Relationship between Coating-Induced Soot Aggregate Restructuring

Publication Date (Web): June 29, 2017 ... from all three sources exhibit the same linear dependence on the number of primary particles per aggregate...
0 downloads 0 Views 1MB Size
Subscriber access provided by Olson Library | Northern Michigan University

Article

Relationship between Coating-Induced Soot Aggregate Restructuring and Primary Particle Number Kaiser Leung, Elijah G Schnitzler, Ramin Dastanpour, Steven Rogak, Wolfgang Jaeger, and Jason S Olfert Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b01140 • Publication Date (Web): 29 Jun 2017 Downloaded from http://pubs.acs.org on July 1, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 31

Environmental Science & Technology

1

Relationship between Coating-Induced Soot

2

Aggregate Restructuring and Primary Particle

3

Number

4

Kaiser K. Leung,† Elijah G. Schnitzler,‡,¶ Ramin Dastanpour,§ Steven N. Rogak, § Wolfgang

5

Jäger,‡ and Jason S. Olfert*,†

6



7

2G8

8



Department of Chemistry, University of Alberta, Edmonton, AB, Canada T6G 2G2

9

§

Department of Mechanical Engineering, University of British Columbia, Vancouver, BC,

10

Department of Mechanical Engineering, University of Alberta, Edmonton, AB, Canada T6G

Canada V6T 1Z4

11 12

AUTHOR INFORMATION

13

Corresponding Author

14

*Phone: (780) 492-2341; fax: (780) 492-2200; e-mail: [email protected].

15

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 31

16

ABSTRACT

17

The restructuring of mono-disperse soot aggregates due to coatings of secondary organic aerosol

18

(SOA) was investigated in a series of photo-oxidation chamber experiments. Soot aggregates

19

were generated by one of three sources (an ethylene premixed burner, a methane inverted

20

diffusion burner, or a diesel generator), treated by denuding, size-selected by a differential

21

mobility analyzer, and injected into a smog chamber, where they were exposed to the photo-

22

oxidation products of p-xylene, which partitioned to form SOA coatings. The evolution of

23

aggregates from their initial to final morphologies was investigated in situ by mobility and mass

24

measurements and ex situ by transmission electron microscopy. At a given initial aggregate

25

mobility diameter, diesel aggregates are less dense and composed of smaller primary particles

26

than those generated by the two burners, and they restructure to a smaller final mobility

27

diameter. Remarkably, the final degrees of restructuring of aggregates from all three sources

28

exhibit the same linear dependence on the number of primary particles per aggregate. The

29

observed linear relationship, valid for the atmospherically-relevant SOA coating investigated

30

here, could allow modelers to predict the evolution of aggregate morphology based on a single

31

property of the aggregates.

32

ACS Paragon Plus Environment

2

Page 3 of 31

33

Environmental Science & Technology

TOC GRAPHICS

34

ACS Paragon Plus Environment

3

Environmental Science & Technology

35

Page 4 of 31

Introduction

36

Soot aggregates in the atmosphere have a significant warming effect on Earth’s climate,

37

second only to that of carbon dioxide,1 due to strong absorption of all visible wavelengths of

38

light by their constituent primary particles of elemental carbon.2 Soot aggregates form at their

39

sources, which include fossil fuel combustion3,4 and biomass burning,5,6 by collisions between

40

primary particles, giving the soot a branched or fractal-like structure.7 Depending on the source

41

and the combustion conditions, the primary particles may vary from 5 to 60 nm in diameter,8,9

42

and aggregates may comprise tens to hundreds of primary particles.7

43

The structural and optical properties of soot aggregates can be affected by internally-

44

mixed non-absorbing species. For example, restructuring of soot aggregates from initially

45

branched structures to comparably compact structures can be induced by coatings of many

46

species, including water,10–12 heptane,13 ethanol,13 dioctyl sebecate,14 oleic acid,14,15 glutaric

47

acid,16 and sulfuric acid.10,17 The evolution of the optical properties of soot aggregates as a result

48

of restructuring and/or lensing has recently been investigated, both experimentally18 and

49

theoretically.19 For a bare soot aggregate of a given mass, a compact structure results in a greater

50

extinction cross section than a branched structure; however, this is due to increased scattering,

51

not absorption, so compaction may not necessarily lead to an increased warming effect on

52

climate.18 On the other hand, the presence of a coating on the aggregate may result in increased

53

absorption even for a compact structure due to lensing, and this enhancement may lead to an

54

increased warming effect.20 Thorough knowledge of the morphological evolution of soot induced

55

by coatings is necessary to further our understanding of the complex climate effects of soot.

56

In the atmosphere, soot aggregates are exposed to many gas-phase species that may

57

condense and contribute to coatings. Accordingly, it is important to investigate the

ACS Paragon Plus Environment

4

Page 5 of 31

Environmental Science & Technology

58

morphological evolution of aggregates upon coating with mixtures, in addition to neat species.

59

Secondary organic aerosol (SOA), which forms from the photo-oxidation of biogenic and

60

anthropogenic volatile organic compounds,21,22 is a mixture of numerous oxygenated organic

61

species. Soot aggregate restructuring has been induced by SOA coatings generated from

62

isoprene,23 α-pinene,24,25 benzene,26 toluene,26,27 ethylbenzene,26 and m- and p-xylene.26,28 Very

63

recently, the morphological and optical properties of burner-generated soot aggregates exposed

64

to ambient air in Beijing and Houston were investigated.29 Since fossil fuel combustion is a

65

significant source of ambient soot aggregates, it is also important to investigate restructuring of

66

engine-generated aggregates, in addition to aggregates generated by typical laboratory burners.

67

The evolution of soot aggregates produced by a diesel engine upon photo-oxidation of gas

68

emissions has been investigated in the past.30 More specifically, previous studies have

69

demonstrated how the properties of the coatings affect restructuring; for example, it was found

70

that the extent of restructuring increases with increasing coating mass14,26 and/or surface

71

tension.31 However, the question of how the properties of the aggregates themselves, including

72

primary particle size and number, dictate restructuring has not been the focus of any previous

73

studies.

