8$}
238- &\multicolumn{2}{|c|}{$67.7$}
239-& $1.79$
240-\\
241-120 Km/h
242- & \multicolumn{2}{|c|}{\rm Outlet}
243- & \multicolumn{2}{|c|}{$8.3\cdot 10^{7}$}
244- &\multicolumn{2}{|c|}{}
245- &\multicolumn{2}{|c|}{$ 88.5$}
246-& $1.69$
247-\\
248202
249- \hline
250-% \vspace*{0.cm}
251-\end{tabular}
252- \caption{Characteristic time scales for the aerosol processes considered in the two sets of experiments and residence time of the aggregates in the transfer line.}
253- \label{table statistics}
254-% \end{center}
255-\end{table}
256-
257203 \section{Experiments}
204+
205+% Experimental set up
206+
207+% Particle size distributions, in particular that they
208+% freeze at the CVS inlet.
209+
258210 \subsection{Experimental set-up and particle sampling}
259-The vehicle used in this study was a common rail DI FIAT Stilo JTD (Euro 3) with common market fuel with a sulfur content below $50$ppm and with lubrificant [{it find out!}].
260-The vehicle was coupled to the dilution tunnel transfer line by a metal-to-metal join during testing to avoid the possibility of exhaust contamination by the high-temperature breakdown of elastomer coupling elements. The exhaust was transported to the dilution tunnel through a $9$m long heated at $70\,^{\circ}\mathrm{C}$ corrugated stainless steel tube. This tube, anaconda, was introduced along the dilution tunnel axis, near an orifice plate that ensured rapid mixing with the dilution air (position d). The results at the dilution tunnel will be presented elsewhere and will not be further discussed.
261-Aerosol samples for particle number measurements were drawn with a modified Fine Particle Sampler (FPS) (Dekati Ltd.) with as short as possible not heated sampling lines. The same length probes were used at all different sampling positions at the dilution tunnel.
262-The schematic of the sampling system is shown in Fig.~\ref{schematic experiment}.
263-\begin{figure}
264-\begin{center}
265-\includegraphics[width=0.8\columnwidth, height=6cm]{JRC_CVS.jpg}
266-\end{center}
267-\caption{Schematic of the sampling system used in the experiments conducted in the VELA-2 laboratories at JRC, Italy.}\label{schematic experiment}
268-\end{figure}
269-With FPS, dilution is carried out in two phases. The first dilution phase is conducted as close to a sampling point as possible, i.e., in a perforated tube diluter. This method assists reproducible results and prevents losses. The first, i.e., primary dilution can be hot or cold. With hot dilution, condensation of volatile species can be prevented, where as with cold dilution, the nucleation of volatile species can be maximized. The second dilution phase is an ejector type diluter, located downstream the primary dilution. The ejector diluter acts as a pump sucking a known amount of a diluted sample from the primary dilution. Simultaneously, a secondary dilution is carried out. In the FPS, critical parameters of the dilution are controlled and simultaneously monitored. The dilution ratio of the measurement can be calculated reliably, thus the measured concentration of particles can be converted to an actual particle concentration.
270-In this study most measurements were close to the PMP protocol specifications (hot dilution). The dilution ratios at FPS were between $30:1$ and $40:1$ (probe diluter dilution ratio $1.5:1$ to $2:1$ and ejector dilutor dilution ratio $17:1$ to $20:1$). The dilution ratio was further checked with $CO_2$ concentrations when FPS was measuring at the tailpipe. For $50$ Km/h there was no difference, but for $120$ Km/h there was (the DR was overestimated $20\%$). This factor was taken into account at the calculations. The temperature was set to $400\,^{\circ}\mathrm{C}$ for the probe heater and $150\,^{\circ}\mathrm{C}$ for the (ejector) dilution air in order to evaporate volatile particles and reduce the partial pressures of the gas phase species to prevent recondensation at the diluter exit. These temperatures led to diluted aerosol temperatures at the exit of the diluter $\sim 120\,^{\circ}\mathrm{C}$.
271-Downstream FPS an evaporation tube (ET) at $400\,^{\circ}\mathrm{C}$ was connected in order to evaporate all volatiles and semi-volatile compounds. The residence time in this tube was estimated $0.2$s (needs to be confirmed).
272-Immediately downstream of the evaporation tube there was an ejector diluter in order to minimize the diffusion losses, cool the hot diluted exhaust gas and reduce the particle number concentration below $10^5$ cm$^{-3}$ in order to be within the detection limit of the instruments (CPCs). The dilution ratio was constant $10.5:1$ for the specific overpressure (1 bar) in this set up. An extra ejector diluter was used in some measurements (e.g. tailpipe and end anaconda) where the particle number concentrations were higher. The dilution ratio of this diluter was calculated $13.5:1$. The calculations of the dilution ratios were done with particles upstream and downstream the diluters as no calibration gas or flowmeters were available at that period. The ejector dilutors were cleaned only at the beginning of the measurement campaign.
273-Some measurements were conducted without any heating at FPS and the ET in order to measure volatile particles also (as nucleation mode or condensed material on the soot particles). These measurements are noted as “cold” dilution measurements because ambient temperature air was used for the dilution. The particles measured with cold dilution are called “wet” particles in order to distinguish them from the normal measurements (PMP protocol or “hot” dilution that give as a result non-volatile or “dry” particles).
211+The vehicle used in this study was a common rail DI FIAT Stilo JTD (Euro 3) fueled with common market fuel (sulfur $<50$ ppm).
212+The schematic of the experimental setup and sampling system is given in Fig.~\ref{schematic experiment}.
213+The vehicle was coupled to the dilution tunnel transfer line by a metal-to-metal join during testing to avoid the possibility of exhaust contamination by the high-temperature breakdown of elastomer coupling elements (position 1). The exhaust was transported to the dilution tunnel through a $9$m-long (inner diameter $10$cm) heated at $70\,^{\circ}\mathrm{C}$ corrugated stainless steel tube (the end of the tube is considered position 2). This tube was introduced along the (full) dilution tunnel axis, near an orifice plate that ensured rapid mixing with the dilution air. Particle samples were drawn at the beginning of the transfer tube (position 1) and at the end (position 2) before entering the dilution tunnel (Fig.~\ref{schematic experiment}). Samples were also drawn from the normal legislated position at the dilution tunnel $10$ tunnel diameters after the mixing point.
214+Aerosol samples for particle number measurements were drawn with a modified Fine Particle Sampler (FPS) (Dekati Ltd.) with as short as possible not heated sampling lines. The same length probes were used at the two different sampling positions. With FPS, dilution is carried out in two phases. The first dilution phase is conducted as close to a sampling point as possible, i.e., in a perforated tube diluter. This method assists reproducible results and prevents losses. The second dilution phase is an ejector type diluter, located downstream the primary dilution. The ejector diluter acts as a pump sucking a known amount of a diluted sample from the primary dilution. Simultaneously, a secondary dilution is carried out. In the FPS, critical parameters of the dilution are controlled and simultaneously monitored. The dilution ratio of the measurement can be calculated reliably, thus the measured concentration of particles can be converted to an actual particle concentration.
215+In this study the measurements were according to the PMP protocol specifications (hot dilution). The dilution ratios at FPS were between $30:1$ and $40:1$ (probe diluter dilution ratio $1.5:1$ to $2:1$ and ejector dilutor dilution ratio $17:1$ to $20:1$). The temperature was set to $400\,^{\circ}\mathrm{C}$ for the probe heater and $150\,^{\circ}\mathrm{C}$ for the (ejector) dilution air in order to evaporate volatile particles and reduce the partial pressures of the gas phase species to prevent recondensation at the diluter exit. These temperatures led to diluted aerosol temperatures at the exit of the diluter $\sim 120\,^{\circ}\mathrm{C}$.
216+Downstream FPS an evaporation tube at $400\,^{\circ}\mathrm{C}$ was connected in order to evaporate all volatiles and semi-volatile compounds. The residence time in this tube was estimated $0.15$s. Immediately downstream of the evaporation tube there were two ejector diluters \citep{makis2004} in order to minimize the diffusion losses, cool the hot diluted exhaust gas and reduce the particle number concentration below $10^5$cm$^{-3}$ in order to be within the detection limit of the instruments (CPCs). The dilution ratio was constant $\sim 150$ for the specific overpressure ($1$bar) in this set up. The estimations of the dilution ratios were done with particles upstream and downstream the diluters as no calibration gases or flowmeters were available at that period. The ejector dilutors were cleaned only at the beginning of the measurement campaign.
274217
275218 \subsection{Instrumentation and PM Sampling}
276-Two CPCs (TSI Inc. 3025A, Dekati prototype), one SMPS (TSI Inc. 3936) and one DMM (Dekati Ltd.) were measuring total particle number concentration, particle number size distribution and mass concentration respectively downstream the ejector dilutor.
277-Mass samples were drawn downstream the FPS with a flowrate of $30$lpm. A long metal tube was used to cool the hot ($\sim 120°C$) diluted exhaust gas.
278-Mass samples were also drawn from the CVS at the normal sampling position at a constant flowrate of 25 lpm at normal conditions ($0\,^{\circ}\mathrm{C}$ and $1$ bar). Mass samples were collected on a single $47$mm Quarzo filters to permit chemical analysis with MEXA. The results of the chemical analysis are not available yet as the MEXA unit had to be sent back to the manufacturer for calibration.
279-For some tests normal T60A20 filters were used at CVS. After their weight was determined the Non Volatile Fraction (NVF) content was determined indirectly with the following procedure: The previously loaded and weighted filters were heated in a furnace under continuous flux of $N_2$. Afterwards filters remained in the conditioning room for another $24$h and then heated filters weights were taken. NVF emissions of the vehicle were calculated as the difference between tare and heated filter weights. It should be noted that even when a blank filter follows this procedure losses some of its mass ($\sim30\mu$g for T60A20 filters) due to its volatile content. This weight loss was also taken into account, but it didn’t affect the results as the weight of the filters at CVS was $>1500\mu$g.
219+One Condensation Particle Counter (CPC, TSI Inc. 3025A) and one Scanning Mobility Particle Sizer (SMPS, TSI Inc. 3936) were measuring total particle number concentration and particle number size distribution downstream the sampling system. The CPC is capable of measuring the number concentration of submicrometer airborne particles that are larger than $3$nm in diameter. The particles are detected and counted by a simple optical detector after a supersaturated vapor condenses onto the particles, causing them to grow into larger droplets. The range of particle concentration detection extends from less than $0.01$ particle/cm$^33$ to $9.99 \cdot 10^4$ particles/cm$^3$.
220+In SMPS particles are classified with an Electrostatic Classifier ($103$ bins) and their concentration is measured with a 3010 CPC. The SMPS system also uses a personal computer and custom software to control individual instruments and perform data reduction. Losses in the SMPS were taken into account with the manufacturer’s new software. In this study sheath/sample flowrates used were $3/0.3$. With these flowrates it was possible to measure particles in the range of $15-640$nm and it was ensured that the activity of the neutralizer of the SMPS (model 3077) was adequate for the charging of the particles. The sample flowrate was checked with a bubble flowmeter (BUCK calibrator M-5). The upscan and downscan times for the SMPS were 90/30 s respectively.
221+Mass samples were also drawn from the CVS at the normal sampling position at a constant flowrate of $25$lpm at normal conditions ($0\,^{\circ}\mathrm{C}$ and $1$bar). Mass samples were collected on a single $47$mm Quarzo filters to permit chemical analysis with a Sunset PM analyser. For some tests T60A20 filters were used at CVS. After their weight was determined, the Non Volatile Fraction (NVF) content was determined indirectly with the following procedure: The previously loaded and weighted filters were heated in a furnace under continuous flux of N2. Afterwards filters remained in the conditioning room for another 24 h and then heated filters weights were taken. NVF emissions of the vehicle were calculated as the difference between tare and heated filter weights. It should be noted that even when a blank filter follows this procedure losses some of it’s mass ($\sim30\mu$g for T60A20 filters) due to its volatile content. This weight loss was also taken into account, but it didn’t affect the results as the weight of the filters at CVS was $>1500\mu$g.
280222 \subsection{Steady-state tests: stability and repeatability}
281-Steady-state tests at $50$ and $120$Km/h were run. The steady-state tests lasted for $20$ minutes. In these $20$ minutes two CVS flowrates were tested: $6$ and $12$ m$^3$/min.
282-% The detailed procedure will be given in the results section.
283-The stability of the measurements (for DMM, CPC1=TSI, CPC2=DEKATI and total concentration of SMPS) can be checked by estimating the CoV, defined as the standard deviation divided by the mean. Usually it was better than $5\%$ with the exception of some measurements at the mixing point (between) where it was around $10\%$. These values are within the common experimental uncertainties.
284-For SMPS the size dependent stability was also checked.
285-% Figure 7 shows a size distribution and the mean variation at 50 and 120 km/h.
286-At the size range of $40-200$nm the CoV is $<10\%$ and for $60-120$nm is $<5\%$. In these size ranges the concentration is high. Outside this range the concentration of particles is very low leading to high CoVs.
287-The repeatability of the measurements was checked by comparing the results at the same sampling position of two different days (beginning and end of measurement campaign).
288-% Figure 8 shows the results at the end of anaconda for 120 km/h where the FPS was used with exactly the same settings and temperatures. Red lines are the results of the first period and green lines for the second after a -10\% correction. The size distributions are exactly the same (mean, stdev) indicating no change of the size distribution. The -10\% difference of the concentration can be attributed to the dirtiness of the sampling device (FPS and ejector) or uncertainties in the ejector dilution air pressure setting.
289- The second repeatition was conducted after the sampling system had measured at the tailpipe. When the sampling system gets dirty its dilution ratio increases and this cannot be taken into account from the instruments dilution ratio indication. The results presented in this report have been corrected for this $-10\%$ for the second period of measurements (after the measurements at the tailpipe).
290-\section{Aggregate structure}
223+Steady state tests at $50$ and $120$km/h of $20$ minutes duration were run. Only the last $10$ minutes of each test were taken into account to be sure that the vehicle had stabilized \citep{makis2007}. The measured exhaust gases flowrates were $1$ and $2$m$^3$/min respectively leading to Re number in the transfer tube (taking into account the high exhaust temperature) of $\sim 8000-13000$. The stability of the measurements was checked with the CoV (=stdev/mean) of each test. Usually it was better than 5\%. For SMPS the size dependent stability was also checked. At the size range of $60-120$nm the stability was very good (<5\%), but outside this range it was higher due to the low particle number concentration measured. The peak of the size distributions was also quite stable during the measurements (CoV $<2\%$ or $<2$nm)
224+The repeatability of the measurements was checked by comparing the results at the same sampling position (end of transfer tube) of two different days (beginning and end of measurement campaign with other measurements in between). The two tests had a difference of $10\%$ probably due to the dilution ration change of the system as it got dirty during the measurements. When the sampling system get’s dirty its dilution ratio increases and this cannot be taken into account from the instruments dilution ratio indication.
225+It should be also emphasised that due to the high temperature and pressure fluctuations at the tailpipe the value given by the instrument might differ from the actual one. At the end of the transfer tube the temperature and the pressure fluctuations are lower so the values given by the instrument should be close to the actual ones. For this reason the size distributions measured at the end of the transfer tube were considered the “accurate” ones and from the tailpipe size distributions the concentration was fitted in order to have mass balance (mean and standard deviation were kept the same) as it will be explained later. However it was found that the differences between theoretical and experimental size distributions were within $10\%$.
226+Other uncertainties of the experimental data are associated with the measurement instruments. Generally the CPC was $10\%$ higher than the total concentration given by SMPS. However, the measurement at the tailpipe at $50$km/h showed a $25\%$ difference between the two instruments, which was taken into account by correcting the SMPS size distribution by this factor. Although there is no clear explanation for this difference the theoretical analysis later also showed that the CPC results were the correct ones. Finally it should be mentioned that depending on the SMPS parameters (or the company of SMPS) small differences at the size distributions can be obtained (mean, standard deviation and total concentration). Although this was not investigated, the effect of different size distribution parameters is discussed in the Discussion section of the paper.
227+\section{Aggregate structure}
228+
291229 \subsection{Effective density and fractal dimension}
230+
292231 Aggregate morphology has a large effect on the transport and thermal properties of
293- aerosol particles. Solid diesel-exhaust carbonaceous particles do not coalesce perfectly
294- but give rise to complex, quasi-fractal structures \citep{friedlander book}.
295-The morphology of large aggregates consisting of a number $n_{\rm{agg}}$ of smaller primary spherules of size $d_0$ is often characterised by their fractal dimension $\df$ defined by:
232+aerosol particles. The morphology of large aggregates consisting of a number $n_{\rm{agg}}$ of smaller
233+primary particles of effective
234+diameter $d_0$ is often characterised by their fractal dimension $\df$ defined by:
296235 \beq\label{rel_nagg}
297236 n_{\rm{agg}}=a\lro\f{\dagg}{d_0}\rro^{\df},
298237 \eeq
299238 where $\dagg$ is the characteristic size of the aggregate.
300-In the experiments, we measured the aggregate number-size distribution using an SMPS. In the SMPS, the aerosol particles are classifed by an electrostatic method, which is based on
301-the electrical mobility of electrically charged particles. Electrical mobility is defined as the velocity
302-that an electrically charged aerosol particle acquires in an electric field of unit strength.