74

Here, we investigate the effects of the soot aggregate properties by systematically

75

comparing the coating-induced restructuring of engine- and burner-generated aggregates. We

76

describe a series of smog chamber experiments in which SOA coatings derived from p-xylene, a

77

representative anthropogenic precursor,32 were used to restructure mono-disperse soot aggregates

78

from three sources: a McKenna premixed burner, an inverted diffusion burner, and a commercial

79

diesel generator. Adopting a technique introduced by Pagels et al.,10 we characterize the

80

morphological evolution of both the coated particles and their soot cores in terms of effective

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 31

81

densities, dynamic shape factors, and mass-mobility relationships, derived from masses and

82

mobilities measured using a centrifugal particle mass analyzer (CPMA) and differential mobility

83

analyzer (DMA), respectively. The initial and final morphologies of the aggregates were also

84

characterized using transmission electron microscopy (TEM). The primary particle sizes and

85

numbers were determined using an indirect method, based on in situ CPMA and the DMA

86

measurements, and a direct method, based on ex situ transmission electron microscopy (TEM)

87

samples. The dependence of restructuring on aggregate properties is explored in detail, and the

88

atmospheric implications are discussed. Our study provides key insights for modeling the

89

evolution of the morphological and, in turn, optical properties of soot aggregates.

90 91

Experimental Section

92

Soot generation and treatment. Soot was produced using a McKenna premixed burner

93

(Holthuis & Associates), a custom-built inverted diffusion burner, and a commercial diesel

94

generator (Yanmar, YDG3700). The respective experimental setups for the treatment and

95

injection of soot are shown in Figure 1. The treatment setup for the McKenna burner experiments

96

is shown in Figure 1a, and it has been described in detail elsewhere.26 Briefly, ethylene was used

97

as fuel, and the equivalence ratio was set to two, achieved with ethylene and air flow rates of

98

1.1 and 8.0 L min-1, respectively. Soot was sampled 27 cm above the flame into an ejector dilutor

99

(Air-Vac, AVR038H), set to a dilution ratio of three, using 30 psi of nitrogen. The soot particles

100

were then treated by passing them through a 25 cm diffusion drier consisting of tubular mesh

101

surrounded by anhydrous calcium sulfate, a thermo-denuder set to 300 °C, and a counter-flow

102

parallel-plate membrane denuder described elsewhere.26 A representative size distribution is

103

shown in Figure S1.

ACS Paragon Plus Environment

6

Page 7 of 31

Environmental Science & Technology

104

The treatment setup for the inverted burner experiments is shown in Figure 1b. The

105

inverted burner is a successor to the one used by Ghazi et al.33 Methane was used as fuel, and the

106

global equivalence ratio was set to 0.76, achieved with methane and air flow rates of 1.3 and

107

16.0 L min-1, respectively. In addition to the combustion air, clean filtered air was used for

108

dilution, at a flow rate of approximately 40 L min-1. The soot was sampled downstream of the

109

dilution phase and directed to the treatment train, which is similar to that used for the McKenna

110

premixed burner, with one addition. In control experiments performed with the above treatment,

111

OH exposure (without injection of p-xylene) induced restructuring of inverted burner aggregates

112

but had no effect on McKenna burner aggregates. This discrepancy indicates that an appreciable

113

fraction of trace gas emissions of the inverted burner penetrated the treatment train, and were

114

oxidized in the chamber to form SOA. Since our objective is to compare restructuring of soot

115

from different sources, it is critical that the SOA coatings are derived solely from p-xylene, so

116

the background SOA formation is unsatisfactory. Consequently, we introduced a 91-cm tube

117

consisting of tubular mesh surrounded by activated carbon between the thermo- and membrane

118

denuders. In subsequent control experiments, OH exposure alone did not cause restructuring of

119

inverted burner aggregates, so this treatment is sufficient.

120

The treatment setup for the diesel generator experiments is illustrated in Figure 1c. A

121

series of resistors was used to run the engine at set loads. As shown in Figure S2, an

122

approximately 100% load (3.5 kW) resulted in particles with the lowest volatile mass fraction

123

(13%). Flexible metal tubing was fastened to the exhaust of the engine, and a sample port was

124

drilled 23 cm downstream. The soot particles were then passed through the same treatment train

125

as was used for the inverted burner, including the activated carbon denuder.

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 31

126

After treatment, the soot aggregates were neutralized using an X-ray source (TSI, 3087)

127

and classified by mobility diameter using a DMA (TSI, 3081), which was operated with sample

128

and sheath flow rates of 1.0 and 10.0 L min-1, respectively. The size-selected, mono-disperse

129

aggregates were then injected into the smog chamber, which has been described in detail in the

130

past.34,35 The 1.8 m3 cubic chamber is constructed of perfluoroalkoxy film (Ingeniven) and

131

equipped with three sets of eight 32 W black-lights. Clean air is provided by a pure air generator

132

(Aadco, 737), and the contents of the chamber are circulated by a mixing fan. Ultraviolet

133

differential optical absorption spectroscopy was used to measure concentrations of p-xylene by

134

passing broadband emission from a deuterium lamp (Ocean Optics, D-2000-S) through the

135

chamber to a spectrometer (Ocean Optics, HR-2000+).

136

Precursor photo-oxidation and soot restructuring. When soot injection was

137

completed, the soot in the chamber was directed, alternately with and without thermo-denuding,

138

into parallel systems to measure the mean mobility diameter and mass of the particles. A DMA

139

and condensation particle counter (CPC; TSI, 3776) were used to measure the mean mobility

140

diameter, by stepping the DMA through a range of diameters and measuring the particle

141

concentration at each step with by the CPC. The mean particle mass was measured by stepping a

142

CPMA, which classifies particles by their mass-to-charge ratio, through a range of particle

143

masses and measuring the particle concentration with another CPC (TSI, 3771). When the DMA

144

voltage was set to select, for example, singly-charged 150 nm aggregates from the diesel

145

generator, McKenna burner, and inverted burner, about 5, 22, and 38%, respectively, of the

146

classified particles were estimated to be larger doubly-charged aggregates. The transmission of

147

multiply-charged particles reduces the ability to resolve the mass of singly charged particles in

148

the CPMA; as shown by Radney et al., the mass can be better resolved by using a second

ACS Paragon Plus Environment

8

Page 9 of 31

Environmental Science & Technology

149

neutralizer (TSI, 3077) upstream of the CPMA.36 Soot was monitored continuously for

150

approximately one hour after injection to confirm that there was no background change in

151

mobility diameter or mass. Then p-xylene (Fisher, 99.9%) was injected into the chamber to give

152

a concentration of approximately 1 ppm. This concentration, which is too high to be

153

atmospherically representative, is required to probe a wide range of coating masses; also, it has

154

been shown that different initial precursor concentrations lead to the same extent of restructuring

155

at a given coating mass,27 so the effects reported here, particularly at low coating masses, are

156

likely relevant to atmospheric conditions. Precursor concentrations higher than 1 ppm have been

157

used, for example, in studies of the optical properties of SOA.37 Hydrogen peroxide (Sigma, 30%

158

w/w in water) was then injected into the chamber to yield hydroxyl radicals by photolysis with

159

UV radiation,38 which was applied continually throughout the experiment.

160

Restructuring, or the collapse of the soot particles, is described in terms of mass-mobility

161

exponent, effective density, and shape factor. The mass-mobility exponent,  , describes how

162

the particle mass,  , scales with mobility diameter,  , according to 

 = 

163 164

(1)

where  is the mass-mobility prefactor.4 For an irregular particle, the mass-mobility exponent is

less than three. The effective density, , of a particle is the mass of the particle divided by its

165

mobility-equivalent volume, (6 )/(  ). In the following discussion, the effective densities

166

 of coated and coated-denuded particles are denoted as and  , respectively. The dynamic

167

shape factor, , is calculated as [  ( )]/[  ( )], where  is the Cunningham slip

168

correction factor for a certain diameter, and ve is the volume equivalent diameter, which is the

169

hypothetical diameter of a spherical particle with the same volume. Assuming there are no

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 31

170

internal voids (i.e. empty regions within the particle that are isolated from the bulk gas phase39),

171

ve is  

6  =    

)

 ! −  #  + '(

$%&

(2)

172

 where  and $%& are the material densities of soot and SOA, respectively, and  and 

173

are the masses of the initial soot aggregates (before the injection of p-xylene) and the coated

174

particles (after photo-oxidation), respectively. Again, the superscript denotes that neither type of

175

 particles was thermo-denuded after sampling from the chamber. The difference  − 

176

gives the mass of the SOA coating. The material density of soot was taken as 1.8 g cm-3,40 and

177

the density of the p-xylene-derived SOA was taken as 1.46 g cm-3.41 For spherical particles, 

178 179

will be equivalent to  , and the shape factor will be unity. The shape factors of coated and coated-denuded particles are denoted as   and   respectively.

180

Restructuring was also probed using TEM. To this end, mobility-classified particles were

181

collected on TEM grids using a thermophoretic particle sampler (TPS). Images were produced

182

by a Hitachi H7600 transmission electron microscope operated at 80 kV. Images were analyzed

183

by an open-source automated TEM image processing program, which uses the pair correlation

184

method (PCM).42,43 This program enables measuring of the average primary particle diameter

185

and various morphology parameters of aggregates.

186

Results and Discussion

187

Initial aggregate morphology. The mass-mobility relationships for the initial aggregates

188

are depicted in Figure 2a. The mass-mobility exponents for aggregates from the three sources are

189

all significantly less than three, indicating that their effective densities decrease with increasing

190

mobility diameter. This trend is well-known and has been previously observed for soot generated

ACS Paragon Plus Environment

10

Page 11 of 31

Environmental Science & Technology

191

by laboratory burners26,33 and diesel engines.3,4 The mass-mobility relationships of the diesel,

192

inverted burner, and McKenna burner aggregates determined here are broadly consistent with

193

earlier measurements (see Figure S3).3,4,33,44 At mobility diameters of 100 and 150 nm,

194

aggregates from both burners are considerably denser than those from the diesel generator.

195

At a given mobility diameter, the mass of soot aggregates can be related to the diameters

196

of the primary particles constituting the aggregates generated by each of the three sources.

197

Primary particle diameters were investigated both indirectly, using mass and mobility

198

measurements, and directly, using TEM microscopy. If the aggregate mobility diameter is taken

199

to be equivalent to the projected-area equivalent diameter (determined by TEM),45,46 then the

200

surface-area equivalent primary particle diameter, * , can be estimated from the following

201

relationship:47  *

202

)

,- . +*

( ),- ' = 6

(3)

where +* and 0 are parameters dictated by the formation of and collisions between primary

203

particles. The values of +* and 0 ‒ 1.13 and 1.1, respectively ‒ were optimized based on mass-

204

mobility and TEM data for two engine sources (diesel and gasoline direct injection) over a wide

205

range of operating conditions, and they are expected to be an estimate for other soot sources.48

206

The estimated * values for each experiment are listed in Table S1, and they are compared to

207

the ex situ, but direct, results of the TEM image analysis obtained by the PCM program.42 We

208

note that small relative differences in primary particle diameter between the two methods lead to

209

more significant differences in primary particle number. In general, the primary particles

210

produced by the diesel generator were estimated to have smaller diameters (15.6 nm for 100 nm

211

aggregates) than those produced by the inverted and McKenna burners (27.5 and 28.5 nm for

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 31

212

100 nm aggregates, respectively). Similarly, direct TEM observations of 100 nm aggregates

213

indicate that the diesel generator produces the smallest primary particles: for the diesel generator,

214

the mean * is 18.0 nm (with a 95% confidence interval of 13.7–22.3 nm); for the McKenna

215

and inverted burners, the mean * values are 24.6 and 24.5 nm, respectively (with respective

216

95% confidence intervals of 21.8–27.4 and 21.3–27.7 nm). The average number of primary

217

particles, 1, is the ratio of aggregate mass to primary particle mass, the latter of which is derived

218

from * and the material density of soot, and listed in Table S1. Based on TEM observations,

219

aggregates initially having a 100 nm mobility diameter produced by the diesel generator

220

comprise about 46 primary particles, whereas those produced by the McKenna and inverted

221

burners comprise about 29 and 30 primary particles, respectively. This difference explains why

222

aggregates from the diesel generator are the least dense; the greater number of primary particles

223

accommodates a more branched structure.