303-If the aerosol
304-particle is non-spherical, the SMPS returns a mobility equivalent diameter that indicates the diameter of
239+The aggregate number distribution was measured with an SMPS; thus,
240+particles were classified according to their electrical mobility.
241+For non-spherical particles the SMPS estimates a mobility equivalent diameter that indicates the diameter of
305242 a sphere with the same electrical mobility as the actual non-spherical particle.
306- We then choose this diameter as the characteristic size of an aggregate $\dagg$.
307-The prefactor $a$, weakly dependent on $\df$ \citep{naumann code}, is often taken as a constant equal to one, as in the present study. The exact value of $a$ is not crucial, since it can be absorbed in a re-definition of $d_0$, the only important parameter being the ratio $a/d_0^{d_f}$.
308-Here we do not allow for a time evolution of the fractal dimension (see at this regard \citet{evolving fractal 1}).
243+We chose this diameter as the characteristic size of an aggregate $\dagg$, in agreement
244+with previous works \citep{maricq experimental}, where
245+the effective density is expressed as a function of the
246+mobility diameter instead of the aggregate radius
247+of gyration. The value of the prefactor $a$, weakly dependent on $\df$ \citep{naumann code},
248+is not crucial, since it can be absorbed in $d_0$,
249+the only important parameter being the ratio $a/d_0^{d_{\rm f}}$.
250+We, thus, take it to be unity.
309251 The mass of an aggregate made up of $\nagg$ primary particles can then be expressed as:
310-\beq\label{mass aggregate}
311-\magg=\nagg \f{\pi}{6}d_0^3=\reff \vagg,
252+\beq
253+\label{mass aggregate}
254+\magg=\nagg \f{\pi}{6}d_0^3 \equiv \reff \vagg,
312255 \eeq
313-where $\vagg=\pi\dagg^3/6$ is the volume of the aggregate and $\reff$ is the effective density given by:
256+where $\vagg=\pi\dagg^3/6$ is the volume of the aggregate and $\reff$ is the effective density given by:
314257 \beq\label{eff dens}
315-\rho_{\rm{eff}}=\rho_0\lro \f{\dagg}{d_0} \rro^{\lro{\df-3}\rro},
258+\rho_{\rm{eff}}=\rho_0 \lro \f{\dagg}{d_0} \rro^{\lro{\df-3}\rro},
316259 \eeq
317260 where $\rho_0$ is the primary particle density.
318261 % The effective density in Eq.~(\ref{eff dens}) is useful to determine the mass of an aggregate once its
319262 % fractal dimension and mobility equivalent size are known.
320-A large body of experimental data (see e.g. \cite{maricq experimental} ) suggests an empirical expression for the effective density of light-duty diesel engine exhaust particles:
263+A large body of experimental data (see, e.g., \cite{maricq experimental, park fractal, virtanen fractal} ) indicate that the power-law in Eq.~(\ref{eff dens}) breaks down at aggregates sizes $\dagg\simeq 50$nm, below which the effective density is approximately constant with value $\reff\simeq 1$g/cm$^3$.
264+Therefore we choose the following expression to interpolate the power-law behavior at large aggregate sizes to the constant value of $\reff$ at for small aggregates:
321265 \beq\label{rho eff revised}
322266 \reff=\min\lro\rho_0,\rho_0\lro \f{\dagg}{d_0} \rro^{\lro{\df-3}\rro}\rro,
323267 \eeq
324268 with $\rho_0\simeq1$g/cm$^3$, $d_0\simeq50$nm, and a fractal dimension in the interval $[1.8,2.8]$.
325-This reflects the fact that aggregates acquire a real fractal nature only when above a size threshold.
326-
269+Equation (\ref{rho eff revised}) implies that the primary particle diameter $d_0$
270+is the aggregate diameter beyond which aggregates become fractal-like.
271+We tested numerically the robustness of the results we obtain against small ($\sim 10-20\%$) variations of $\rho_0$ and $d_0$.
272+% Aggregates of a given effective diameter may be considered fractal objects
273+% beyonf a threshold.
327274
328-\subsection{Particle-size distribution}
275+\subsection{Particle number distribution}
329276
330-Ambient aerosol number-size distributions are typically multi-lognormal distributions \citep{whitby classical}.
331-The SMPS used in the experiments binned the number-size distribution into $103$ logarithmically-spaced channels.
332-Since the experiments showed only a single-modal distribution, the measured particle number in a unit volume, as a function of the aggregate mobility diameter, $dN(\dagg)$ was fitted according to:
277+%Ambient aerosol number-size distributions are typically multi-lognormal distributions \citep{whitby classical}.
278+The SMPS used in the experiments binned the number distribution into $103$ logarithmically-spaced channels.
279+Since the experiments showed only a single-modal distribution, the
280+measured particle number per unit volume, as a function of the aggregate mobility diameter, $dN(\dagg)$
281+was fitted to a single lognormal distribution according to:
333282 \beq\label{lognormal}
334283 dN=
335284 \f{N}{\sqrt{2\pi}\log\sigma}\exp\lsq-\f{(\log \dagg-\log\mu)^2}{2\log^2\sigma}\rsq d\log\dagg,
336285 \eeq
337-where $N$ is the total particle number in a unit aerosol volume and $\mu$ and $\sigma$ are the geometric mean and standard deviation of the distribution, respectively.
338- We choose $N$, $\mu$ and $\sigma$ as fit parameters of the lognormal distribution. The nonlinear
286+where $N$ is the total particle number per unit volume,
287+and $\mu$ and $\sigma$ are the geometric mean and standard deviation of the distribution, respectively.
288+We choose $N$, $\mu$ and $\sigma$ as the fitting parameters of the lognormal distribution. The nonlinear
339289 least-square optimisation used for the fit relies on Levenberg-Marquardt algorithm
340- \citep{numerical recipes} as implemented in the Minpack library of R statistical software \citep{minpack}.
341-Once the number-size distribution in Eq.~(\ref{lognormal}) is known, the corresponding mass-size distribution can be easily obtained:
290+\citep{numerical recipes} as implemented in the Minpack library of R statistical software \citep{minpack}.
291+
292+Once the number distribution in Eq.~(\ref{lognormal}) is known,
293+the corresponding mass distribution can be easily obtained
342294 \beq\label{mass size}
343295 dM=\reff\f{\pi}{6}\dagg^3dN.
344-\eeq
296+\eeq
345297
346298
347-\begin{figure}
348-\begin{center}
349-\includegraphics[width=0.5\columnwidth, height=6cm]{mass-sizes.pdf}
350-\end{center}
351-\caption{Cumulative mass distribution function for the particle-size distributions measured at the inlet for a representative $d_f=2.3$, $d_0=50$nm and $\rho_0=1$g/cm$^3$. Most of the total mass is due to aggregates with $\dagg$ larger than $100$nm. This proves true for any $d_f$ in the range $[1.8,2.8]$.}\label{mass-size distr}
352-\end{figure}
353-The total particle mass in a unit volume, $M_{\rm{tot}}$, is obtained from the integration of Eq.~(\ref{mass size}), after choosing the fractal dimension $d_f$.
354-Although we generally evaluate it numerically, if we express the effective density as in Eq.~(\ref{eff dens}), i.e. we neglect its deviation from a power-law for aggregate sizes below $d_0$, we deduce the approximate expression:
299+
300+The total particle mass per unit volume, $M_{\rm{tot}}$, is obtained from the integration of Eq.~(\ref{mass size}), after choosing the fractal dimension $d_{\rm f}$.
301+Although we generally evaluate it numerically, if we express the effective density as in Eq.~(\ref{eff dens}),
302+i.e. we neglect its deviation from a power-law for aggregate sizes below $d_0$, we obtain the approximate expression:
355303 \beq\label{total mass}
356304 % M=\f{N\pi}{6d_0^{(\df-3)}}\exp\lsq\df\ln\mu_{\rm geo}+\f{\df^2}{2}\ln^2\sigma_{\rm geo}\rsq.
357-M_{\rm tot}=\i_0^{\infty}dM\simeq N\rho_0\f{\pi}{6d_0^{(d_f-3)}}\exp\lro d_f\ln\mu+\f{d_f^2}{2}\ln^2\sigma\rro.
305+M_{\rm tot}=\i_0^{\infty}dM \simeq N\rho_0\f{\pi}{6d_0^{(d_{\rm f}-3)}}\exp\lro d_{\rm f}\ln\mu+\f{d_{\rm f}^2}{2}\ln^2\sigma\rro.
358306 \eeq
359-It turns out that most of the mass is found at aggreagate sizes above $100$nm, irrespective of the fractal dimension, so this estimate is not very sensitive to the behavior of $\reff$ for aggregate diameters smaller than $d_0$. This is shown in Fig.~\ref{mass-size distr} plotting, for a characteristic fractal dimension $d_f=2.3$, the cumulative mass distribution function (CDF) for the initial state measured at the inlet of the transfer line in both the experimental cases investigated, namely car speed $120$ and $50$Km/h. A similar conclusion holds in general for $d_f$ in the range from $1.8$ to $2.8$.
360-In the case of diesel-engine nanoparticles, we expect the smallest generated particles to be of the size of a few nanometers, so we limit the applicability of Eq.~(\ref{rho eff revised}) down to a monomer size $d_{\rm mon}=2$nm.
361-Again, the precise value of $d_{\rm mon}$, in this order of magnitude, is irrelevant for the mass balances and for the aerosol processes considered in this work.
307+It turns out that most of the mass is found at aggregate sizes above $100$nm, irrespective of the fractal dimension: hence,
308+the approximate expression for the total mass, Eq. (\ref{total mass}), is not very sensitive to the behavior
309+of $\reff$ for aggregate diameters smaller than $d_0$.
310+This is shown in Fig.~\ref{mass-size distr} where the normalized cumulative mass distribution function (CDF)
311+at the transfer line inlet is plotted for both experiments, vehicle speed $50$ and $100$Km/h.
312+The normalized mass CDF in Fig.~\ref{mass-size distr} was calculated for a characteristic fractal dimension $d_{\rm f}=2.3$, but
313+a similar conclusion holds in general for $d_{\rm f}$ in the range from $1.8$ to $2.8$.
314+In the case of diesel-engine nanoparticles, the smallest generated particles are of the order of a few nanometers,
315+so we limit the applicability of Eq.~(\ref{rho eff revised}) to a monomer size $d_{\rm mon}=2$nm.
316+Again, the precise value of $d_{\rm mon}$, as long as it is of the same order of magnitude, is irrelevant for the mass balance
317+and for the aerosol processes considered in this work.
318+
362319 % \subsection{Application to the experimental number-size distributions}
363320
364-The statistics described in the previous section can be applied to the experimentally measured number-size distributions.
365-Figure~\ref{experimental number-size} shows the experimental results, together with their statistical uncertainty, at the inlet and outlet of the trasfer line for the two steady-state velocities $50$ and $120$Km/h.
366-For any fractal dimension in the range $[1.8,2.8]$, the calculation of the total particle mass, for steady-state velocity $50$Km/h, leads to an excess mass around $10\%$ at the outlet with respect to the inlet.
367-This is most likely due to the presence of strong pressure fluctuations when sampling at the inlet of the transfer line which can affect the determination of the sampling system dilution ratio.
368-In Section~\ref{numerical method} we show how to deal with that experimental uncertainty.
369-For car speed $120$Km/h, instead, we find about $10\%$ extra particle mass at the inlet. This experimental result is not problematic and, as we show in Section~\ref{results}, can be almost entirely explained by taking into account thermophoretic deposition.
370- \begin{figure}
371-\includegraphics[width=0.5\columnwidth]{experimental-distr-with-error-bars_50.pdf}
372-\includegraphics[width=0.5\columnwidth]{experimental-distr-with-error-bars.pdf}
373-\caption{Left: experimental number-size distributions, together with statistical uncertainty, at inlet and outlet, for car speed equal to $120$Km/h. Right: as on the left, but now the car speed is $120$Km/h.}\label{experimental number-size}
374-\end{figure}
321+A comparison of the lognormal fit of the experimental distributions, along with the experimental statistical uncertainty
322+estimated from repeatability experiments, is shown in Figure~\ref{experimental number-size}.
323+The experimental number size distributions at the inlet and outlet of the transfer line for the two steady-state velocities $50$ and $120$Km/h
324+are shown. The experimental data and the lognormal fits are practically indistinguishable.
325+
326+%In addition to the analysis of the particle size distribution, the total particle mass was calculated,
327+%both via numerical integration and via the approximate expression. For the experiments at the steady-state
328+%velocity $50$Km/h we find that, for any fractal dimension in the range $[1.8,2.8]$,
329+%the calculated total particle mass at the outlet is $10\%$ greater than the mass at the inlet.
330+%This discrepancy is most likely due to strong pressure fluctuations during sampling
331+%at the transfer line inlet: these fluctuations affect the determination of the sampling system dilution ratio,
332+%rendering the corresponding measurements less reliable than the measurements at the outlet.
333+%In Section~\ref{numerical method} we show that the proposed algorithm
334+%for the determination of the average fractal dimension
335+%can be used to reconstruct the initial state, and, hence, how to deal with this experimental uncertainty.
336+%For car speed $120$Km/h, instead, the total calculated particle mass at the inlet is about $10\%$ larger
337+%than at the outlet. We show in Section~\ref{results} that the extra inlet particle mass
338+%can be almost entirely explained by taking into account thermophoretic deposition.
339+
375340
376341
377342
378343 \section{Aerosol dynamics}
344+
379345 \subsection{One-dimensional treatment}
380-The dynamics of solid aerosol particle (e.g. transport and
346+The dynamics of solid aerosol particles (e.g. transport and
381347 agglomeration) in the transfer line is described by the
382-General Dynamic Equation (GDE) for the particle-size distribution
348+General Dynamic Equation (GDE) for the number distribution
383349 (gas-to-particle conversion processes are not considered because
384-we are only interested in the dynamics of solid particles, as described
385-in Section 2). As the main interest of our work is the effect of the
350+we only model dynamics of solid particles).
351+As the main interest of our work is the effect of
386352 aggregate structure on particle dynamics, and the development of a
387353 simplified approach to estimate the main aerosol process
388354 within the transfer line we make a number of approximations to simplify the GDE.
@@ -392,26 +358,22 @@
392358 turbulent GDE an extremely complex task.
393359
394360 The effect of particle inertia on transport and deposition
395-can be estimated by considering the appropriate
396-dimensionless number, the particle
397-Stokes number. The particle response time is given by:
398-\beq\label{tau_part}
399-\tau_p=\f{\rho_0\dagg^2}{18\nu\rho_f},
400-\eeq
401-where $\rho_{f}$ is the fluid density and $\nu$ is the fluid kinematic viscosity.
402-If we consider $\dagg\simeq 70$nm a typical aggregate size, see the corresponding
403-figure of the experimental distributions, the characteristic particle response time
404-becomes $\tau_p\sim 10^{-8}$s. The characteristic fluid time scale in a turbulent flow
405-is the Kolmogorov time scale \citep{pope turbulence}:
406-\beq
407-\tau_f=\sqrt{\f{\nu}{\varepsilon}},
408-\eeq
409-where $\varepsilon$ is the turbulent energy dissipation rate. Computational Fluid Dynamics
410-calculations of the fluid flow in the transfer line estimate the
411-Kolmogorov time-scale to be $\tau_f\sim 10^{-3}$s. Therefore, the
412-corresponding Stokes number becomes ${\rm St}=\tau_p/\tau_f\sim 10^{-5}$.
413-Particle inertia is, thus, not important and particles may be considered fluid points,
414-namely, they follow the fluid streamlines.
361+can be estimated by considering the dimensionless particle
362+Stokes number, the ratio of the particle response time
363+to a characteristic fluid time scale. The particle response time
364+is $\tau_{\rm p} = \rho_0\dagg^2/(18\nu\rho_{\rm f})$, where $\rho_{\rm f}$ is the fluid density
365+and $\nu$ is the fluid kinematic viscosity.
366+If we consider $\dagg\simeq 70$nm a typical aggregate size, cf. Fig. \ref{experimental number-size},
367+the characteristic particle response time
368+becomes $\tau_{\rm p}\sim 10^{-8}$s. The characteristic fluid time scale in a turbulent flow
369+is the Kolmogorov time scale \citep{pope turbulence}, $\tau_{\rm f}=\sqrt{\nu/\varepsilon}$,
370+where $\varepsilon$ is the turbulent energy dissipation rate. We performed
371+Computational Fluid Dynamics calculations (relying on the $k-\epsilon$ turbulence module as implemented in the finite-element software \cite{comsol}) of the fluid flow in the transfer line
372+that estimate the Kolmogorov time-scale to be $\tau_{\rm f}\sim 10^{-3}$s. The
373+corresponding Stokes number becomes ${\rm St}~=~\tau_{\rm p}/\tau_{\rm f}\sim 10^{-5}$.
374+Hence, particle inertia is not important and, in the absence of
375+diffusion, particles may be considered fluid points,
376+in that they follow the fluid streamlines.
415377
416378 % The first, and foremost, approximation is neglect of
417379 % turbulent fluctuations (that break the cylindrical-tube azimuthal
@@ -428,80 +390,87 @@
428390 The effects of instantaneous particle concentration inhomogeneities, solely induced by turbulent
429391 flow fluctuations, on agglomeration are neglected.