224

Evolution of aggregate morphology. To generate atmospherically-relevant coatings,

225

we exposed p-xylene, a representative anthropogenic aromatic hydrocarbon, to hydroxyl radical,

226

which initiated photo-oxidation and led to the formation of semi-volatile oxygenated organic

227

compounds. Since the concentration of hydroxyl radical was constant, the reaction of p-xylene

228

was governed by pseudo-first-order kinetics, and the product concentrations increased

229

exponentially. Once the products saturated the gas phase, they began partitioning onto the soot

230

aggregates as SOA, leading to exponential growth in particle mass. To facilitate comparisons

231

between aggregates from different sources and of different initial mobility diameters, we divide

232

mobility diameters and masses by their initial values to give diameter and mass growth factors,

233

Gfd and Gfm, respectively. Typical trends in diameter and mass growth factors are shown in

234

Figure S4 for initially 100 nm aggregates from both the inverted burner and the diesel generator.

ACS Paragon Plus Environment

12

Page 13 of 31

Environmental Science & Technology

235

As shown in Figure S4b, the mass of the coated-denuded aggregates was constant throughout the

236

experiment, indicating that the thermo-denuder temperature (300 °C) was high enough to

237

evaporate all the SOA coating. (The thermo-denuder likely also affected the physical properties

238

of the SOA coatings before evaporation; in a future study, we will investigate the role of SOA

239

coating viscosity, in particular, under a variety of relative humidity conditions.) As shown in

240

Figure S4a, the diameter of the coated-denuded aggregates from the generator decreases more

241

than that of the aggregates from the burner, a first indication that the morphological evolution of

242

soot aggregates from different sources varies.

243

The evolution of aggregate morphology occurs in three stages that can be identified, for

244

example, in Figure 3, which depicts effective densities and shape factors versus mass growth

245

factor for initially 250 nm aggregates from the two burners. In the first stage, the particles

246

become denser as coating mass increases, but the core soot aggregates retain their initial

247

morphologies. This induction period is not captured by an earlier exponential model derived

248

from experiments with a lower coating mass resolution.14 In the second stage, commencing at a

249

mass growth factor of about 1.5, the soot aggregates begin to restructure under the influence of

250

the SOA coating. The final stage commences at a mass growth factor of about five, and the

251

effective density of the coated-denuded soot plateaus at approximately 0.6 g cm-3, which is

252

considerably lower than the material density of soot. The shape factors of the coated particles

253

continue to decrease with increasing coating mass before reaching unity at a mass growth factor

254

of approximately 10, indicating that a thickly-coated spherical particle is then present. Despite

255

the initial difference in the shape factors of the coated-denuded aggregates from the two burners,

256

they both plateau at approximately 1.8, which implies that the restructured soot is not spherical,

257

consistent with the final effective density.

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 31

258

In one experiment, significantly greater coating masses were applied to initially 150 nm

259

aggregates from the inverted burner. As shown in Figure S5, a mass growth factor of over 30 is

260

reached. As discussed above, at a mass growth factor of approximately 10, the shape factor of

261

the coated particles plateau, and the effective density ceases to increase. As the mass growth

262

factor continues to increase, the shape factor is of course constant, but the effective density

263

exhibits a slight decrease, because the mass fraction of the soot, which is denser than the SOA

264

coating, decreases significantly. The asymptote corresponds to the material density of the SOA

265

coating.

266

To compare the evolution of soot aggregates from all three sources, we must consider

267

smaller aggregates, because too few 250 nm aggregates are produced by the diesel generator to

268

allow classification and injection into the smog chamber. Accordingly, trends in effective

269

densities and shape factors versus mass growth factor are shown in Figure 4 for initially 100 nm

270

aggregates from the three sources. As noted above for the 250 nm aggregates, restructuring

271

began at a mass growth factor of about 1.5. For aggregates from the two burners, restructuring

272

continued until the mass growth factor reached about three, at which point the effective densities

273

and shape factors of the coated-denuded particles plateaued. In contrast, the aggregates from the

274

diesel generator continued to restructure until the mass growth factor reached a value of

275

approximately six. The initial effective densities of aggregates from the McKenna and inverted

276

burners are similar (~ 0.6 g cm-3), whereas that of the aggregates from the diesel generator is

277

lower (~ 0.4 g cm-3). Despite the difference in initial effective densities, the coated diesel

278

aggregate effective densities increase less rapidly to approximately 1.4 g cm-3 than those of the

279

other aggregates. When restructuring is complete, the coated-denuded aggregates from the

280

McKenna and inverted burners reach effective densities of approximately 0.75 g cm-3, while the

ACS Paragon Plus Environment

14

Page 15 of 31

Environmental Science & Technology

281

diesel aggregates plateau at an effective density of 0.6 g cm-3. The shape factors of the coated

282

aggregates from the burners converge to unity at a mass growth factor of approximately four,

283

while that of the diesel aggregates converge at mass growth factors closer to 10.

284

Despite the above differences between 100 nm aggregates from the burners and the diesel

285

generator, the evolution of initially 100 nm aggregates from the diesel generator is remarkably

286

similar to that of initially 200 nm aggregates from the burners, as depicted in Figure S6. In both

287

cases, aggregates cease restructuring at a mass growth factor of six, and coated particles reach

288

sphericity at a mass growth factor of about 10. Interestingly, the 100 nm diesel aggregates and

289

200 nm inverted burner aggregates comprise similar numbers of primary particles, 70 and 67,

290

respectively (see Table S1), suggesting that restructuring is related to primary particle number.

291

This notion is explored in greater detail below.

292

Final aggregate morphology. The compaction of coated-denuded aggregates can be

293

visualized using representative TEM samples collected before and after the photo-oxidation

294

stage of a typical experiment, shown in Figure 5a-b for initially 100 nm inverted burner

295

aggregates. As shown in Figure 5c, the final diameter growth factor, the ratio of the final

296

restructured aggregate mobility diameter to the initial aggregate mobility diameter, decreases for

297

all three sources as a function of initial mobility diameter. However, at a given initial mobility

298

diameter, the degree of restructuring varies significantly with the source of the soot. For

299

example, at an initial mobility diameter of about 100 nm, aggregates from the diesel generator

300

collapse appreciably (Gfd = 0.92), but those from the burners do not (Gfd = 0.99). Diesel and

301

inverted burner aggregates initially 100 nm in mobility diameter comprise about 70 and 22

302

primary particles, respectively; these estimates, based on the indirect method described above

303

(see Eq. 3), suggest that in general an aggregate comprised of more primary particles

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 31

304

accommodates more restructuring. If we compare initially 100 nm diesel aggregates and initially

305

200 nm inverted burner aggregates, as above, there is a striking similarity between the final

306

diameter growth factors (0.92 compared to 0.90) and primary particle numbers (70 compared to

307

67).

308

Motivated by the above comparisons between aggregates comprising roughly the same

309

number of primary particles, we consider here whether there is a systematic relationship between

310

the final degree of restructuring and primary particle number. A plot of the final diameter growth

311

factor as a function of the number of primary particles per aggregate (see Figure 5d) reveals an

312

approximately linear relationship across the range of primary particle numbers within the data

313

set, which is most relevant to ambient soot; this relationship may reflect the dilation symmetry of

314

the fractal-like aggregates. If separate fits are performed for each source, all three are nearly

315

identical, indicating that the source of the soot is important only to the extent that it dictates the

316

size and number of primary particles. Note that the linear fits cannot be used for extrapolation, in

317

the event the aggregates are composed of more than about 150 primary particles. Furthermore,

318

aggregates with similar numbers of primary particles collapse to a similar extent at any given

319

coating ratio, even before reaching the final morphology; for example, at mass growth factor of

320

about five, 100 nm diesel aggregates and 200 nm inverted burner aggregates both collapse by

321

about 6% of their initial mobility diameters. Consequently, if modelers know only the coating

322

mass and the number of primary particles per aggregate for a given ambient sample, which can

323

be derived from mass-mobility relationships that are known for many types of soot,33,48, they can

324

predict the evolution of aggregate morphology. Since the extinction cross section of soot

325

aggregates increases with compaction,18 the above fits may equip modelers to make better

326

estimations of the evolution of the optical properties of soot aggregates.49

ACS Paragon Plus Environment

16

Page 17 of 31

Environmental Science & Technology

327

We note that the final coated-denuded aggregates from the three sources were not

328

spherical. Consider the final mass-mobility exponents of aggregates from the McKenna burner,

329

inverted burner, and diesel generator, as shown in Figure 2b: 2.63, 2.68, and 2.23, respectively.

330

These values are significantly greater than the initial mass-mobility exponents, but they are still

331

well below three, indicating that indeed these aggregates were non-spherical when restructuring

332

ceased. Further restructuring could occur if the surface tension of the coating increased; for

333

example, coatings of glycerol (64 mN m-1 at 20 °C) have been shown to result in final mass-

334

mobility exponents of three for aggregates generated by the McKenna burner.31 Consequently,

335

the linear relationships reported here, although accurate for the representative anthropogenic

336

SOA coating, could change for coatings of higher surface tension. In the atmosphere, the surface

337

tension of SOA may increase via uptake of water at high relative humidity or other

338

atmospherically-relevant species, such as sulfuric acid.50,10 Furthermore, photo-chemical aging

339

and the attendant increase in the oxygen-to-carbon ratio of the SOA, may also increase the

340

surface tension and will certainly increase the hygroscopicity.51

341 342

ASSOCIATED CONTENT

343

Supporting Information

344

The Supporting Information is available free of charge on the ACS Publications website at DOI:

345

___.

346

Figures of representative size distributions, volatile mass fractions and initial effective

347

densities of diesel aggregates, representative time series of diameter and mass growth factors,

348

trends in effective density and shape factor for very thickly-coated inverted burner aggregates,

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 31

349

comparisons between 100 nm diesel aggregates and 200 nm inverted burner aggregates; table of

350

primary particle diameters and numbers.

351

AUTHOR INFORMATION

352

Corresponding Author

353

*Phone: (780) 492-2341; fax: (780) 492-2200; e-mail: [email protected].

354

Present Address

355



356

Notes

357

The authors declare no competing financial interest.

358

ACKNOWLEDGMENTS

359

This project was undertaken with the financial support of the Government of Canada through the

360

Federal Department of the Environment and the Natural Sciences and Engineering Research

361

Council of Canada (NSERC). We thank René Zepeda for assisting in the use of the TPS. E.G.S.

362

thanks NSERC and Environment Canada for scholarships. W.J. holds a Tier I Canada Research

363

Chair in Cluster Science.

364

REFERENCES

365

(1)

Department of Chemistry, University of Toronto, Toronto, ON, Canada M5S 3H6

366 367 368

Ramanathan, V.; Carmichael, G. Global and Regional Climate Changes due to Black Carbon. Nat. Geosci. 2008, 1, 221–227.

(2)

Bond, T. C.; Bergstrom, R. W. Light Absorption by Carbonaceous Particles: An Investigative Review. Aerosol Sci. Technol. 2006, 40, 27–67.

ACS Paragon Plus Environment

18

Page 19 of 31

369

Environmental Science & Technology

(3)

Olfert, J. S.; Symonds, J. P. R.; Collings, N. The Effective Density and Fractal Dimension

370

of Particles Emitted from a Light-Duty Diesel Vehicle with a Diesel Oxidation Catalyst. J.