430392 In essence, this implies that the carrier flow is not highly turbulent otherwise,
431- even in the case of a simple duct geometry, a
432- two-dimensional treatment becomes unavoidable.
433- Accordingly, the flow field is described by its mean
434-value having only an axial component, $\vec U=(U_z,U_r)=(U_m,0)$, with
435- $U_m$ the mean axial velocity along the transfer line.
436-Furthermore, the experiments were performed with a
437-constant flow field and under steady-state conditions.
438-Axial turbulent diffusion is easily shown to be negligible but radial diffusion and thermophoresis may become significant.
439-They are accounted for by introducing thermophoretic and diffusional velocities directed towards the duct walls.
393+even in the case of a simple duct geometry, a
394+two-dimensional treatment becomes unavoidable.
395+Our Computational Fluid Dynamics calculations estimate the
396+average contribution of fluid velocity fluctuations to the
397+total fluid kinetic energy to be of the order of 1\%.
398+Accordingly, the leading order approximation of
399+the flow field becomes the mean flow with
400+only an axial component, $\vec U=(U_z,U_r)=(U_m,0)$,
401+ $U_m$ being the mean axial velocity along the transfer line.
402+Axial turbulent diffusion is easily shown to be negligible (we estimated a Peclet number ${\rm Pe\simeq 80}$ along the transfer line), but radial
403+diffusion and thermophoresis may become significant.
404+They are accounted for by introducing
405+thermophoretic and diffusional velocities directed towards the duct walls.
406+As the Stokes number is small turbophoresis may be neglected,
407+and temperature fluctuations, induced by fluid flow fluctuations,
408+are also neglected.
440409 Since the emitted particles are in the nanoparticle
441410 regime sedimentation is not significant.
442-% Although but other external processes,
443-% like .
444-% Turbulent
445-% diffusion and correlation of velocity fluctuations
446-% with particle concentration fluctuations are, consequently, neglected.
447-
448-
449-% : particle concentration
450-% is considered radially uniform.
451-
452-
453-
411+Furthermore, the experiments were performed with a
412+constant flow field and under steady-state conditions.
454413
455414 Under these assumptions,
456415 the full GDE may be approximated by a one-dimensional, steady-state
457-equation. Therefore, simple mass balance considerations convert the
458-GDE (a population balance equation) into an effective one-dimensional equation for
416+equation. Therefore, a simple balance on the number concentration,
417+or equivalently an average over the pipe cross section, converts the
418+GDE into an effective one-dimensional equation for
459419 $n_q$, the mean (flux-averaged) axial aggregate concentration
460-(aggregate number per unit volume) of characteristic size $\dagg=d_q$, to
420+(aggregate number per unit volume) of characteristic
421+aggregate size $d_q$, to
461422 \beq
462423 \label{1D-advection-intermediate}
463-U_m \, \f{\p n_q}{\p z}=-\f{2(v_d+v_{th})}{R}n_q + \omega_q(z),
424+U_m \, \f{d n_q}{d z}=-\f{2(v_{\rm dif}+v_{\rm th})}{\rm R}n_q + \omega_q(z),
464425 \eeq
465426
466-where $z$ is the axial coordinate and $R$ is the radius of the transfer line.
427+where $z$ is the axial coordinate and ${\rm R}$ is the radius of the transfer line.
467428 The first term on the right-hand side of Eq.~(\ref{1D-advection-intermediate})
468429 describes changes of the size distribution due to the external processes considered:
469-diffusional and thermophoretic deposition. The corresponding deposition
470-velocities are denoted $v_d$ and $v_{th}$. The last term $\omega_q$
430+turbulent diffusional and thermophoretic deposition. The corresponding deposition
431+velocities are denoted $v_{\rm dif}$ and $v_{\rm th}$. The last term $\omega_q$
471432 describes particle aggregation, the only internal particle processes taken into account.
472433 It is worth pointing out that the aggregation term has no effect on
473434 mass balance since aggregation conserves the total mass; it only affects
474-particle number by reducing the concentration.
435+particle number by reducing its concentration.
475436 A similar argument leads to a one-dimensional steady-state equation for the
476-steady-state, axially averaged fluid temperature (see, for example,
477-Ref.~\cite{yannis 1d deposition}).
437+steady-state, axially averaged fluid temperature [see, for example,
438+Ref.~\cite{yannis 1d deposition}]
439+\beq
440+\label{1D-temperature}
441+\frac{d T_m}{dz} = - \frac{2 \textrm{Nu}}{\rm{R} \textrm{Re Pr}} \, (T_w -T_m) ,
442+\eeq
443+where the mean temperature is $T_m$, the wall temperature $T_w$, \textrm{Re} is
444+the flow Reynolds number, $\textrm{Nu}$ the (turbulent) Nusselt number,
445+and $\textrm{Pr}$ the (turbulent) Prandtl number.
478446
479447 According to Eq.~(\ref{1D-advection-intermediate}) agglomeration becomes
480448 a function of the aggregate position within the tube. Under steady-state conditions
481-the axial position of an aggregate is trivially related to its residence time
449+the axial position of an aggregate is related to its residence time
482450 in the tube, denoted by
483451 $\tau$ to be distinguished from the physical time $t$, via $z=U_m\tau$. Agglomeration, thus, becomes
484452 a function of the residence time of the aggregate along the
485453 duct. In other words, aggregates
486454 at different axial penetrations along the duct spent
487-a longer time inside the pipe, and they are thus more affected by agglomeration.
455+a longer time inside the pipe, and they are, thus,
456+ more affected by agglomeration.
488457 By making the $\tau$-dependence explicit in Eq.~(\ref{1D-advection-intermediate}) one obtains:
489458 \beq\label{smolu-tau}
490-\f{\p n_q(\tau)}{\p\tau}=-\f{2(v_d+v_{th})}{R}n_q(\tau)+\omega_q(\tau),
459+\f{d n_q(\tau)}{d\tau}=-\f{2(v_{\rm dif}+v_{\rm th})}{\rm R}n_q(\tau)+\omega_q(\tau),
491460 \eeq
492461 namely, a Smoluchowski equation in the particle residence time along the duct, $\tau$,
493462 with additional linear terms for particle losses.
494463 We have, thus, mapped the one-dimensional dynamics of aerosol particles convected inside
495464 the transfer line into the dynamics of particles ageing in a chamber.
496465
497-The agglomeration term $\omega_q$ is given by:
466+The collision term $\omega_q$ is given by:
498467 \beq\label{omega q as in books}
499468 \omega_q=\f{1}{2}\sum_{i+j=q}\mathcal{K}_{ij}n_in_j+n_q\sum_i\mathcal{K}_{iq}n_i,
500469 \eeq
501470 where $\mathcal{K}_{ij}$ is the collision kernel between aggregates of solid volume $\nu_i$ and $\nu_j$.
502-For low mass particles the gravitational kernel does not contribute
503-significantly. For aggregates smaller than $1\mu$m, such as in the
504-experiments, and $Re\simeq 8000-16000$, Brownian motion is the leading mechanism
471+The gravitational kernel does not contribute
472+significantly for collisions of nanoparticles. For aggregates smaller than $1\mu$m, such as in the
473+experiments, and $\textrm{Re}\simeq 8000-16000$, Brownian motion is the leading mechanism
505474 for particle collisions \citep{pandis book}.
506475 We, thus, consider the Brownian collision kernel for collisions of aggregates of solid volume $\nu_i$ and $\nu_j$
507476 in the continuum regime. The effect of aggregate structure is
@@ -509,18 +478,17 @@
509478 on the aggregate fractal dimension \citep{continuum formula},
510479 \beq
511480 \label{cont kernel}
512-\mathcal{K}_{ij}^{\rm cont}=\f{2k_BT}{3\mu_{\rm{air}}}\lro \nu_i^{1/\df}+\nu_j^{1/\df}\rro\lro
481+\mathcal{K}_{ij}^{\rm cont}=\f{2k_B T_m}{3\mu_{\rm{air}}}\lro \nu_i^{1/\df}+\nu_j^{1/\df}\rro\lro
513482 \f{C_i}{\nu_i^{1/\df}}+ \f{C_j}{\nu_j^{1/\df}} \rro,
514483 \eeq
515-where $k_B$ is the Boltzmann constant, $T$ the temperature, $\mu_{\rm{air}}$ the air dynamic
516-viscosity, $C_i=1+\Kn_i(1.17+0.53\exp(-0.78/\Kn_i))$
517-the Cunningham slip factor, and $\Kn_i$ is the Knudsen number,
518-that ratio of the air mean free path
519-to the particle diameter, calculated for a particle in the $i$-th bin.
484+where $k_B$ is the Boltzmann constant, $T_m$ the mean temperature, $\mu_{\rm{air}}$ the air dynamic
485+viscosity, $C_i=1+\Kn_i[1.17+0.53\exp(-0.78/\Kn_i)]$
486+the Cunningham slip factor, and $\Kn_i$the Knudsen number,
487+the ratio of the air mean free path
488+to the particle diameter, calculated for a $\nu_i$ particle.
520489 For a characteristic particle size $d_q~\sim ~70$nm, and an exhaust temperature $T~\sim ~380$K,
521-the particle Knudsen number is in between the continuum and the
522-free molecular regime, i.e., in the transition regime.
523-Accordingly, we consider a modification of the continuum Brownian kernel
490+the particle Knudsen number is in the transition regime, between the continuum and the
491+free molecular regime. Accordingly, we consider a modification of the continuum Brownian kernel
524492 via the Fuchs interpolation factor $\beta$ \citep{pandis book},
525493 a quantity considered herein independent of the aggregate fractal dimension. The Fuchs factor
526494 interpolates the kernel from the continuum to the free molecular regime.
@@ -529,65 +497,81 @@
529497 \label{fuchs_kernel}
530498 \mathcal{K}_{ij}=\mathcal{K}_{ij}^{\rm cont}\beta \ .
531499 \eeq
500+The collision kernel depends explicitly on the aggregate fractal dimension.
501+Here we do not allow for a time evolution of the aggregate
502+fractal dimension. The bivariate distribution was
503+studied extensively \citet{evolving fractal 1}) where they
504+found that the fractal dimension eventually settles to an average
505+fractal dimension as long as agglomeration progresses.
506+This is surely the case of the experiments we model in this study, since the
507+typical size of the aggregates is well above the few nanometers typical of the smallest diesel combustion-generated nanoparticles.
532508
533509 For low Stokes number the effect of homogeneous turbulence
534-on particle collisions can be modelled by the Saffman and Turner kernel \citep{pandis book}.
535-The ratio of the continuum to the turbulent kernels may be easily estimated
510+on particle collisions can be modelled by the Saffman and Turner kernel \citep{saffman}.
511+The ratio of the continuum to the turbulent kernel may be easily estimated
536512 for typical experimental conditions to be:
537513 \beq
538-\f{\mathcal{K^{TS}}}{\mathcal{K}^{\rm cont}}=\f{3\mu_{\rm air}(\pi\varepsilon/120\nu)^{1/2}\dagg^3}{k_BT}\sim 10^{-4},
514+\f{\mathcal{K^{TS}}}{\mathcal{K}^{\rm cont}}=\f{3\mu_{\rm air}(\pi\varepsilon/120\nu)^{1/2}\dagg^3}{k_B T_m}\sim 10^{-4},
539515 \eeq
540-i.e. agglomeration is dominated by Brownian collisions and the role of turbulence
516+namely, the collision kernel is dominated by Brownian collisions, and the role of turbulence
541517 in particle collisions is negligible, in agreement with usual estimates \citep{pandis book}.
542518
543519 The kernel in Eq.~(\ref{fuchs_kernel}) and the thermophoretic and diffusional deposition velocities are
544-$\tau$-dependent. Indeed the temperature $T$ varies along the transfer line as a function of the position $z=U_m\tau$
545-and so do, as a consequence, air viscosity, Knudsen number and particle diffusion coefficient.
546-In the effective one-dimensional transport equation (\ref{smolu-tau}), we are neglecting the effect of turbulent
520+$\tau$-dependent. Indeed the mean temperature $T_m$ varies along the transfer line as a function of the position $z=U_m\tau$,
521+Eq. (\ref{1D-temperature}), and so do, as a consequence, air viscosity, Knudsen number and particle diffusion coefficient.
522+In the effective one-dimensional transport equation (\ref{smolu-tau}), we neglect the effect of turbulent
547523 collisions which could introduce a radial dependence in the agglomeration process.
548-The thermophoretic and diffusional velocities [see Eqs.~\ref{expression vt}-\ref{expression vd}], are estimated by means of one-dimensional correlations.
549-We resort to a correlation also for the calculation of the mean one-dimensional temperature along the anaconda \citep{yannis 1d deposition}.
550-The details about all the correlations used in this study can be found in appendix \ref{appendix corr}.
524+The thermophoretic and diffusional velocities, as well as the one-dimensional
525+mean temperature, are estimated by means of one-dimensional correlations.
526+% Details are summarized in appendix \ref{appendix corr}, see, also, Ref.~\citep{yannis 1d deposition}.
527+The thermophoretic velocity is calculated as:
528+ \beq
529+ \vth= -\kt\;\f{\nu}{T}\f{T_w-T}{2{\rm R}/\nusselt},
530+ \eeq
531+
532+
551533
552534 The relative importance of the aerosol processes in
553535 Eq.~(\ref{smolu-tau}) is most conveniently estimated by
554536 recasting it in a dimensionless form.
555-Convection fixes the total residence time of the aggregates in the transfer line, $\tau_{\rm res}=L/U_m$, where $L$ is the total length of the transfer line.
556- We introduce appropriate
557-characteristic time scale for agglomeration, $\tau_{\rm ag}$, thermophoretic
537+Convection fixes the total residence time of the aggregates in the transfer line, $\tau_{\rm res}=L/U_m$,
538+where $L$ is the total length of the transfer line.
539+We introduce appropriate characteristic time scale for agglomeration, $\tau_{\rm ag}$, thermophoretic
558540 deposition, $\tau_{\textrm{th}}$, and diffusional deposition,
559541 $\tau_{\textrm{dif}}$, as follows
560542 \begin{align}
561543 \label{dimensionless times}
562-\tau_{\rm ag} &\equiv \lro N_0\f{2k_BT_0}{3\mu_{\rm{air}}}\beta(\mu_0,T_0)\rro^{-1},
563- &\tau_{\rm{dif}} \equiv &
564- \gdif^{-1}(\mu_0,T_0) ,&\tau_{\rm{th}} \equiv & \gth^{-1}(\mu_0^,T_0),
544+\tau_{\rm ag}^{-1} &\equiv N_0 \, \f{2k_BT_0}{3\mu_{\rm{air}}} \, \beta(\mu_0,T_0) ,
545+ &\tau_{\rm{dif}}^{-1} \equiv &
546+ \gdif(\mu_0,T_0) ,&\tau_{\rm{th}}^{-1} \equiv & \gth(\mu_0, T_0),
565547 \end{align}
566548
567-
568549 % \begin{align}
569550 % \label{dimensionless times}
570551 % \nonumber \tau_{\rm ag} &\equiv \lro N_0\f{2k_BT_0}{3\mu_{\rm{air}}}\beta(\mu_0,T_0)\rro^{-1},
571552 % &\tau_{\rm{dif}} \equiv &
572553 % \gdif^{-1}(\mu_0,T_0) ,&\tau_{\rm{th}} \equiv & \gth^{-1}(\mu_0^,T_0),
573554 % \end{align}
574-where $N_0$, $\mu_0$ and $T_0$
575-are the total number of aggregates, arithmetic mean of the number-size distribution, and the
576-average temperature at the inlet, respectively. The thermophoretic and
555+where $N_0$ are the total number of aggregates per unit volume,
556+$\mu_0$ the arithmetic mean of the number distribution, and
557+$T_0$ the mean temperature at the inlet. The thermophoretic and
577558 diffusional time scales are determined from the corresponding deposition
578-velocities via $\gamma_{{\rm th(dif)}}=2v_{{\rm th(dif)} }/R$.
559+velocities via $\gamma_{{\rm th(dif)}}=2v_{{\rm th(dif)} }/{\rm R}$.
579560 The agglomeration time scale, $\tau_{\rm ag}$, is a
580-generalization, through Fuchs interpolation factor, of
581-the agglomeration characteristic time used in \cite{pratsinis preserving adimensional}.
582- Its inverse value is obtained multiplying the initial total aggregate concentration, $N_0$, by the collision kernel evaluated for a typical temperature and aggregate size, a quantity we call $\mathcal{K}$.
583-We illustrate the role of $\tau_{\rm ag}$ by considering the case of an initially monodispersed aerosol interacting with a constant kernel $\mathcal{K}$.
584-One can prove that \citep{friedlander book} the total aerosol concentration decreases according to:
561+generalization, through the Fuchs interpolation factor, of
562+the characteristic agglomeration time used in \cite{pratsinis preserving adimensional}.
563+Its inverse is obtained by multiplying the initial total aggregate concentration, $N_0$,
564+times the collision kernel evaluated at a typical temperature and aggregate size,
565+a quantity we call $\mathcal{K}_0$.
566+The role of $\tau_{\rm ag}$ becomes apparent
567+by considering an initially monodispersed aerosol interacting with a constant kernel $\mathcal{K}_0$.