371

Aerosol Sci. 2007, 38, 69–82.

372

(4)

373 374

Park, K.; Cao, F.; Kittelson, D. B.; McMurry, P. H. Relationship between Particle Mass and Mobility for Diesel Exhaust Particles. Environ. Sci. Technol. 2003, 37, 577–583.

(5)

Keywood, M.; Kanakidou, M.; Stohl, A.; Dentener, F.; Grassi, G.; Meyer, C. P.; Torseth,

375

K.; Edwards, D.; Thompson, A. M.; Lohmann, U.; et al. Fire in the Air: Biomass Burning

376

Impacts in a Changing Climate. Crit. Rev. Environ. Sci. Technol. 2013, 43, 40–83.

377

(6)

Schwarz, J. P.; Gao, R. S.; Spackman, J. R.; Watts, L. A.; Thomson, D. S.; Fahey, D. W.;

378

Ryerson, T. B.; Peischl, J.; Holloway, J. S.; Trainer, M.; et al. Measurement of the Mixing

379

State, Mass, and Optical Size of Individual Black Carbon Particles in Urban and Biomass

380

Burning Emissions. Geophys. Res. Lett. 2008, 35, L13810.

381

(7)

382 383

2011, 45, 765–779. (8)

384 385

Sorensen, C. M. The Mobility of Fractal Aggregates: A Review. Aerosol Sci. Technol.

Barone, T. L.; Storey, J. M. E.; Youngquist, A. D.; Szybist, J. P. An Analysis of DirectInjection Spark-Ignition (DISI) Soot Morphology. Atmos. Environ. 2012, 49, 268–274.

(9)

Gysel, M.; Laborde, M.; Mensah, A. A.; Corbin, J. C.; Keller, A.; Kim, J.; Petzold, A.;

386

Sierau, B. Technical Note: The Single Particle Soot Photometer Fails to Reliably Detect

387

PALAS Soot Nanoparticles. Atmos. Meas. Tech. 2012, 5, 3099–3107.

388

(10) Pagels, J.; Khalizov, A. F.; McMurry, P. H.; Zhang, R. Y. Processing of Soot by

389

Controlled Sulphuric Acid and Water Condensation—Mass and Mobility Relationship.

390

Aerosol Sci. Technol. 2009, 43, 629–640.

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 31

391

(11) Ma, X.; Zangmeister, C. D.; Gigault, J.; Mulholland, G. W.; Zachariah, M. R. Soot

392

Aggregate Restructuring during Water Processing. J. Aerosol Sci. 2013, 66, 209–219.

393

(12) Mikhailov, E. F.; Vlasenko, S. S. Structure and Optical Properties of Soot Aerosol in a

394

Moist Atmosphere: 1. Structural Changes of Soot Particles in the Process of Condensation.

395

Izv. Atmos. Ocean. Phys. 2007, 43, 181–194.

396

(13) Miljevic, B.; Surawski, N. C.; Bostrom, T.; Ristovski, Z. D. Restructuring of

397

Carbonaceous Particles upon Exposure to Organic and Water Vapours. J. Aerosol Sci.

398

2012, 47, 48–57.

399 400

(14) Ghazi, R.; Olfert, J. S. Coating Mass Dependence of Soot Aggregate Restructuring due to Coatings of Oleic Acid and Dioctyl Sebacate. Aerosol Sci. Technol. 2013, 47, 192–200.

401

(15) Bambha, R. P.; Dansson, M. A.; Schrader, P. E.; Michelsen, H. A. Effects of Volatile

402

Coatings and Coating Removal Mechanisms on the Morphology of Graphitic Soot.

403

Carbon 2013, 61, 80–96.

404

(16) Xue, H.; Khalizov, A. F.; Wang, L.; Zheng, J.; Zhang, R. Effects of Coating of

405

Dicarboxylic Acids on the Mass−Mobility Relationship of Soot Particles. Environ. Sci.

406

Technol. 2009, 43, 2787–2792.

407

(17) Khalizov, A. F.; Xue, H.; Wang, L.; Zheng, J.; Zhang, R. Enhanced Light Absorption and

408

Scattering by Carbon Soot Aerosol Internally Mixed with Sulfuric Acid. J. Phys. Chem. A

409

2009, 113, 1066–1074.

410

(18) Radney, J. G.; You, R.; Ma, X.; Conny, J. M.; Zachariah, M. R.; Hodges, J. T.;

411

Zangmeister, C. D. Dependence of Soot Optical Properties on Particle Morphology:

412

Measurements and Model Comparisons. Environ. Sci. Technol. 2014, 48, 3169–3176.

ACS Paragon Plus Environment

20

Page 21 of 31

413 414

Environmental Science & Technology

(19) Liu, F.; Yon, J.; Bescond, A. On the Radiative Properties of Soot Aggregates – Part 2: Effects of Coating. J. Quant. Spectrosc. Radiat. Transf. 2016, 172, 134–145.

415

(20) Cappa, C. D.; Onasch, T. B.; Massoli, P.; Worsnop, D. R.; Bates, T. S.; Cross, E. S.;

416

Davidovits, P.; Hakala, J.; Hayden, K. L.; Jobson, B. T.; et al. Radiative Absorption

417

Enhancements Due to the Mixing State of Atmospheric Black Carbon. Science 2012, 337,

418

1078–1081.

419

(21) Mentel, T. F.; Wildt, J.; Kiendler-Scharr, A.; Kleist, E.; Tillmann, R.; Dal Maso, M.;

420

Fisseha, R.; Hohaus, T.; Spahn, H.; Uerlings, R.; et al. Photochemical Production of

421

Aerosols from Real Plant Emissions. Atmos. Chem. Phys. 2009, 9, 4387–4406.

422

(22) Shao, P.; An, J.; Xin, J.; Wu, F.; Wang, J.; Ji, D.; Wang, Y. Source Apportionment of

423

VOCs and the Contribution to Photochemical Ozone Formation during Summer in the

424

Typical Industrial Area in the Yangtze River Delta, China. Atmos. Res. 2016, 176–177,

425

64–74.