568+Then, the total aerosol concentration decreases according to \citep{friedlander book}
585569 \beq\label{explanation tau agglomeration}
586-N(\tau)=\f{N_0}{1+\mathcal{K}N_0t/2},
587-\eeq
588-which shows the role of the $\tau_{\rm ag}=1/(\mathcal{K}N_0)$ as the appropriate time scale for agglomeration.
570+N(\tau)=\f{N_0}{1+\mathcal{K}_0 N_0 t/2} .
571+\eeq
572+Hence, $\tau_{\rm ag}=1/(\mathcal{K}_0 N_0)$ is the appropriate time scale for agglomeration.
589573
590-Equation~(\ref{smolu-tau}) expressed in terms of following dimensionless
574+Equation~(\ref{smolu-tau}) expressed in terms of the dimensionless
591575 variables
592576 \begin{align}
593577 \label{dimensionless variables}
@@ -597,7 +581,7 @@
597581 becomes
598582 \beq
599583 \label{smolu-dimensionless}
600-\f{\p \tilde n_q}{\p\tilde \tau}= \f{\tau_{\rm res}}{\tau_{\rm ag}}\lro\f{1}{2}\s_{i,j}^{N_b}\tilde{\mathcal{K}}_{ij}\tilde n_i\tilde n_j-\s_i^{N_b}\tilde{\mathcal{K}}_{iq}\tilde n_i\tilde n_q\rro
584+\f{d \tilde n_q}{d\tilde \tau}= \f{\tau_{\rm res}}{\tau_{\rm ag}}\lro\f{1}{2}\s_{i,j}^{N_b}\tilde{\mathcal{K}}_{ij}\tilde n_i\tilde n_j-\s_i^{N_b}\tilde{\mathcal{K}}_{iq}\tilde n_i\tilde n_q\rro
601585 -\f{\tau_{\rm res}}{\tau_{\rm{dif}}} \tilde n_q -\f{\tau_{\rm res}}{\tau_{\rm{th}}} \tilde n_q.
602586 \eeq
603587 % Equation~(\ref{smolu-dimensionless}) explicitly shows that there are independent time-scales
@@ -605,196 +589,382 @@
605589 % processes included in this study. Furthermore, it shows that there are two important
606590 % combinations the may be used to estimate the relative importance of each process,
607591 % a procedure that we follow in the Results and Discussion Section.
608-Equation~(\ref{smolu-dimensionless}) explicitely shows that there are four independent time scales that govern the system dynamics, three corresponding to the aerosol
609-processes included in this study and one given by convection. Furthermore, it shows that there are three important
610-combinations the may be used to estimate the relative importance of each process,
592+Equation~(\ref{smolu-dimensionless}) explicitly shows the four independent time scales
593+that govern the system dynamics, three corresponding to aerosol
594+processes and one determined by convection. Furthermore, it shows that there are three important
595+combinations the may be used to estimate the relative importance of the processes,
611596 a procedure that we follow in Section \ref{results}.
612597
613-
598+\subsection{Numerical method}
599+\label{numerical method}
614600
615-\subsection{Numerical method and determination of the fractal dimension} \label{numerical method}
616-The dynamics of particles on a nanosize scale dispersed throughout the fluid is governed by the aerosol general dynamic equation (GDE) \citep{friedlander book}.
617-Its formulation for a discrete size distribution amounts to a population balance equation for each particle size. Such a formulation quickly becomes computationally unpractical when the number of primaries in an aggregate reaches $n_{\rm{agg}}\sim 1000$.
618-Here we follow a sectional approach: the initial number-size distribution is split into $N_b$ bins and particle properties are averaged on each bin \citep{garrick section, jacobson sectional}.
619-Despite relying heavily on the lognormality of the number-size distribution for mass balances and statistics, the sectional method does not require any assumption about its functional form.
620-Equation~(\ref{smolu-tau}) is solved on a logarithmically-spaced bin structure.
621-Each bin is labelled according to the average size of the aggregate it contains.
622-After choosing an interval $[d_{\rm mon},d_{\rm{max}}]$, where $d_{\rm{max}}$ is the maximum aggregate size included in the simulation, we fix the grid in aggregate diameters:
601+The formulation of the GDE for a discrete size distribution,
602+a population balance equation for every particle size,
603+quickly becomes computationally impractical when the number of primary
604+particles in an aggregate reaches $n_{\rm{agg}}\sim 1000$.
605+Herein, we follow a sectional approach: the initial number distribution is
606+split into $N_b$ bins and particle properties are averaged on each
607+bin \citep{jacobson sectional,garrick section}.
608+%Even though the experimental size distributions
609+%are accurately represented by a lognormal distribution we chose a numercial
610+%scheme to solve the GDE, the sectional method, that does not require any assumption about its functional form.
611+Equation~(\ref{smolu-tau}) is solved on a logarithmically-spaced bin structure, where
612+each bin is labelled according to the average aggregate size.
613+For a diameter interval $[d_{\rm mon},d_{\rm{max}}]$, where $d_{\rm{max}}$ is the maximum
614+aggregate size included in the simulation, we fix the grid in aggregate diameters
623615 \beq
624616 d_q=\exp\lsq \f{(q-1)\ln(d_{\rm{max}}/d_{\rm mon})}{N_b}\rsq d_{\rm mon} ,\;\;\;\; q=1\dots N_b+1.
625-\eeq
626-The $q$-th bin is given by the interval $[d_q,d_{q+1}]$ and its population $n_q$ can be determined from the numerical integration of the lognormal distribution in Eq.~(\ref{lognormal}).
627-We identify a characteristic size in each bin $\overline{d}_q=(d_q+d_{q+1})/2$ and a corresponding aggregate mass $\overline{m}_q=\reff\pi\overline{d}_q^3/6$ and aggregate volume of solid $\nu_q=\overline{m}_q/\rho_0$.
617+\eeq
618+The $q$-th bin is the interval $[d_q,d_{q+1}]$ and the number concentration $n_q$
619+is obtained by integrating numerically the lognormal distribution in Eq.~(\ref{lognormal}).
620+We identify a characteristic size in each bin $\overline{d}_q=(d_q+d_{q+1})/2$,
621+an aggregate mass $\overline{m}_q=\reff\pi\overline{d}_q^3/6$, and
622+an aggregate solid volume $\nu_q=\overline{m}_q/\rho_0$.
628623
629-The $\omega_q$ term in Eq.~(\ref{smolu-tau}) accounts for the effect of particle collisions.
630-We now re-cast it in a suitable form for the application of the sectional method on a non-uniform grid:
624+The collision term $\omega_q$ in Eq.~(\ref{smolu-tau})
625+may be rewritten into a form suitable for the application of the sectional method on a non-uniform grid
631626 \beq\label{omega_q}
632627 \omega_q=\f{1}{2}\s_{i,j}^{N_b}\chi_{ijq}\mathcal{K}_{ij}n_in_j-\s_i^{N_b}\mathcal{K}_{iq}n_in_q,
633628 \eeq
634-where $\chi_{ijq}$ is the splitting operator (introduced by \citet{jacobson sectional} and here implemented as in \citet{garrick section}) which accounts for the non-uniformity of the volume grid and re-distributes the newly created aggregates into different bins.
635-Equation~(\ref{smolu-tau}) becomes a system of first-order nonlinear coupled ODE's we solve using the LSODA algorithm \citep{lsoda1,lsoda2}.
636-We tested the procedure by comparing the numerical solution to analytical expressions when the latter are available.
637-The accuracy of the solution is checked by ensuring its independency on the number of bins used to discretize the initial number-size distribution.
638-In the absence of diffusional and thermophoretic losses, Eq.~(\ref{smolu-tau}) conserves the total aerosol particle mass in a unit volume. This is also tested in our numerical simulations to make sure that the employed grid contains all the significant particle sizes for the system in exam.
639-Finally, the use of the splitting operator $\chi_{ijk}$ has been tested by comparing the results obtained with a uniform grid and $6000$ bins to those we get using $100$ logarithmically spaced bins.
640-They turn out to be equivalent, but the numerical advantage of such a reduction in the number of bins, i.e. equations to solve, is obvious.
629+where $\chi_{ijq}$ is the splitting operator (introduced by \citet{jacobson sectional}, here implemented as
630+in \citet{garrick section}) that accounts for the non-uniformity of the volume grid
631+by re-distributing newly created aggregates into different bins.
632+Equation~(\ref{smolu-tau}) becomes a system of first-order nonlinear coupled
633+ODE's that we solve using the LSODA algorithm \citep{lsoda1}.
634+We tested the numerical algorithm
635+by comparing the numerical solution to analytical expressions, when available,
636+e.g. for a constant collision kernel.
637+The accuracy of the solution was checked by ensuring numerical
638+results were independent of the number of bins used to discretize the
639+initial number distribution.
640+In the absence of particle deposition, Eq.~(\ref{smolu-tau}) conserves
641+total particle mass concentration. Mass conservation was tested
642+to ensure that the chosen grid contained all relevant particle sizes.
643+Finally, the splitting operator $\chi_{ijk}$ was tested by comparing results obtained in
644+a uniform grid of $6000$ bins and on a grid of $100$ logarithmically spaced bins.
645+Numerical results were identical, but the computational advantage
646+of the reduction in the number of bins,
647+i.e. equations to solve, was noticeable.
641648
642-In the experiments, the uncertainty associated with the determination of the dilution ratio at the inlet of the transfer line, connected directly to the car tailpipe, is larger than the one at the outlet.
643-This affects the determination of the total concentration $N_{\rm in}$ at the inlet, but not the estimates of the geometric mean and standard deviation.
644-After choosing a fractal dimension $d_f$ we first use Eq.~(\ref{total mass}) to determine the total mass at the outlet $M_{\rm out}$.
645-The same procedure is then applied at the inlet. Of course, the uncertainty about $N_{\rm in}$ leads to the same uncertainty in the total mass at the inlet, $M_{\rm in}$.
646-As a first approximation, one can reconstruct $N_{\rm in}$ by imposing $M_{\rm in}=M_{\rm out}$ and adjusting the total particle concentration at the inlet accordingly.
647- It is also possible to account for some extra mass at the inlet if particle losses along the transfer line are not negligible.
648-The fractal dimension (which is not experimentally determined) affects both the mass balances and the particle dynamics.
649-We determine it self-consistently until a good agreement (i.e. minimization of the least-square or $L_2$-norm distance) is reached between the mass-size distribution
650-returned by our numerical simulations and its value determined from the fit of the corresponding experimental number-size distribution (see Eq.~(\ref{mass-size distr})).
651-The mass-size distribution is favoured over the number-size distribution since the methodology developed in this study relies on mass balances.
652-The algorithm we follow is sketched in Fig.~\ref{diagram for d_f}.
653- \begin{figure}
654-\includegraphics[width=\columnwidth]{diagram.pdf}
655-\caption{Algorithm used to determine the fractal dimension of the aggregates.}\label{diagram for d_f}
656-\end{figure}
649+\subsection{Determination of the aggregate fractal dimension}
650+\label{DetermineFractalDimension}
651+
652+The algorithm we use for the self-consistent calculation
653+of the average fractal dimension is based on the strong
654+dependence of the system dynamics, through the collision kernel, on the fractal dimension.
655+In fact, the kernel in Eq.~(\ref{fuchs_kernel}) exhibits a $1/d_{\rm f}$ power-law dependency on the fractal dimension, leading to enhanced collsions between aggregates with the same volume of solids but lower $d_{\rm f}$.
656+As in \cite{maricq coagulation}, we treat both the total aggregate
657+number concentration and the fractal dimension as fitting parameters.
658+However, in our experiments, the uncertainty of the overall aggregate concentration
659+is probably greater at the transfer-line inlet than at
660+the outlet. This is probably due to strong pressure fluctuations
661+during sampling at the inlet that affect the determination of the
662+sampling-system dilution ratio. The geometric mean and standard deviation of the
663+experimental distribution, however, may be considered weakly dependent on
664+the dilution-ratio uncertainty. Consequently, we base out algorithm on the assumption
665+that outlet measurements are more accurate than inlet measurements.
666+
667+Accordingly, an aggregate fractal dimension is chosen, and then
668+the total particle mass per unit volume at the inlet, $M_{\rm in}$, and at the outlet, $M_{\rm out}$,
669+are calculated via Eq.~(\ref{total mass}). Unlike Ref.~\cite{maricq coagulation}, we explicitly
670+impose, as a first approximation, mass conservation from inlet to
671+outlet and reconstruct $N_{\rm in}$ such that $M_{\rm in} = M_{\rm out}$.
672+This reconstruction uses the chosen fractal dimension, and the experimentally determined
673+$\mu$ and $\sigma$ at the inlet to determine $N_{\textrm{in}}$.
674+If the dynamics shows non-negligible particle losses, we account for the
675+additional mass concentration at the inlet.
676+We then propagate the reconstructed inlet number distribution for
677+the appropriate residence time, thus obtaining the outlet number distribution
678+and the mass distribution (calculated via the effective density [see Eq.~(\ref{mass size})]).
679+The experimental mass distribution at the outlet is similarly calculated
680+using the lognormal fit of the experimental number distribution.
681+Since we rely explicitly on mass-balance considerations to fit $N_{\rm in}$,
682+we iterate the procedure described above for different values of $d_{\rm f}$
683+until we minimize the least-square (or $L_2$-norm) distance
684+between the experimental and simulated mass distributions at the outlet.
685+Another natural choice would be the minimization
686+the least-square distance between the experimental and calculated number distributions.
687+The two minimizations are not equivalent in general and a comparison
688+between them is shown in Section \ref{results}.
689+The algorithm we follow is sketched in Fig.~\ref{diagram for d_f}.
690+
691+% after choosing the fractal dimension, outlet particle distribution is calculated propagating
692+% the inlet distribution for the appropriate
693+% residence time. Due to the experimental uncertainties, we consider
694+% A similar methodology was used in Ref.~
657695
658696
659-% \subsection{Statistical Treatment of Particle Distributions }
660-\section{Results}\label{results}
661-We investigated the dynamics of diesel-exhaust aggregates emitted by a Euro-3 vehicle at constant speeds $50$ and $120$Km/h.
662-Table~\ref{table time scales} shows a the relevant time scales.
663-At $120$Km/h the dynamics is dominated by agglomeration and thermophoretic deposition, whereas diffusional losses are not significant.
664-The exhaust velocity inside the transfer line is $4.3-4.5$m/s, the line being about $9$m long, which leads to a total aggregate residence time $\tau_{\rm res}~\simeq 2$s.
665-The exhaust mean temperature at the inlet can be estimated as about $200\,^{\circ}\mathrm{C}$.
666-Thermophoretic losses are estimated to account for about $7\%$ of the total mass.
667-% In \cite{lee_deposition}, a correlation similar to the one from \cite{yannis 1d deposition}, i.e. the one employed in this study, was used to estimate thermophoretic deposition in a tailpipe.
668-% There it was found to be insufficient but the experimental conditions (decreasing instead of constant wall temperature) were, as pointed out by the authors, outside the realm of validity of the correlation used.
669-% It is simply worth pointing out that \cite{lee_deposition} also found thermophoretic deposition to be non-negligible for inlet temperatures and number-size distributions comparable with those measured at the VELA-2 laboratories.
670-
671-The experiments at car speed equal to $50$Km/h show some similarities in the role of agglomeration on particle dynamics.
672- The total residence time of aerosol aggregates inside the transfer line is $\tau_{\rm res}\simeq4-4.5$s, due to the lower exhaust flow rate, and is again comparable to the characteristic time for agglomeration.
673-The exhaust initial temperature is about $60\,^{\circ}\mathrm{C}$ lower than in the case of the $120$Km/h.
674-This significantly alters the role played by thermophoresis, which is mirrored by a new value of $\tau_{\rm th}~\sim 10^2$s and in this case thermophoretic losses amount to roughly $1\%$ of the total particle mass.
675-% With a characteristic time $\tau_{\rm ag}~\sim 4$s, agglomeration remains by far the dominating process.
676-Figure~\ref{mass-size fig} shows the numerical results in this study.
677- For the experiments at car speed $120$Km/h (top row), we estimated a fractal dimension $\df=2.2$. One observes the combined effects of agglomeration and losses in decreasing particle concentration and increasing the mean of the number-size distribution (left).
678-On the right we also show the mass-size distribution [see Eq.~(\ref{mass-size fig})] as obtained from the numerics and from fitting the experimental number-size distribution at outlet.
679-For car speed equal to $50$Km/h, we determine the fractal dimension to be $\df=2$ and we show the quantities of interest in the bottom row of Fig.~\ref{mass-size fig} as for the $120$Km/h case.
680-\begin{table}
681-\begin{center}
682-\begin{tabular}{|c|r@{}lr@{}lr@{}l||r|}
683- \hline
684-\multicolumn{8}{|c|}
685- {\rule[-3mm]{0mm}{8mm} \textbf{Time scales for the two sets of experiments}} \\
686- Car Speed
687- & \multicolumn{2}{|c|}{Agglomeration}
688- & \multicolumn{2}{|c|}{Diffusion}
689- & \multicolumn{2}{|c|}{Thermophoresis}
690- & Residence Time \\ \hline \hline
691-50 Km/h & \multicolumn{2}{|c|}{$\tau_{\rm ag}\simeq 4$s} & \multicolumn{2}{|c|}{$\tau_{\rm dif}\simeq 10^3$s} &\multicolumn{2}{|c|}{$\tau_{\rm th}\simeq 10^2$s} & $\tau_{\rm res}\simeq 4$s \\
692-120 Km/h & \multicolumn{2}{|c|}{$\tau_{\rm ag}\simeq 2$s} & \multicolumn{2}{|c|}{$\tau_{\rm dif}\simeq 10^3$s} &\multicolumn{2}{|c|}{$\tau_{\rm th}\simeq 13$s} & $\tau_{\rm res}\simeq 2$s \\
693- \hline
694-% \vspace*{0.cm}
695-\end{tabular}
696- \caption{Characteristic time scales for the aerosol processes considered in the two sets of experiments and residence time of the aggregates in the transfer line.}
697- \label{table time scales}
698-\end{center}
699-\end{table}
700-\begin{figure}
701-\includegraphics[width=0.5\columnwidth, height=6cm]{2D-distribution.pdf}
702- \includegraphics[width=0.5\columnwidth, height=6cm]{2D-mass_fitting.pdf}
703-% \includegraphics[width=6.cm, height=6cm]{errors_120.pdf}
704-% \includegraphics[width=10.cm, height=6cm]{evolution_3d_120.pdf}
705- \includegraphics[width=0.5\columnwidth, height=6cm]{2D-distribution_50.pdf}
706- \includegraphics[width=0.5\columnwidth, height=6cm]{2D-mass_fitting_50.pdf}
707-\caption{Top: results for for fractal dimension $\df=2.2$ and car speed $120$Km/h. Top left: fit of the measured number-size distribution at the inlet (dashed line) together with experimental (crosses) and numerical (continuous line) findings at the outlet.
708-Top right: mass-size distribution derived from the corresponding fitted number-size distributions both at the inlet (dashed line) and at the outlet (crosses) together with the output of our simulations at the outlet (continuous line).
709-Bottom: as on top, but results for $\df=2$ and car speed $50$Km/h. }\label{mass-size fig}
710-\end{figure}
711697
712698
713-The dependence of the distance (in a least-square sense) on $\df$, between the calculated and fitted mass-size distributions at outlet, is shown in Fig.~\ref{mass-size residuals}.
714-On the left, we plot the square root of the sum of the residuals of the mass-size distribution as a function, normalized by the total mass, of the fractal dimension $\df$.
699+% The proposed algorithm for the self-consistent calculation
700+% of the average fractal dimension is based on the strong
701+% dependence of the collision kernel, and consequently
702+% on the propagation of the number distribution
703+% along the transfer line, on the fractal dimension.
704+% Specifically, an inlet fractal dimension is chosen,
705+% the inlet distribution is propagated for the appropriate
706+% residence time, and the calculated outlet particle
707+% size distribution is compared to the experimental distribution.
708+% If they differ, a new initial fractal dimension is chosen till the
709+% difference of the calculated and experimental outlet distributions
710+% is minimized. A similar procedure was used by Maricq to
711+% determine the fractal dimension and the total particle number (CHECK).
712+
713+% However, for the experiments at steady-state velocity $50$Km/h we found that
714+% the calculated total particle mass at the outlet
715+% was $5\%$ greater than the mass at the inlet
716+% for a typical fractal dimension of $d_f = 2.0$.
717+% For a vehicle speed of $120$Km/h, instead, the total calculated particle mass
718+% at the inlet was about $10\%$ larger than at the outlet.
719+% We attribute this discrepancy to the experimental uncertainty associated with the
720+% determination of the dilution ratio at the inlet of the transfer line.
721+% This uncertainty is greater at the inlet than at
722+% the outlet, as the transfer line is directly connected
723+% to the tail pipe. It is most likely due to strong pressure fluctuations
724+% during sampling at the transfer line inlet: these fluctuations
725+% affect the determination of the sampling system dilution ratio,
726+% rendering the corresponding measurements less reliable than the
727+% measurements at the outlet. In Section~\ref{results} we
728+% show that the additional inlet particle mass can be
729+% almost entirely explained by taking into account thermophoretic
730+% deposition. The greater outlet mass is dealt with by
731+% reconstructing the initial state such that the total inlet
732+% particle mass is equal (or greater in case of significant
733+% deposition losses) than the outlet mass.
734+% It is, thus, necessary not only to determine
735+% the fractal dimension self consistently but the initial state
736+% has to be reconstructed, a reconstruction based on the calculation
737+% of the particle mass distribution.
738+
739+% The algorithm is modified accordingly. We choose a fractal
740+% dimension $d_f$ and then we use Eq.~(\ref{total mass}) to calculate the
741+% total mass at the outlet $M_{\rm out}$. The same procedure is then
742+% applied at the inlet.
743+
744+% As a first approximation, we reconstruct
745+% $N_{\rm in}$ by requiring $M_{\rm in}=M_{\rm out}$ and adjusting the total
746+% particle concentration at the inlet accordingly.
747+% We consider that the
748+% uncertainty in the sampling dilution ratio
749+% affects only the determination of the total
750+% concentration $N_{\rm in}$ at the inlet,
751+% but not the estimates of
752+% the geometric mean and standard deviation:
753+% hence the reconstructed initial state retains the
754+% experimental $\mu$ and $\sigma$.
755+% Of course, the uncertainty of $N_{\rm in}$
756+% leads to the same uncertainty in the total mass at the inlet,
757+% $M_{\rm in}$. It is also possible
758+% to account for additional mass at the inlet if particle losses along
759+% the transfer line are not negligible.
760+% Consequently we determine the (average) fractal dimensions
761+% self-consistently until a
762+% good agreement (i.e. minimization of the least-square or $L_2$-norm
763+% distance) is reached between the mass-size distribution returned by
764+% our numerical simulations and its value determined from the fit of
765+% the corresponding experimental number-size distribution (see
766+% Eq.~(\ref{mass-size distr})).
767+% As the fractal dimension affects both mass balance and
768+% particle dynamics the minimization procedure could be used
769+% either on the particle size distribution or the mass-size distribution.
770+% The mass-size distribution is favoured
771+% over the number-size distribution since the methodology developed
772+% in this study relies on mass balances (??) and also
773+% the figure in the next section (Rewrite!!!). A sketch of the
774+% algorithm we follow is shown in Fig.~\ref{diagram for d_f}.
775+
776+
777+
778+\section{Results}
779+\label{results}
780+
781+We compare experimental and calculated number distributions
782+for two sets of experiments performed at constant speeds $50$ and $120$Km/h
783+of a Euro-3 light-duty vehicle. Our results are summarized
784+in Table~\ref{table statistics} that shows the statistics of the number
785+distributions and the calculated fractal dimensions.
786+The difference between $N_{\rm in}$, as calculated from the experimental data (fit)
787+and as reconstructed according to mass-balance considerations (rec), amounts to only
788+a few percents, and it is within experimental uncertainties. The same, of course,
789+is true for the total inlet mass concentration. Note, however,
790+that the fitted inlet mass concentration is smaller than
791+the outlet mass concentrations (experiments at $50$Km/h).
792+For the experiments at car speed $120$Km/h, we
793+estimated a fractal dimension $\df=2.2$, whereas
794+$d_{\rm f} = 2$ the $50$Km/h experiments.
795+
796+Mass losses are almost entirely due to thermophoretic deposition. This is illustrated in
797+Table~\ref{table time scales} where the relevant time scales are presented.
798+At steady-state speed $120$Km/h the dynamics is dominated by agglomeration and
799+thermophoretic deposition, whereas diffusional losses are not
800+significant. The exhaust velocity inside the $9$m long transfer line is
801+$4.3-4.5$m/s, leading to a total aggregate residence time $\tau_{\rm res}~\simeq 2$s.
802+The mean exhaust temperature at the inlet can be estimated to be
803+$200\,^{\circ}\mathrm{C}$. Thermophoretic losses are estimated to
804+account for about $7\%$ of the total mass.
805+% In \cite{lee_deposition}, a correlation similar to the one from \cite{yannis 1d deposition}, i.e.
806+%the one employed in this study, was used to estimate thermophoretic deposition in a tailpipe.
807+% There it was found to be insufficient but the experimental conditions (decreasing instead of constant wall temperature)
808+%were, as pointed out by the authors, outside the realm of validity of the correlation used.
809+% It is simply worth pointing out that \cite{lee_deposition} also found thermophoretic deposition to be non-negligible for inlet t
810+%emperatures and number-size distributions comparable with those measured at the VELA-2 laboratories.
811+The experiments at $50$Km/h show similarities in the role of agglomeration on particle dynamics.
812+The aggregate residence time is $\tau_{\rm res}\simeq4-4.5$s, due to the lower exhaust flow rate:
813+as before it is comparable to the characteristic time for agglomeration.
814+The exhaust initial temperature is about $60\,^{\circ}\mathrm{C}$ lower than the $120$Km/h experiments.
815+This significantly alters the role played by thermophoresis,
816+a change that is mirrored in a new value of $\tau_{\rm th}~\sim 10^2$s:
817+in these experiments thermophoretic losses amount to roughly $1\%$ of the total particle mass.
818+% With a characteristic time $\tau_{\rm ag}~\sim 4$s, agglomeration remains by far the dominating process.
819+
820+The evolution of the number and mass distributions is plotted in Fig.~\ref{mass-size fig}
821+for both experiments at $120$Km/h (top row) and $50$Km/h (bottom row).
822+The combined effects of agglomeration and losses in decreasing particle concentration and increasing
823+the mean of the number-size distribution (left) are easily noted (the same effect is shown in
824+Table~\ref{table statistics}). On the right we also show the
825+mass distributions [see Eq.~(\ref{mass-size fig})] as obtained from the simulations and the fit of
826+the experimental outlet number distribution.
827+
828+The difference between the calculated and fitted
829+mass distributions at the outlet, defined as the
830+distance (in a least-square sense, or equivalently the $L_2$ norm),
831+is shown in Fig.~\ref{mass-size residuals}.
832+On the left we plot (in arbitrary units) the square root of the
833+sum of the residuals of the mass distribution
834+as a function of the fractal dimension $\df$.
715835 This function exhibits a clear minimum justifying its use for the self-consistent determination of the fractal dimension.
716-On the right, we carry out the same calculations replacing the mass-size with the number-size distribution.
717-We notice that in particular for car speed $50$Km/h, the resulting function is much flatter and the determination of the minimum is less obvious. This is somehow expected, since the number-size distribution does not distinguish between small and large sizes, thus weighting equally every aggregate size.
718-\begin{figure}
719-\includegraphics[width=0.5\columnwidth, height=6cm]{errors_mass.pdf}
720-\includegraphics[width=0.5\columnwidth, height=6cm]{errors_number.pdf}
721-\caption{Left: square root of the sum of the residuals of the mass-size distribution, normalized by the total mass, as a function of the fractal dimension, for $\df$ in $[1.8,2.8]$, normalized by the minimum of value assumed by the sum. Right: as on the left, but now the residuals of the number-size distribution are normalized by the total aggregate number. }\label{mass-size residuals}
722-\end{figure}
836+On the right, the same calculations for the mass distribution are shown.
837+We notice that the distance function is much flatter and the determination of
838+the minimum is less obvious, especially for the $50$Km/h experiments.
839+This is somehow expected, since the number distribution does not
840+distinguish between small and large aggregates, thus weighting equally every aggregate size.
841+Therefore, even though the reconstructed inlet number distribution does not differ
842+significantly from the experimental distribution,
843+the use of the number or mass distribution to determine the fractal dimension is not equivalent.
844+We suggest the use of the mass distribution as the preferable metric for the determination of $d_{\rm f}$,
845+regardless of the specific experimental uncertainties,
846+since the effective density, and therefore $d_{\rm f}$, is intimately bound to
847+the mass rather than to the number concentration of the aggregates.
723848
724849 % Some experimental studies relate the aggregate fractal dimension to the engine load \citep{virtanen fractal, skillas fractal, park fractal}.
725850 % Unlike in this study, there the fractal dimension was found to be a decreasing function of the engine load.
726851 % However, the presence of the a diesel oxidation catalist (DOC), as in our experiments, can increase the sulphur content of the exhaust particles at high temperature \citep{olfert-fractal} thus giving rise to more compact aggregate structures, i.e. with a higher fractal dimension [{\it does it make sense when we do hot sampling???}].
727852
728853
729-\section{Discussion}
730-We can appreciate the importance of the modeling of the whole distribution, rather than a representative particle size, by investigating the
731- dynamics of the system according to Eq.~(\ref{explanation tau agglomeration}).
732-This amounts to approximating the initial number-size distribution with a monodisperse aerosol interacting with a constant collision kernel evaluated for a single aggregate size while neglecting deposition.
733- Such a treatment turns out to be a rather crude approximation, which underestimates the decrease of the particle concentration by several tens of percent. One can see that in Fig.~(\ref{comparison total number}) where $N(\tau)$ is calculated both from the output of the numerical simulations and according to Eq.~(\ref{explanation tau agglomeration}). This shows the importance of including all the aggregate sizes and the deposition mechanisms in the study of aerosol dynamics.
734-\begin{figure}
735-\includegraphics[width=0.5\columnwidth, height=6cm]{Total_particle_number_vs_time_comparison_monodisperse_approx.pdf}
736- \includegraphics[width=0.5\columnwidth, height=6cm]{Total_particle_number_vs_time_comparison_monodisperse_approx_50.pdf}
737-\caption{Left: for car speed $120$Km/h, $\tau$-evolution of the total aggregate concentration obtained from the numerics and the analytical expression in Eq.~(\ref{explanation tau agglomeration}). Right: as on left, but now the car speed is $50$Km/h. In both cases, Eq.~(\ref{explanation tau agglomeration}) underestimates the decrease in the number concentration. }\label{comparison total number}
738-\end{figure}
739-
740-\begin{figure}
741-\includegraphics[width=0.5\columnwidth, height=6cm]{2D-distribution_short_anaconda_50.pdf}
742- \includegraphics[width=0.5\columnwidth, height=6cm]{2D-distribution_short_anaconda.pdf}
743-\caption{bla bla }\label{comparison anaconda length}
744-\end{figure}
745-
746-The precise determination of $\df$ is affected by experimental uncertainties (the experimental reading of the exhaust air flow rates assumes air to be in normal conditions always inside the duct, whereas it is generally hotter and with non-constant density due to the decrease in temperature).
747-As a sensitivity study, we repeat the least-square minimization of the mass-size distribution errors to determine the fractal dimension $d_f$ while fictitiously varying the residence time $\tau_{\rm res}$ (in this test we exaggerate the experimental uncertainty) of the aerosol particles inside the transfer line. We allow $d_f$ to vary in $[1.8,2.8]$ in discrete steps of $0.05$ and, for each chosen $\tau_{\rm res}$, we determine the best fractal dimension in a least-square sense.
748-We notice that, while for the $120$Km/h case $d_f$ varies linearly with the residence time, it rather looks like a sum of step functions for car speed $50$Km/h. Such a behavior could be smoothed out by using a more refined grid on $d_f$, but this was not the purpose of the exercise, which was aimed at understanding the sensitivity of the fitted $d_f$ on the residence time.
749-This is indirectly an estimate of the effect of the fractal dimension on the system dynamics: particle-size distributions with different fractal dimensions would need a different residence time to undergo a similar evolution.
750-% For instance, in Fig~(\ref{df_fitting}) we show the values of the fitted $d_f$ as a function of the residence time.
751-% The uncertainty on the residence time is due to the non-constant value of the air density along the anaconda, which affects the determination of $U$.
752854
753855
754856
755-\begin{figure}
756-\includegraphics[width=0.5\columnwidth, height=6cm]{df_fitting_120.pdf}
757-\includegraphics[width=0.5\columnwidth, height=6cm]{df_fitting_50.pdf}
758-\caption{Test on the sensitivity of the fitted fractal dimension on the exhaust particles residence time in the transfer line. The experimental uncertanty on $\tau_{\rm res}$ has been exeggerated in this test. Left: estimate of $d_f$ as a function of the residence time for car speed $120$Km/h. The fitted $d_f$ depends linearly on the residence time. Right: as on the left, but now the car speed is $50$Km/h. The apparent jumps of $d_f$ are due to the finite resolution of the set of fractal dimensions which could be chosen by the least-square minimization. }\label{df_fitting}
759-\end{figure}
760-
761-
762-\begin{figure}
763-\includegraphics[width=0.5\columnwidth, height=6cm]{thermo_losses_50.pdf}
764-\includegraphics[width=0.5\columnwidth, height=6cm]{thermo_losses.pdf}
765-\caption{bla bla thermphoresis }\label{thermo_losses}
766-\end{figure}
767857
768858
769- We can conclude that the dynamics inside the duct is mainly dominated by agglomeration, which is sensitive to the residence time of aggregates inside the transfer line.
770-At high dilution ratios, the number-size distribution remains frozen inside the dilution tunnel and its measurement at the standard sampling point is thus significantly affected by the length of the transfer line. A difference in this length between two experimental set-ups could hinder the reproducibility of the measurements.
771-We also notice that the same conclusion does not necessarily hold true for a Euro-4 or Euro-5 vehicle.