426

(23) Khalizov, A. F.; Lin, Y.; Qiu, C.; Guo, S.; Collins, D.; Zhang, R. Role of OH-Initiated

427

Oxidation of Isoprene in Aging of Combustion Soot. Environ. Sci. Technol. 2013, 47,

428

2254–2263.

429

(24) Schnaiter, M.; Linke, C.; Möhler, O.; Naumann, K.-H.; Saathoff, H.; Wagner, R.;

430

Schurath, U.; Wehner, B. Absorption Amplification of Black Carbon Internally Mixed

431

with Secondary Organic Aerosol. J. Geophys. Res. Atmos. 2005, 110, D19204.

432

(25) Saathoff, H.; Naumann, K.-H.; Schnaiter, M.; Schöck, W.; Möhler, O.; Schurath, U.;

433

Weingartner, E.; Gysel, M.; Baltensperger, U. Coating of Soot and (NH4)2SO4 Particles by

434

Ozonolysis Products of α-Pinene. J. Aerosol Sci. 2003, 34, 1297–1321.

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 31

435

(26) Schnitzler, E. G.; Dutt, A.; Charbonneau, A. M.; Olfert, J. S.; Jäger, W. Soot Aggregate

436

Restructuring Due to Coatings of Secondary Organic Aerosol Derived from Aromatic

437

Precursors. Environ. Sci. Technol. 2014, 48, 14309–14316.

438 439

(27) Qiu, C.; Khalizov, A. F.; Zhang, R. Soot Aging from OH-Initiated Oxidation of Toluene. Environ. Sci. Technol. 2012, 46, 9464–9472.

440

(28) Guo, S.; Hu, M.; Lin, Y.; Gomez-Hernandez, M.; Zamora, M. L.; Peng, J.; Collins, D. R.;

441

Zhang, R. OH-Initiated Oxidation of m-Xylene on Black Carbon Aging. Environ. Sci.

442

Technol. 2016, 50, 8605–8612.

443

(29) Peng, J.; Hu, M.; Guo, S.; Du, Z.; Zheng, J.; Shang, D.; Zamora, M. L.; Zeng, L.; Shao,

444

M.; Wu, Y.-S.; et al. Markedly Enhanced Absorption and Direct Radiative Forcing of

445

Black Carbon under Polluted Urban Environments. Proc. Natl. Acad. Sci. 2016, 113,

446

4266–4271.

447

(30) Tritscher, T.; Jurányi, Z.; Martin, M.; Chirico, R.; Gysel, M.; Heringa, M. F.; DeCarlo, P.

448

F.; Sierau, B.; Prévôt, A. S. H.; Weingartner, E.; et al. Changes of Hygroscopicity and

449

Morphology during Ageing of Diesel Soot. Environ. Res. Lett. 2011, 6, 34026.

450 451

(31) Schnitzler, E. G.; Gac, J. M.; Jäger, W. Coating Surface Tension Dependence of Soot Aggregate Restructuring. J. Aerosol Sci. 2017, 106, 43–55.

452

(32) Simpson, I. J.; Blake, N. J.; Barletta, B.; Diskin, G. S.; Fuelberg, H. E.; Gorham, K.;

453

Huey, L. G.; Meinardi, S.; Rowland, F. S.; Vay, S. A.; et al. Characterization of Trace

454

Gases Measured over Alberta Oil Sands Mining Operations: 76 Speciated C2–C10

455

Volatile Organic Compounds (VOCs), CO2, CH4, CO, NO, NO2, NOy, O3 and SO2. Atmos.

456

Chem. Phys. 2010, 10, 11931–11954.

ACS Paragon Plus Environment

22

Page 23 of 31

Environmental Science & Technology

457

(33) Ghazi, R.; Tjong, H.; Soewono, A.; Rogak, S. N.; Olfert, J. S. Mass, Mobility, Volatility,

458

and Morphology of Soot Particles Generated by a McKenna and Inverted Burner. Aerosol

459

Sci. Technol. 2013, 47, 395–405.

460

(34) Parsons, M. T.; Sydoryk, I.; Lim, A.; McIntyre, T. J.; Tulip, J.; Jäger, W.; McDonald, K.

461

Real-Time Monitoring of Benzene, Toluene, and p-Xylene in a Photoreaction Chamber

462

with a Tunable Mid-Infrared Laser and Ultraviolet Differential Optical Absorption

463

Spectroscopy. Appl. Opt. 2011, 50, A90–A99.

464

(35) Nejad, B. M.; Parsons, M. T.; Sydoryk, I.; Schnitzler, E.; Lim, A.; Tulip, J.; Jaeger, W.;

465

McDonald, K. Monitoring Multiple Trace Gas Species during Secondary Organic Aerosol

466

Formation in a Smog Simulation Chamber Using Mid-IR Laser and UV Spectroscopic

467

Methods. In Lasers, Sources, and Related Photonic Devices; Optical Society of America,

468

2012; LT5B.6.

469

(36) Radney, J. G.; Zangmeister, C. D. Practical Limitations of Aerosol Separation by a

470

Tandem Differential Mobility Analyzer–Aerosol Particle Mass Analyzer. Aerosol Sci.

471

Technol. 2016, 50, 160–172.

472

(37) Liu, P.; Zhang, Y.; Martin, S. T. Complex Refractive Indices of Thin Films of Secondary

473

Organic Materials by Spectroscopic Ellipsometry from 220 to 1200 nm. Environ. Sci.

474

Technol. 2013, 47, 13594–13601.

475 476

(38) Atkinson, R. Product Studies of Gas-Phase Reactions of Organic Compounds. Pure Appl. Chem. 2009, 70, 1335–1343.

477

(39) DeCarlo, P. F.; Slowik, J. G.; Worsnop, D. R.; Davidovits, P.; Jimenez, J. L. Particle

478

Morphology and Density Characterization by Combined Mobility and Aerodynamic

479

Diameter Measurements. Part 1: Theory. Aerosol Sci. Technol. 2004, 38, 1185–1205.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 31

480

(40) Park, K.; Kittelson, D. B.; Zachariah, M. R.; McMurry, P. H. Measurement of Inherent

481

Material Density of Nanoparticle Agglomerates. J. Nanopart. Res. 2004, 6, 267–272.

482

(41) Nakao, S.; Tang, P.; Tang, X.; Clark, C. H.; Qi, L.; Seo, E.; Asa-Awuku, A.; Cocker, D.,

483

III. Density and Elemental Ratios of Secondary Organic Aerosol: Application of a Density

484

Prediction Method. Atmos. Environ. 2013, 68, 273–277.

485 486 487 488 489 490 491 492 493 494

(42) Dastanpour, R.; Boone, J. M.; Rogak, S. N. Automated Primary Particle Sizing of Nanoparticle Aggregates by TEM Image Analysis. Powder Technol. 2016, 295, 218–224. (43) Dastanpour,

R.;

Boone,

J.

M.;

Rogak,

S.

N.

PCM

Program;

2016;

http://www.aerosol.mech.ubc.ca/research/soot-and-nanoparticles. (44) Maricq, M. M.; Xu, N. The Effective Density and Fractal Dimension of Soot Particles from Premixed Flames and Motor Vehicle Exhaust. J. Aerosol Sci. 2004, 35, 1251–1274. (45) Meakin, P.; Donn, B.; Mulholland, G. W. Collisions between Point Masses and Fractal Aggregates. Langmuir 1989, 5, 510–518. (46) Rogak, S. N.; Flagan, R. C.; Nguyen, H. V. The Mobility and Structure of Aerosol Agglomerates. Aerosol Sci. Technol. 1993, 18, 25–47.

495

(47) Eggersdorfer, M. L.; Gröhn, A. J.; Sorensen, C. M.; McMurry, P. H.; Pratsinis, S. E.

496

Mass-Mobility Characterization of Flame-Made ZrO2 Aerosols: Primary Particle Diameter

497

and Extent of Aggregation. J. Colloid Interface Sci. 2012, 387, 12–23.

498

(48) Dastanpour, R.; Rogak, S. N.; Graves, B.; Olfert, J.; Eggersdorfer, M. L.; Boies, A. M.

499

Improved Sizing of Soot Primary Particles Using Mass-Mobility Measurements. Aerosol

500

Sci. Technol. 2016, 50, 101–109.

ACS Paragon Plus Environment

24

Page 25 of 31

Environmental Science & Technology

501

(49) Wu, Y.; Cheng, T.; Zheng, L.; Chen, H. Sensitivity of Mixing States on Optical Properties

502

of Fresh Secondary Organic Carbon Aerosols. J. Quant. Spectrosc. Radiat. Transf. 2017,

503

195, 147–155.

504 505

(50) Lee, J. Y.; Hildemann, L. M. Surface Tensions of Solutions Containing Dicarboxylic Acid Mixtures. Atmos. Environ. 2014, 89, 260–267.

506

(51) Massoli, P.; Lambe, A. T.; Ahern, A. T.; Williams, L. R.; Ehn, M.; Mikkilä, J.;

507

Canagaratna, M. R.; Brune, W. H.; Onasch, T. B.; Jayne, J. T.; et al. Relationship between

508

Aerosol Oxidation Level and Hygroscopic Properties of Laboratory Generated Secondary

509

Organic Aerosol (SOA) Particles. Geophys. Res. Lett. 2010, 37, L24801.

510 511

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 31

512 513

Figure 1. Experimental setups during injection of soot aggregates generated by (a) the McKenna

514

premixed burner, (b) the inverted diffusion burner, and (c) the commercial diesel generator and

515

(d) during sampling. TD: thermo-denuder; CPMD: counter-flow parallel-plate membrane

516

denuder; DMA: differential mobility analyzer; CPC: condensation particle counter; CPMA:

517

centrifugal particle mass analyzer.

518

ACS Paragon Plus Environment

26

Page 27 of 31

Environmental Science & Technology

519 520

Figure 2. Mass-mobility relationships for soot aggregates from the three sources (a) before and

521

(b) after restructuring. Dashed curves show linear least-squares fits to the natural logarithms of

522

aggregate mass and mobility; the data points are weighted by their uncertainties, calculated from

523

experimental precision and instrumental bias. The data point at about 50 nm is the same in both

524

plots; the mass and mobility diameter were measured directly from the diesel generator treatment

525

train, and no photo-oxidation experiment was performed, because the number concentration after

526

classification was too low to counteract losses to the chamber walls during injection. Since very

527

little restructuring was observed for initially 70 nm diesel aggregates, initially 50 nm diesel

ACS Paragon Plus Environment

27

Environmental Science & Technology

528

aggregates are assumed to undergo no change in either mass or mobility. The constant C in the

529

fit functions (Eq. 1) has units of fg nm–7 .

Page 28 of 31

530

ACS Paragon Plus Environment

28

Page 29 of 31

Environmental Science & Technology

531 532

Figure 3. Trends in (a) effective density and (b) shape factor with increasing mass growth factor

533

for coated and coated-denuded burner-generated aggregates initially about 250 nm in mobility

534

diameter.

535

ACS Paragon Plus Environment

29

Environmental Science & Technology

Page 30 of 31

536 537

Figure 4. Trends in (a) effective density and (b) shape factor with increasing mass growth factor

538

for coated and coated-denuded aggregates, initially about 100 nm in mobility diameter, generated

539

by all three sources.

540

ACS Paragon Plus Environment

30

Page 31 of 31

Environmental Science & Technology

541 542

Figure 5. TEM images of initially 100 nm inverted burner soot aggregates (a) before and (b)

543

after restructuring due to SOA coating, and final diameter growth factor as a function of (c)

544

initial mobility diameter and (d) number of primary particles per aggregate, estimated using the

545

indirect method; the uncertainty in the number of primary particles is detailed in the SI

ACS Paragon Plus Environment

31