772-This is due to the depence on $N_0$ of characteristic agglomeration time (see Eq.~(\ref{dimensionless times})) $\tau_{\rm ag}\sim 1/N_0$. Modern diesel-particulate-filter equipped cars have significantly lower number emissions, which translates into a time-scale for agglomeration which can easily be one or two orders of magnitude above the one estimated here. In that case, the role of agglomeration in the transfer line is much more limited.
773-Furthermore, for high exhaust temperature, thermophoresis is not negligible at all at should taken into account. This also raises important issues about the possibility of a later accidental resuspension of deposited material into the carrier flow.
859+
860+\section{Discussion}
861+\label{sec:discussion}
862+
863+The algorithm described in the previous sections provides
864+an efficient and simple procedure for the interpretation of
865+experimental results under different conditions. Accordingly, we
866+performed a number of sensitivity studies corresponding either
867+to simpler modelling approaches or to different experimental geometries.
868+
869+The importance of the modeling the complete number
870+distribution, rather than a representative particle size, is
871+appreciated by investigating
872+the dynamics of particles as given by Eq.~(\ref{explanation tau agglomeration}).
873+The initial number distribution is approximated by a monodisperse aerosol interacting
874+with a constant collision kernel evaluated for a single aggregate size,
875+neglecting deposition. Such a treatment is a rather crude approximation that
876+underestimates the decrease of the particle concentration by several tens of percent.
877+Figure~(\ref{comparison total number}) presents the total aggregate number
878+concentration $N(\tau)$ calculated with the numerical simulations presented
879+herein and according to Eq.~(\ref{explanation tau agglomeration}).
880+The significant difference of the two curves,
881+Eq.~(\ref{explanation tau agglomeration}) underestimates the decrease in the number concentration
882+for both experiments, shows the importance of including
883+all aggregate sizes and the dominant deposition mechanisms in a proper
884+description of aerosol dynamics in the transfer line.
885+
886+
887+
888+Furthermore, Fig.~(\ref{comparison total number}) shows the effect of the length of the transfer line (i.e. the residence time, since $U_m$ is not affected by the length of the duct) on the total aggregate concentration.
889+A simulated decrease from $9$ to $5$m amounts to almost halving $\tau_{\rm res}$ for both car speeds.
890+This is investigated as not all experimental geometries used to measure
891+number and mass concentrations have transfer lines of equal length.
892+ The role of agglomeration on particle concentration is also reflected by the outlet number distributions for different lengths of the
893+transfer line. Figure~\ref{comparison anaconda length} shows that
894+a significant variation of its length
895+would clearly lead to a different output number distribution.
896+This would severely affect solid-particle measurements at the
897+sampling point located at the outlet of the dilution tunnel, especially if the
898+distribution remains frozen in the dilution tunnel.
899+Finally. Fig.~\ref{comparison mu and sigma} shows the influence of the transfer line length on the geometric mean and standard deviation of the number distribution.
900+
901+We should remark, however, that the same conclusion might not necessarily hold for
902+more recent Euro-4 or Euro-5 vehicle. This is due to the dependence
903+of characteristic agglomeration time (see Eq.~(\ref{dimensionless
904+times})) $\tau_{\rm ag}\sim 1/N_0$ on the total number of
905+aggregate concentration at the inlet, $N_0$. Modern diesel-particulate-filter
906+equipped vehicles emit significantly lower number of particles: consequently, the
907+corresponding agglomeration time scale may easily be
908+one or two orders of magnitude above the estimate presented here. In that
909+case the role of agglomeration in the transfer line is much more
910+limited, as convection in the transfer line would be the dominant
911+aerosol process.
912+
913+Finally, we estimated the role of thermophoresis by varying
914+the wall temperature of the transfer line. For variations
915+of the order of $\sim 30-40\,^{\circ}\mathrm{C}$, we predict a
916+moderate (few percent) variation of the thermophoretic mass
917+deposition. In general, for high exhaust temperature, thermophoresis
918+is not negligible and it should taken into account. This also
919+raises important issues about the possibility of a later accidental
920+resuspension of deposited material into the carrier flow.
921+
922+
923+% The determination of $\df$ depends sensitively on the transfer-line
924+% residence time, itself dependent on experimental uncertainties, e.g., the experimental reading of
925+% the exhaust-air flow rates assumes air to be in normal conditions inside the duct,
926+% whereas it is hotter with non-constant, temperature dependent density.
927+% As a sensitivity study we, thus, repeated the least-square minimization of the mass distribution
928+% residuals by artificially varying the residence time $\tau_{\rm res}$ (admittedly an exaggeration of
929+% experimental uncertainties).
930+We finally estimate the effect of the fractal dimension on the aerosol dynamics in the duct.
931+We chose a fractal dimension
932+$d_{\rm f}$ in the interval $[1.8,2.8]$ and, after propagating the reconstructed inlet number distribution, we determine an artificial residence time which minimizes the least-square difference between the simulated and the fitted mass distributions at the outlet.
933+As expected, since a low fractal dimension enhances the collision rate between aggregates with the same volume of solids, the artificial residence time turns out to be a monotonic function of the fractal dimension.
934+We note that for the $120$Km/h experiments $\tau_{\rm res}$ varies linearly with the fractal dimension,
935+whereas for the $50$Km/h experiments it becomes a sum of step functions.
936+Such a behavior could be smoothed by using a more refined grid on $\tau_{\rm res}$, i.e. saving more often the mass distribution along the evolution,
937+but this was not the purpose of the exercise that aimed at understanding the sensitivity of the dynamics on the fractal dimension.
938+% Furthermore, this sensitivity study is indirectly an estimate of the effect of the fractal
939+% dimension on the system dynamics: number distributions with different fractal dimensions would
940+% need a different residence time to exhibit a similar evolution.
941+% For instance, in Fig~(\ref{df_fitting}) we show the values of the fitted $d_f$ as a function of the residence time.
942+% The uncertainty on the residence time is due to the non-constant value of the air density along the anaconda, which affects the determination of $U$.
943+
774944
775945 \section{Conclusion}
946+\label{sec:conclusion}
947+
776948 We developed a method for the study of one-dimensional aerosol dynamics in a turbulent flow.
777-For low Reynolds numbers and under steady-state conditions, we find ourself solving the GDE for aerosol particles in an ageing chamber with additional terms for particle losses.
778- The fractal dimension is a priori an unknown parameter but is determined self-consistently from the output of the dynamics.
779-Statistical considerations on the total aerosol mass per unit volume are used to reconstruct the initial total particle concentration.
949+For low Reynolds numbers and under steady-state conditions, we map the dynamics into a GDE for aerosol particles in an ageing chamber with additional terms for particle losses.
780950 Although the turbulent nature of the flow is neglected, we estimate the magnitude of the errors introduced by this approximation.
781-The fractal dimension of the agglomerates is treated as a constant in this study, but this is not an inherent limitation of the method developed here.
951+Statistical considerations on the total aerosol mass per unit volume are employed to reconstruct the initial total particle concentration.
952+The fractal dimension is a priori an unknown parameter but is determined self-consistently from the output of the dynamics
953+and we explain why we favour the mass over the number distribution for its determination. The sensitivity of the method on the particle total residence time in the transfer line is discussed.
954+We identify the time-scales of the relevant aerosol processes and show the influence of the length of the transfer line on the output number distribution.
955+Finally, we link the results to the case of the more up-to-date Euro-4 and Euro-5 vehicles.
782956
783957 \section*{Acknowledgements}
784-The research conducted here has been supported by the European Commission. The authors thank Dr. Panagiota Dilara
785-for [\emph{to be written}].
786-
787-% \begin{figure}
788-% \includegraphics[width=8cm, height=6cm]{2D-distribution.pdf}
789-% \caption{Evolution of the number-size distribution with chosen fractal dimension $2.25$ for car speed $120$Km/h. }\label{2D-120km}
790-% \end{figure}
958+The authors thank P. Dilara for helpful discussions and
959+G. de Santi for his support.
791960
792961 % The Appendices part is started with the command \appendix;
793962 % appendix sections are then done as normal sections
794- \appendix
795-
796- \section{One-dimensional correlations} \label{appendix corr}
797-In this appendix we list for completeness the 1D correlations employed to model losses and average temperature along the transfer line.
963+% \begin{appendix}
964+% \section{Depositional velocities}
965+% \label{appendix corr}
966+% \section{One-dimensional correlations} \label{appendix corr}
967+%In this appendix we list for completeness the 1D correlations employed to model losses and average temperature along the transfer line.
798968 % \subsection{Air properties}
799969 % Air kinematic viscosity, in units of m$^2$/s, is given as a function of temperature by the following empirical expression [{\it I took the following 2 expressions from some fittings on the web; I tested they are accurate wrt what you find in books, but how should I reference them???}]:
800970 % \begin{align}\label{air kinematic viscosity}
@@ -810,61 +980,169 @@
810980 % \eeq
811981 % where $p$ is the atmospheric pressure in units of bars.
812982 % The dynamic air viscosity, $\mu_{\rm air}$, is of course given by $\rho_{\rm air}\nu_{\rm air}$.
813-
814-
815-
816983 % \label{}
817-\subsection{Thermophoretic and diffusional losses coefficients}
818-The thermophoretic velocity can be expressed as in \cite{yannis 1d deposition}:
819-\beq
820-\vth= -\kt\;\f{\nu}{T}\f{T_w-T}{2R/\nusselt},
821-\eeq
822-where $R$ is the duct radius, $\nusselt$ is the Nusselt number and $\kt$ is Talbot coefficient.
823-Further correlations are needed to obtain the last two quantities.
824-In the case of a pipe, one can estimate the Nusselt number using Dittus-Boelter correlation \citep{book heat mass transfer}:
825-\beq
826-\nusselt=0.023\re^{0.8}\pr^{0.4},
827-\eeq
828-where $T_w$ is the transfer line wall temperature, $\re$ the Reynolds number and $\pr$ the Prandtl number, typically quite insensitive to temperature.
829- More sophisticated correlations, e.g. Glieninski formula \citep{book heat mass transfer}, lead fundamentally to the same estimate for $\nusselt$ when the Reynolds number ranges from $8000$ to $19000$, as in the case of the VELA-2 experiements.
830-Talbot coefficient can be estimated according to \citep{talbot}:
831-\beq
832-\kt=2C_sC_i\f{\ra+C_t\Kn_i}{(1+3C_m\Kn_i)(1+2\ra+2C_t\Kn_i)},
833-\eeq
834-where $C_s$, $C_m$ and $C_t$ are three constants equal to $1.17$, $1.14$ and $2.18$, respectively, $\ra=\kair/\kp$, $\kp$ being the carbonaceous particle thermal conductivity.
835-Then the coefficient for thermophoretic losses is given by \citep{yannis 1d deposition}:
836-\beq\label{expression vt}
837-\gth=\f{2\vdif}{R}.
838-\eeq
984+%\subsection{Thermophoretic and diffusional losses coefficients}
839985
840-The diffusion velocity is:
841-\beq\label{expression vd}
842-\vdif=\sh\f{\mathcal{D}_q}{2R},
843-\eeq
844-where $\sh$ is the Sherwood number which is calculated through the following correlation \citep{sherwood number}:
845-\beq\label{sherwood}
846-\sh=0.042\re f^{0.5}\sc^{1/3},
847-\eeq
848-where $\sc=\nu_{\rm air}/\mathcal{D}_q$ is the Schmidt number and $f$ is Fanning friction factor which in turn can be obtained from \citep{pope turbulence}:
849-\beq
850-f=0.0791\re^{-0.25},
851-\eeq
852-which is practically equivalent to other correlations widespread in literature like Churchill expression \citep{churchill}.
853-Similarly to Eq.~(\ref{expression vd}), the coefficient for diffusional losses is then given by $\gdif=2\vdif/R$.
854-\subsection{Temperature correlation}
855-The duct walls are kept at a constant temperature $T_w=70\,^{\circ}\mathrm{C}$ and $T$ (average temperature along the pipe cross-section) evolves according to \citep{yannis 1d deposition}:
856-\beq
857-T=T_w+(T_0-T_w)\exp\lro\f{-2\nusselt x}{R \re\pr}\rro,
858-\eeq
859-where $T_0$ is the average exhaust temperature at the pipe inlet.
860-As mentioned before, $x=U\tau$ so linking the temperature along the transfer line to the particle residence time is straightforward.
986+% For completeness we summarize the 1D correlations employed to model losses along the transfer line.
861987
988+% The thermophoretic velocity can be expressed as in Ref.~\cite{yannis 1d deposition}:
989+% \beq
990+% \vth= -\kt\;\f{\nu}{T}\f{T_w-T}{2{\rm R}/\nusselt},
991+% \eeq
992+% where ${\rm R}$ is the duct radius, $\nusselt$ is the Nusselt number and $\kt$ is Talbot coefficient.
993+% Further correlations are needed to obtain the last two quantities.
994+% In the case of a pipe, one can estimate the Nusselt number using Dittus-Boelter correlation \citep{book heat mass transfer}:
995+% \beq
996+% \nusselt=0.023\re^{0.8}\pr^{0.4},
997+% \eeq
998+% where $T_w$ is the transfer line wall temperature, $\re$ the Reynolds number and $\pr$ the Prandtl number, typically quite insensitive to temperature.
999+% More sophisticated correlations, e.g. Glieninski formula \citep{book heat mass transfer}, lead fundamentally to the same estimate for $\nusselt$ when the Reynolds number ranges from $8000$ to $19000$, as in the case of the VELA-2 experiements.
1000+% Talbot coefficient can be estimated according to \citep{talbot}:
1001+% \beq
1002+% \kt=2C_sC_i\f{\ra+C_t\Kn_i}{(1+3C_m\Kn_i)(1+2\ra+2C_t\Kn_i)},
1003+% \eeq
1004+% where $C_s$, $C_m$ and $C_t$ are three constants equal to $1.17$, $1.14$ and $2.18$, respectively, $\ra=\kair/\kp$, $\kp$ being the carbonaceous particle thermal conductivity.
1005+% Then the coefficient for thermophoretic losses is given by \citep{yannis 1d deposition}:
1006+% \beq\label{expression vt}
1007+% \gth=\f{2\vdif}{\rm R}.
1008+% \eeq
1009+
1010+% The diffusion velocity is:
1011+% \beq\label{expression vd}
1012+% \vdif=\sh\f{\mathcal{D}_q}{2 \rm R},
1013+% \eeq
1014+% where $\sh$ is the Sherwood number which is calculated through the following correlation \citep{sherwood number}:
1015+% \beq\label{sherwood}
1016+% \sh=0.042\re f^{0.5}\sc^{1/3},
1017+% \eeq
1018+% where $\sc=\nu_{\rm air}/\mathcal{D}_q$ is the Schmidt number and $f$ is Fanning friction factor which in turn can be obtained from \citep{pope turbulence}:
1019+% \beq
1020+% f=0.0791\re^{-0.25},
1021+% \eeq
1022+% which is practically equivalent to other correlations widespread in literature like Churchill expression \citep{churchill}.
1023+% Similarly to Eq.~(\ref{expression vd}), the coefficient for diffusional losses is then given by $\gdif=2\vdif/{\rm R}$.
1024+
1025+%\subsection{Temperature correlation}
1026+%The duct walls are kept at a constant temperature $T_w=70\,^{\circ}\mathrm{C}$ and $T$ (average temperature along the pipe cross-section) evolves according to \citep{yannis 1d deposition}:
1027+%\beq
1028+%T=T_w+(T_0-T_w)\exp\lro\f{-2\nusselt x}{R \re\pr}\rro,
1029+%\eeq
1030+%where $T_0$ is the average exhaust temperature at the pipe inlet.
1031+%As mentioned before, $x=U\tau$ so linking the temperature along the transfer line to the particle residence time is straightforward.
1032+% \end{appendix}
8621033 \begin{thebibliography}{00}
8631034
8641035 % \bibitem[Pippo. R, Rossi. G] {Seinfeld and Pandis, Aerosol Science and athp}
8651036
8661037 % \harvarditem{Name}{Year}{label}
8671038
1039+\hv{Andersson {\it et al.}}{2007}{andersson1}
1040+Andersson, J., Giechaskiel B., Munoz-Bueno, R., \& Dilara, P. (2007). Particle Measurement Programme (PMP): Light-duty Inter-laboratory Correlation Exercise (ILCE LD)-Final report (EUR 22775 EN) GRPE-54-08-Rev.1, \url{http://www.unece.org/trans/main/wp29/wp29wgs/wp29grpe/grpeinf54.html}
1041+Reg. 83 - Rev.3 - Emission of Pollutants According to Engine Fuel Requirements, \url{http://www.unece.org/trans/main/wp29/wp29regs81-100.html}
1042+
1043+\hv{Burtscher}{2004}{burtscher review} Burtscher, H. (2004). Physical characterization of particulate emissions from diesel engines: a review. {\it Journal of Aerosol Science, 36}, 896-932.
1044+
1045+\hv{Casati {\it et al.}}{2007}{casati}
1046+Casati, R, Scheer, V., Vogt R, \& Benter, T. (2007). Measurement of nucleation and soot mode particle emission from a diesel passenger car in real world and laboratory in situ dilution. {\it Atmospheric Environment, 41}, 2125-2135.
1047+
1048+\hv{Comsol Multiphysics}{2006}{comsol} Comsol Multiphysics version 3.2B. \url{http://www.comsol.com/}.
1049+
1050+% \hv{Drossinos \& Housiadas}{2006}{yannis book} Drossinos, Y., \& Housiadas, C. (2006). Aerosol Flows. {\it Multiphase Flow Handbook}. Boca Raton: Taylor \& Francis.
1051+
1052+\hv{Elzhov}{2005}{minpack} Elzhov, T.V. (2005). Minpack.lm: R interface for least squares optimization library.
1053+\url{http://cran.r-project.org/src/contrib/Descriptions/minpack.lm.html}.
1054+
1055+\harvarditem{Friedlander}{2000}{friedlander book} Friedlander S.K. (2000). {\it Smoke, Dust and Haze}. New York: Oxford University Press.
1056+
1057+% \hv{Gelbard \& Seinfeld}{1980}{gelbard-seinfeld} Gelbard, F., \& Seinfeld, J.H. (1980). Simulation of Multicomponent Aerosol Dynamics (1980). {\it Journal of Colloid and Interface Science, 78}, 485-501.
1058+
1059+\hv{Giechaskiel {\it et al.}}{2004}{makis2004}
1060+Giechaskiel, B., Ntziachristos. L., Samaras Z. (2004). Calibration and Modeling of Ejector Dilutors for Automotive Exhaust Sampling, {\it Measurement Science \& Technology, 15}, 2199-2206.
1061+\hv{Giechaskiel {\it et al.}}{2005}{makis2005}
1062+Giechaskiel, B., Ntziachristos, L., Samaras, Z., Casati, R., Scheer, V., \& Vogt, R. (2005). Formation Potential of Vehicle Exhaust Nucleation Mode Particles On-Road and in the Laboratory, {\it Atmospheric Environment, 39}, 3191-3198.
1063+\hv{Giechaskiel {\it at al.} }{2007a}{makis2007}
1064+Giechaskiel, B., Ntziachristos, L., Samaras, Z., Casati, R., Volker, \& S., Rainer, V. (2007a). Effect of speed and speed transition on the formation of nucleation mode particles from a light duty diesel vehicle. {\it SAE 2007-01-1110}.
1065+\hv{Giechaskiel {\it at al.} B}{2007b}{makis2007b}
1066+Giechaskiel, B., Munoz-Bueno, R., Rubino, L., Dilara, P., Urbano, M., \& Andersson, J. (2007b). Emissions of a Euro 4 light-duty diesel vehicle equipped with DPF. Close examination of particle size distributions during regeneration. {\it SAE 2007-01-1944}.
1067+
1068+\hv{Good}{2007}{good}
1069+Good P. (2007). {\it An Overview of European Emissions Legislation for LD vehicles}. Presentation at the International Conference on transport and Environment, Milan 19-21 March 2007, Italy, available at \url{http://transportenv07.jrc.it/proceedings_D.html}
1070+
1071+\hv{Hindmarsh}{1980}{lsoda1} Hindmarsh, A.C. (1980). LSODE and LSODI, two new initial value ordinary differential equation solvers. {\it ACM-SIGNUM newsletter, 15}, 10-11.
1072+
1073+\harvarditem{Housiadas \& Drossinos}{2005}{yannis 1d deposition} Housiadas C., \& Drossinos Y. (2005). Thermophoretic Deposition in Tube Flow. {\it Aerosol Science and Technology, 39}, 304-318.
1074+
1075+% \hv{Hull}{2005}{hull derivatives} Hull, J.C. (2006). {\it Options, Futures and Other Derivatives}, New York: Prentice-Hall.
1076+
1077+\hv{Incropera \& DeWitt}{1996}{book heat mass transfer} Incropera, F.P. \& DeWitt, D.P. (1996), {\it Fundamentals of Heat and Mass Transfer}. New York: John Wiley \& Sons.
1078+
1079+\hv{Jacobson \& Turco}{1994}{jacobson sectional} Jacobson M.Z., \& Turco, R.P. (1994). Modeling Coagulation among particles of different composition and size (1994). {\it Journal of Aerosol Science, 28}, 1327-1338.
1080+
1081+% \hv{Kittelson}{1997}{kittelson review} Kittelson, D.B. (1997). Engines and nanoparticles: a review. {\it Journal of Aerosol Science, 29}, 575-588.
1082+
1083+\hv{Kakac {\it et al.}}{1987}{churchill} Kakac, S., Shah, R. K., and Aung,W. (Eds.) (1987). {\it Handbook of Single-Phase
1084+Convective Heat Transfer.} John Wiley \& Sons: New York.
1085+
1086+\hv{Kostoglou \& Konstandopoulos }{2001}{evolving fractal 1} Kostoglou, M., \& Konstandopoulos, A.G. (2001). Evolution of aggregate size and fractal dimension during Brownian coagulation. {\it Journal of Aerosol Science, 32}, 1399-1420.
1087+
1088+% \hv{Kostoglou {\it et al.}}{2006}{evolving fractal 2} Kostoglou, M., Konstandopoulos, A.G., \& Friedlander, S.K. (2006). Bivariate population dynamics simulation of fractal aerosol aggregate coagulation and restructuring. {\it Jouarnal of Aerosol Science, 37}, 1102-1115.
1089+
1090+% \hv{Lee {\it et al.}}{2006}{lee_deposition} Lee, B.U., Byun, D.S., Bae, G., Lee, J.(2006). Thermophoretic Deposition of ultrafine particles in a turbulent pipe flow: Simulation of ultrafine particle behaviour in an automobile exhaust pipe. {\it Journal of Aerosol Science, 37}, 1788-1796.
1091+
1092+\hv{Maricq {\it et al.}}{1999}{maricq tailpipe}
1093+Maricq, M., Chase, R., Podsiadlik, D., Vogt, R. (1999). Vehicle Exhaust particle Size Distributions: A Comparison of Tailpipe and Dilution Tunnel Measurements. SAE 1999-01-1461
1094+
1095+
1096+
1097+\hv{Maricq \& Xu}{2004}{maricq experimental} Maricq, M.M., \& Xu, N. (2004). The effective density and fractal dimension of soot particles from premixed flames and motor vehicle exhaust. {\it Journal of Aerosol Science, 35}, 1251-1274.
1098+
1099+\hv{Maricq }{2006}{maricq coagulation} Maricq, M.M., (2006). Coagulation dynamics of fractal-like soot aggregates. {\it Journal of Aerosol Science, 38}, 141-156.
1100+
1101+\hv{Miller \& Garrick}{2004}{garrick section} Miller, S.E., \& Garrick, S.C. (2004). Nanoparticle Coagulation in a Planar Jet. {\it Aerosol Science and Technology, 38}, 79-89.
1102+
1103+
1104+
1105+\hv{Mountain \& Mulholland}{1986}{continuum formula} Mountain, R. D., Mulholland, G.W., \& Baum, H. (1986). Simulation of aerosol agglomeration in the free molecular and continuum flow regimes.
1106+ {\it Journal of Colloid and Interface Science, 114}, 67–81.
1107+
1108+\hv{Naumann}{2003}{naumann code} Naumann, K.H. (2003). COSIMA--a computer program simulating the dynamics of fractal aerosols. {\it Journal of Aerosol Science, 34}, 1371-1397.
1109+
1110+% \hv{Olfert {\it et al.}}{2006}{olfert-fractal} Olfert, J.S., Symonds, J.P.R., \& Collings, N. (2006). The effective density and fractal dimension of particles emitted from a light-duty diesel vehicle with a diesel oxidation catalyst. {\it Journal of Aerosol Science, 38}, 69-82.
1111+\hv{Ntziachristos {\it at al.}}{2004}{leon2}
1112+Ntziachristos, L., Giechaskiel, B., Pistikopoulos, P., Samaras, Z., Mathis, U., Mohr, M., Ristimäki, J., Keskinen, J., Mikkanen, P., Casati, R., Scheer, \& V., Vogt, R. (2004). Performance Evaluation of a Novel Sampling and Measurement System for Exhaust Particle Characterization. {\it SAE 2004-01-1439}.
1113+\hv{Ntziachristos {\it at al.}}{2005}{leon1}
1114+Ntziachristos, L., Giechaskiel, B., Pistikopoulos, P., \& Samaras, Z. (2005). Comparative Assessment of Two Different Sampling Systems for Particle Emission Type-Approval Measurements, {\it SAE Transactions 2005-01-0198}.
1115+
1116+ \hv{Park {\it et al.}} {2003}{park fractal} Park, K., Cao, F., Kittelson, D.B., \& McMurry, P.H. (2003). Relationship between Particle Mass and Mobility for Diesel Exhaust Particles. {\it Environmental Science and Technology, 37}, 577-583.
1117+% \hv{Petzold}{1983}{lsoda2} Petzold, L.R. (1983). Automatic Selection of Methods for Solving Stiff and Nonstiff Systems of Ordinary Differential Equations. {\it SIAM Journal on Scientific Computing, 4}, 136-148.
1118+
1119+\hv{Pope}{2000}{pope turbulence} Pope, S.B. (2000). {\it Turbulent Flows}, Cambridge: Cambridge University Press.
1120+
1121+\hv{Press {\it et al.}}{1988}{numerical recipes} Press, W.H., Flannery, B.P., Teukolsky, S.A., \& Vetterling W.T. (1988). {\it Numerical Recipes}, New York: Cambridge University Press.
1122+
1123+
1124+\hv{R\"{o}nkk\"{o} {\it et al.}}{2006}{ronkko}
1125+R\"{o}nkk\"{o}, T., Virtanen, A., Vaaraslahti, K., Keskinen, J., Pirjola, L., \& Lappi, M. (2006). Effect of dilution conditions and driving parameters on nucleation mode particles in diesel exhaust: Laboratory and on-road study. {\it Atmospheric Environment, 40}, 2893-2901.
1126+
1127+\hv{Saffman \& Turner}{1956}{saffman} Saffman, P.G., \& Turner, J.S. (1956). On the collision of drops in turbulent clouds. {\it J. Fluid Mech., 1}, 16-30.
1128+
1129+\harvarditem{Seinfeld \& Pandis}{1998}{pandis book} Seinfeld, J.H., \& Pandis, S.N. (1998). {\it Atmospheric chemistry and physics}. NewYork: Wiley.
1130+% \hv{Skillas G. {\it et al.}}{1998}{skillas fractal} Skillas, G., Kunzel, S., Burtscher, H., Baltensperger, U., \& Siegmann, K. (1998). High Fractal-Like Dimension of Diesel Exhaust Soot Agglomerates. {\it Journal of Aerosol Science, 29,} 411-419.
1131+\hv{Talbot {\it et al.}}{1980}{talbot} Talbot, L., Cheng, R.K., Schefer, R.W., \& Willis, D.R. (1980). Thermophoresis of particles in a heated boundary layer. {\it J. Fluid. Mechanics, 101}, 737-758.
1132+\hv{Van Gulijk C. {\it et al.}}{2004}{fractal experimental} Van Gulijk, C., Marijnissen, J.C.M., Makkee, M., Moulijn, J.A. \& Schmidt-Ott, A. (2004). Measuring diesel soot witha scanning mobility particle sizer
1133+and an electrical low-pressure impactor: performance. {\it Journal of Aerosol Science, 35}, 633-655.
1134+assessment witha model for fractal-like agglomerates
1135+\hv{Vemury \& Pratsinis}{1994}{pratsinis preserving adimensional} Vemury, S., \& Pratsinis, S.E. (1994). Self-preserving size distributions of agglomerates. {\it Journal of Aerosol Science, 26}, 175-185.
1136+\hv{Virtanen {\it et al.}}{2004}{virtanen fractal} Virtanen, A.K.K., Ristimaki, J.M., Vaaraslahti, K.M., \& Keskinen, J.(2004). Effect of Engine Load on Diesel Soot Particles. {\it Environmental Science and Technology, 38}, 2551-2556.
1137+\hv{Vogt \& Scheer}{2002}{vogt}
1138+Vogt, R., \& Scheer, V. (2002). {\it Particles in diesel vehicle exhaust: A comparison of laboratory and chasing experiments}. Proceedings of the 11th International Symposium Ttransport and Air Pollution, Graz 19-21.6.2002, pp 79-84, Univ.-Prof. Dr. R. Pischinger, Graz, ISBN 3-901351-59-0
1139+
1140+\hv{Voutsis {\it et al.}}{2005}{vouitsis} Vouitsis, E., Ntziachristos, L., \& Samaras, Z. (2005). Modelling of diesel exhaust aerosol during laboratory sampling.{\it Atmospheric Environment, 39}, 1335-1345.
1141+% \hv{Whitby}{1978}{whitby classical} Whitby, K.T. (1978). The physical characteristics of sulfur aerosols. {\it Journal of Aerosol Science, 12}, 135-159.
1142+
1143+\hv{Williams {\it et al.}}{1980}{sherwood number} Williams, M.M.R., Loyalka, S.K., 1991. {\it Aerosol Science.
1144+Theory and Practice}. Pergamon Press: Oxford.
1145+
8681146
8691147
8701148
@@ -883,93 +1161,10 @@
8831161
8841162
8851163
886-\hv{Burtscher}{2004}{burtscher review} Burtscher, H. (2004). Physical characterization of particulate emissions from diesel engines: a review. {\it Journal of Aerosol Science, 36}, 896-932.
887-
888-
889-
890-\hv{Drossinos \& Housiadas}{2006}{yannis book} Drossinos, Y., \& Housiadas, C. (2006). Aerosol Flows. {\it Multiphase Flow Handbook}. Boca Raton: Taylor \& Francis.
891-
892-
893-\hv{Elzhov}{2005}{minpack} Elzhov, T.V. (2005). Minpack.lm: R interface for least squares optimization library.
894-~http:~//~cran.r-project.org~/~src~/~contrib/~Descriptions~/~minpack.lm.html.
895-
896-\harvarditem{Friedlander}{2000}{friedlander book} Friedlander S.K. (2000). {\it Smoke, Dust and Haze}. New York: Oxford University Press.
897-
898-
899-\hv{Gelbard \& Seinfeld}{1980}{gelbard-seinfeld} Gelbard, F., \& Seinfeld, J.H. (1980). Simulation of Multicomponent Aerosol Dynamics (1980). {\it Journal of Colloid and Interface Science, 78}, 485-501.
900-
901-
902-\hv{Hindmarsh}{1980}{lsoda1} Hindmarsh, A.C. (1980). LSODE and LSODI, two new initial value ordinary differential equation solvers. {\it ACM-SIGNUM newsletter, 15}, 10-11.
903-
904-\harvarditem{Housiadas \& Drossinos}{2005}{yannis 1d deposition} Housiadas C., \& Drossinos Y. (2005). Thermophoretic Deposition in Tube Flow. {\it Aerosol Science and Technology, 39}, 304-318.
905-
906-\hv{Hull}{2005}{hull derivatives} Hull, J.C. (2006). {\it Options, Futures and Other Derivatives}, New York: Prentice-Hall.
907-
908-
909-\hv{Incropera \& DeWitt}{1996}{book heat mass transfer} Incropera, F.P. \& DeWitt, D.P. (1996), {\it Fundamentals of Heat and Mass Transfer}. New York: John Wiley \& Sons.
910-
911-\hv{Jacobson \& Turco}{1994}{jacobson sectional} Jacobson M.Z., \& Turco, R.P. (1994). Modeling Coagulation among particles of different composition and size (1994). {\it Journal of Aerosol Science, 28}, 1327-1338.
912-
913-\hv{Kittelson}{1997}{kittelson review} Kittelson, D.B. (1997). Engines and nanoparticles: a review. {\it Journal of Aerosol Science, 29}, 575-588.
914-
915-
916-
917-
9181164
9191165
9201166
921-\hv{Kakac {\it et al.}}{1987}{churchill} Kakac, S., Shah, R. K., and Aung,W. (Eds.) (1987). {\it Handbook of Single-Phase
922-Convective Heat Transfer.} John Wiley \& Sons: New York.
923-
924-
925-\hv{Kostoglou \& Konstandopoulos }{2001}{evolving fractal 1} Kostoglou, M., \& Konstandopoulos, A.G. (2001). Evolution of aggregate size and fractal dimension during Brownian coagulation. {\it Journal of Aerosol Science, 32}, 1399-1420.
926-
927-\hv{Kostoglou {\it et al.}}{2006}{evolving fractal 2} Kostoglou, M., Konstandopoulos, A.G., \& Friedlander, S.K. (2006). Bivariate population dynamics simulation of fractal aerosol aggregate coagulation and restructuring. {\it Jouarnal of Aerosol Science, 37}, 1102-1115.
928-
929-
930-\hv{Lee {\it et al.}}{2006}{lee_deposition} Lee, B.U., Byun, D.S., Bae, G., Lee, J.(2006). Thermophoretic Deposition of ultrafine particles in a turbulent pipe flow: Simulation of ultrafine particle behaviour in an automobile exhaust pipe. {\it Journal of Aerosol Science, 37}, 1788-1796.
931-
932-
933-
934-\hv{Maricq }{2006}{maricq coagulation} Maricq, M.M., (2006). Coagulation dynamics of fractal-like soot aggregates. {\it Journal of Aerosol Science, 38}, 141-156.
935-
936-\hv{Miller \& Garrick}{2004}{garrick section} Miller, S.E., \& Garrick, S.C. (2004). Nanoparticle Coagulation in a Planar Jet. {\it Aerosol Science and Technology, 38}, 79-89.
937-
938-
939-
940-\hv{Maricq M.M. \& Xu N.}{2004}{maricq experimental} Maricq, M.M., \& Xu, N. (2004). The effective density and fractal dimension of soot particles from premixed flames and motor vehicle exhaust. {\it Journal of Aerosol Science, 35}, 1251-1274.
941-
942-\hv{Mountain \& Mulholland}{1986}{continuum formula} Mountain, R. D., Mulholland, G.W., \& Baum, H. (1986). Simulation of aerosol agglomeration in the free molecular and continuum flow regimes.
943- {\it Journal of Colloid and Interface Science, 114}, 67–81.
944-
945-\hv{Naumann}{2003}{naumann code} Naumann, K.H. (2003). COSIMA--a computer program simulating the dynamics of fractal aerosols. {\it Journal of Aerosol Science, 34}, 1371-1397.
946-
947-\hv{Olfert {\it et al.}}{2006}{olfert-fractal} Olfert, J.S., Symonds, J.P.R., \& Collings, N. (2006). The effective density and fractal dimension of particles emitted from a light-duty diesel vehicle with a diesel oxidation catalyst. {\it Journal of Aerosol Science, 38}, 69-82.
948-
949-
950-\hv{Park {\it et al.}} {2003}{park fractal} Park, K., Cao, F., Kittelson, D.B., \& McMurry, P.H. (2003). Relationship between Particle Mass and Mobility for Diesel Exhaust Particles. {\it Environmental Science and Technology, 37}, 577-583.
951-\hv{Petzold}{1983}{lsoda2} Petzold, L.R. (1983). Automatic Selection of Methods for Solving Stiff and Nonstiff Systems of Ordinary Differential Equations. {\it SIAM Journal on Scientific Computing, 4}, 136-148.
952-
953-\hv{Pope}{2000}{pope turbulence} Pope, S.B. (2000). {\it Turbulent Flows}, Cambridge: Cambridge University Press.
954-
955-\hv{Press {\it et al.}}{1988}{numerical recipes} Press, W.H., Flannery, B.P., Teukolsky, S.A., \& Vetterling W.T. (1988). {\it Numerical Recipes}, New York: Cambridge University Press.
956-
957-\harvarditem{Seinfeld \& Pandis}{1998}{pandis book} Seinfeld, J.H., \& Pandis, S.N. (1998). {\it Atmospheric chemistry and physics}. NewYork: Wiley.
958-\hv{Skillas G. {\it et al.}}{1998}{skillas fractal} Skillas, G., Kunzel, S., Burtscher, H., Baltensperger, U., \& Siegmann, K. (1998). High Fractal-Like Dimension of Diesel Exhaust Soot Agglomerates. {\it Journal of Aerosol Science, 29,} 411-419.
959-\hv{Talbot {\it et al.}}{1980}{talbot} Talbot, L., Cheng, R.K., Schefer, R.W., \& Willis, D.R. (1980). Thermophoresis of particles in a heated boundary layer. {\it J. Fluid. Mechanics, 101}, 737-758.
960-\hv{Van Gulijk C. {\it et al.}}{2004}{fractal experimental} Van Gulijk, C., Marijnissen, J.C.M., Makkee, M., Moulijn, J.A. \& Schmidt-Ott, A. (2004). Measuring diesel soot witha scanning mobility particle sizer
961-and an electrical low-pressure impactor: performance. {\it Journal of Aerosol Science, 35}, 633-655.
962-assessment witha model for fractal-like agglomerates
963-\hv{Vemury \& Pratsinis}{1994}{pratsinis preserving adimensional} Vemury, S., \& Pratsinis, S.E. (1994). Self-preserving size distributions of agglomerates. {\it Journal of Aerosol Science, 26}, 175-185.
964-
965-\hv{Virtanen {\it et al.}}{2004}{virtanen fractal} Virtanen, A.K.K., Ristimaki, J.M., Vaaraslahti, K.M., \& Keskinen, J.(2004). Effect of Engine Load on Diesel Soot Particles. {\it Environmental Science and Technology, 38}, 2551-2556.
966-
967-\hv{Whitby}{1978}{whitby classical} Whitby, K.T. (1978). The physical characteristics of sulfur aerosols. {\it Journal of Aerosol Science, 12}, 135-159.
968-
969-\hv{Williams {\it et al.}}{1980}{sherwood number} Williams, M.M.R., Loyalka, S.K., 1991. {\it Aerosol Science.
970-Theory and Practice}. Pergamon Press: Oxford.
971-
972-\hv{Xiong \& Pratsinis}{1993}{xiong pratsinis} Xiong, Y., \& Pratsinis, S.E. (1993). Formation of agglomerate particles by coagulation and sintering-part I. A two-dimensional solution of the population balance equation. {\it Journal of Aerosol Science, 24}, 283-300.
1167+% \hv{Xiong \& Pratsinis}{1993}{xiong pratsinis} Xiong, Y., \& Pratsinis, S.E. (1993). Formation of agglomerate particles by coagulation and sintering-part I. A two-dimensional solution of the population balance equation. {\it Journal of Aerosol Science, 24}, 283-300.
9731168
9741169 % notes:
9751170 % \bibitem{label} \note
@@ -986,5 +1181,218 @@
9861181
9871182 \end{thebibliography}
9881183 % \end{linenumbers}
1184+
1185+\clearpage
1186+\begin{table}
1187+ \begin{center}
1188+\begin{tabular}{|c|r@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}l||r|}
1189+ \hline
1190+%\multicolumn{16}{|c|}
1191+% {\rule[-3mm]{0mm}{8mm} \textbf{Overview of the Numerical Results}} \\
1192+ Speed
1193+ & \multicolumn{2}{|c|}{Position}
1194+ & \multicolumn{2}{|c|}{$N$ [cm$^{-3}$]}
1195+ & \multicolumn{2}{|c|}{$N$ [cm$^{-3}$]}
1196+& \multicolumn{2}{|c|}{$M$ [g cm$^{-3}$] }
1197+ & \multicolumn{2}{|c|}{$M$ [g cm$^{-3}$]}
1198+ & \multicolumn{2}{|c|}{$\mu$ [nm]}
1199+ & \multicolumn{2}{|c|}{$\sigma$[nm]}
1200+ & $d_{\rm f}$ \\
1201+
1202+ & \multicolumn{2}{|c|}{}
1203+ & \multicolumn{2}{|c|}{fit}
1204+ & \multicolumn{2}{|c|}{rec}
1205+& \multicolumn{2}{|c|}{fit}
1206+ & \multicolumn{2}{|c|}{rec}
1207+ & \multicolumn{2}{|c|}{}
1208+ & \multicolumn{2}{|c|}{}
1209+ & \\ \hline \hline
1210+50 Km/h
1211+ & \multicolumn{2}{|c|}{\rm Inlet}
1212+ & \multicolumn{2}{|c|}{$7.6\cdot 10^7$}
1213+ &\multicolumn{2}{|c|}{$7.9\cdot 10^7$}
1214+ & \multicolumn{2}{|c|}{$1.41\cdot 10^{-8}$}
1215+ &\multicolumn{2}{|c|}{$1.47\cdot 10^{-8}$}
1216+ &\multicolumn{2}{|c|}{$62$}
1217+&\multicolumn{2}{|c|}{$1.75$}
1218+& $2$
1219+\\
1220+50 Km/h
1221+ & \multicolumn{2}{|c|}{\rm Outlet}
1222+ & \multicolumn{2}{|c|}{$4.5\cdot 10^{7}$}
1223+ &\multicolumn{2}{|c|}{}
1224+ & \multicolumn{2}{|c|}{$1.45\cdot 10^{-8}$}
1225+ &\multicolumn{2}{|c|}{}
1226+ &\multicolumn{2}{|c|}{$ 87.3$}
1227+&\multicolumn{2}{|c|}{$ 1.63$}
1228+& $2$
1229+\\
1230+
1231+120 Km/h
1232+ & \multicolumn{2}{|c|}{\rm Inlet}
1233+ & \multicolumn{2}{|c|}{$1.4\cdot 10^8$}
1234+ &\multicolumn{2}{|c|}{$1.38\cdot10^{8}$}
1235+ & \multicolumn{2}{|c|}{$4.1\cdot 10^{-8}$}
1236+ &\multicolumn{2}{|c|}{$4\cdot10^{-8}$}
1237+ &\multicolumn{2}{|c|}{$67.7$}
1238+&\multicolumn{2}{|c|}{$1.79$}
1239+& $2.2$
1240+\\
1241+120 Km/h
1242+ & \multicolumn{2}{|c|}{\rm Outlet}
1243+ & \multicolumn{2}{|c|}{$8.3\cdot 10^{7}$}
1244+ &\multicolumn{2}{|c|}{}
1245+ &\multicolumn{2}{|c|}{$3.7\cdot 10^{-8}$}
1246+&\multicolumn{2}{|c|}{}
1247+ &\multicolumn{2}{|c|}{$ 88.5$}
1248+&\multicolumn{2}{|c|}{$1.69$}
1249+& $2.2$
1250+\\
1251+
1252+ \hline
1253+% \vspace*{0.cm}
1254+\end{tabular}
1255+ \caption{Parameters of the experimental and calculated number distributions.}
1256+ \label{table statistics}
1257+\end{center}
1258+\end{table}
1259+
1260+% \clearpage
1261+\begin{table}
1262+\begin{center}
1263+\begin{tabular}{|c|r@{}lr@{}lr@{}l||r|}
1264+ \hline
1265+%\multicolumn{8}{|c|}
1266+% {\rule[-3mm]{0mm}{8mm} \textbf{Time scales for the two sets of experiments}} \\
1267+ Car Speed
1268+ & \multicolumn{2}{|c|}{Agglomeration}
1269+ & \multicolumn{2}{|c|}{Diffusion}
1270+ & \multicolumn{2}{|c|}{Thermophoresis}
1271+ & Residence Time \\ \hline \hline
1272+50 Km/h & \multicolumn{2}{|c|}{$\tau_{\rm ag}\simeq 4$s} & \multicolumn{2}{|c|}{$\tau_{\rm dif}\simeq 10^3$s} &\multicolumn{2}{|c|}{$\tau_{\rm th}\simeq 10^2$s} & $\tau_{\rm res}\simeq 4$s \\
1273+120 Km/h & \multicolumn{2}{|c|}{$\tau_{\rm ag}\simeq 2$s} & \multicolumn{2}{|c|}{$\tau_{\rm dif}\simeq 10^3$s} &\multicolumn{2}{|c|}{$\tau_{\rm th}\simeq 13$s} & $\tau_{\rm res}\simeq 2$s \\
1274+ \hline
1275+% \vspace*{0.cm}
1276+\end{tabular}
1277+ \caption{Characteristic time scales of the aerosol processes.}
1278+ \label{table time scales}
1279+\end{center}
1280+\end{table}
1281+
1282+
1283+\clearpage
1284+\begin{figure}
1285+\begin{center}
1286+\includegraphics[width=0.8\columnwidth, height=6cm]{JRC_CVS.jpg}
1287+\end{center}
1288+\caption{Schematic of the experimental set-up and sampling system at the VELA-2 laboratories.}\label{schematic experiment}
1289+\end{figure}
1290+
1291+\clearpage
1292+\begin{figure}
1293+ \begin{center}
1294+ \includegraphics[width=0.5\columnwidth, height=6cm]{mass-sizes.pdf}
1295+ \end{center}
1296+ \caption{Calculated cumulative mass distribution function at the inlet with $d_{\rm f}=2.3$, $d_0=50$nm and $\rho_0=1$g/cm$^3$.}
1297+\label{mass-size distr}
1298+ \end{figure}
1299+
1300+\clearpage
1301+\begin{figure}
1302+\includegraphics[width=0.5\columnwidth]{experimental-distr-with-error-bars_50.pdf}
1303+\includegraphics[width=0.5\columnwidth]{experimental-distr-with-error-bars.pdf}
1304+\caption{Left: Experimental number distributions, with statistical uncertainty, at inlet and outlet,
1305+vehicle speed $50$Km/h. Right: Vehicle speed $120$Km/h.}
1306+\label{experimental number-size}
1307+\end{figure}
1308+
1309+
1310+\clearpage
1311+\begin{figure}
1312+\includegraphics[width=\columnwidth]{diagram.pdf}
1313+\caption{Algorithm used to determine the fractal dimension of the aggregates.}
1314+\label{diagram for d_f}
1315+\end{figure}
1316+
1317+
1318+\clearpage
1319+\begin{figure}
1320+\includegraphics[width=0.5\columnwidth, height=6cm]{2D-distribution.pdf}
1321+ \includegraphics[width=0.5\columnwidth, height=6cm]{2D-mass_fitting.pdf}
1322+% \includegraphics[width=6.cm, height=6cm]{errors_120.pdf}
1323+% \includegraphics[width=10.cm, height=6cm]{evolution_3d_120.pdf}
1324+\includegraphics[width=0.5\columnwidth, height=6cm]{2D-distribution_50.pdf}
1325+\includegraphics[width=0.5\columnwidth, height=6cm]{2D-mass_fitting_50.pdf}
1326+\caption{Top row: Experiments at $120$Km/h and $d_{\rm f} = 2.2$.
1327+Top left: Reconstructed (solid line) and experimental (crosses) number distributions at the inlet,
1328+ experimental (circles) and numerical (dashed line) number distributions at the outlet.
1329+Top right: Mass distribution derived from the fitted number distributions at the inlet (dashed line)
1330+and at the outlet (crosses), simulations at the outlet (continuous line).
1331+Bottom row: Experiments at $50$Km/h and $\df=2$.}
1332+\label{mass-size fig}
1333+\end{figure}
1334+
1335+\clearpage
1336+\begin{figure}
1337+\includegraphics[width=0.5\columnwidth, height=6cm]{errors_mass.pdf}
1338+\includegraphics[width=0.5\columnwidth, height=6cm]{errors_number.pdf}
1339+\caption{Left: Square root of the sum of the mass distribution residuals (arbitrary units)
1340+as a function of the fractal dimension.
1341+Right: Square root of the sum of the number distribution residuals.}\label{mass-size residuals}
1342+\end{figure}
1343+
1344+
1345+\clearpage
1346+\begin{figure}
1347+\includegraphics[width=0.5\columnwidth, height=6cm]{Total_particle_number_vs_time_comparison_monodisperse_approx.pdf}
1348+\includegraphics[width=0.5\columnwidth, height=6cm]{Total_particle_number_vs_time_comparison_monodisperse_approx_50.pdf}
1349+\caption{Left: calculated $\tau$-evolution of the total aggregate concentration, numerical simulations (circles),
1350+ the analytical expression Eq.~(\ref{explanation tau agglomeration}) (crosses), experiments at $120$Km/h.
1351+ The arrow indicates the total particle concentration corresponding to a shorter transfer line
1352+ (i.e. a shorter residence time), namely a $5$m-long line.
1353+Right: Experiments at $50$Km/h.}
1354+\label{comparison total number}
1355+\end{figure}
1356+
1357+
1358+\clearpage
1359+\begin{figure}
1360+\includegraphics[width=0.5\columnwidth, height=6cm]{2D-distribution_short_anaconda_50.pdf}
1361+\includegraphics[width=0.5\columnwidth, height=6cm]{2D-distribution_short_anaconda.pdf}
1362+\caption{Left: Reconstructed inlet number distribution (solid) and calculated outlet number
1363+distributions for a transfer line length of $5$m (crosses) and $9$m (dashed), vehicle speed $50$Km/h.
1364+Right: Vehicle speed $120$Km/h.}
1365+\label{comparison anaconda length}
1366+\end{figure}
1367+
1368+\clearpage
1369+\begin{figure}
1370+\includegraphics[width=0.5\columnwidth, height=6cm]{mu_geo.pdf}
1371+\includegraphics[width=0.5\columnwidth, height=6cm]{sig_geo.pdf}
1372+\caption{Left: Evolution of the geometric mean of the number distribution
1373+ as a function of the residence time and indication
1374+of the outlet value for a $5$m-long transfer line for car speed $120$ (circles) and $50$Km/h (crosses)..
1375+Right: The evolution of the geometric standard deviation is plotted.}
1376+\label{comparison mu and sigma}
1377+\end{figure}
1378+
1379+\clearpage
1380+\begin{figure}
1381+\includegraphics[width=0.5\columnwidth, height=6cm]{df_fitting_120.pdf}
1382+\includegraphics[width=0.5\columnwidth, height=6cm]{df_fitting_50.pdf}
1383+\caption{Determination of the artificial residence time $\tau_{\rm res}$ minimizing the least-square difference between the numerical and fitted mass distributions at the outlet for a set of choices of the fractal dimension.
1384+Left: Experiments at $120$Km/h. Right: Experiments at $50$Km/h.}
1385+\label{df_fitting}
1386+\end{figure}
1387+
1388+
1389+
1390+
1391+% \begin{figure}
1392+% \includegraphics[width=8cm, height=6cm]{2D-distribution.pdf}
1393+% \caption{Evolution of the number-size distribution with chosen fractal dimension $2.25$ for car speed $120$Km/h. }\label{2D-120km}
1394+% \end{figure}
1395+
1396+
1397+
9891398 \end{document}
990-