Measured Saturation Vapor Pressures of Phenolic ... - ACS Publications

Mar 6, 2017 - Carl J. Percival,. † and David Topping. †,‡. †. School of Earth, Environmental and Atmospheric Science, University of Manchester...
0 downloads 0 Views 1MB Size
Subscriber access provided by Queen Mary, University of London

Article

Measured saturation vapour pressures of phenolic and nitro-aromatic compounds Thomas J Bannan, Alastair Murray Booth, Benjamin T Jones, Simon O'Meara, Mark Howard Barley, Ilona Riipinen, Carl John Percival, and David O. Topping Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b06364 • Publication Date (Web): 06 Mar 2017 Downloaded from http://pubs.acs.org on March 7, 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 19

Environmental Science & Technology

1

Measured saturation vapour pressures of

2

phenolic and nitro-aromatic compounds

3

1*

Thomas J Bannan, 1A Murray Booth, 1Benjamin T Jones, 1Simon O’Meara, 1Mark H Barley, 3Ilona

4

Riipinen, 1Carl J. Percival, 1,2David Topping.

5

1

6

UK

7

2

National Centre for Atmospheric Science, University of Manchester, Manchester, UK.

8

3

Department of Environmental Science and Analytical Chemistry (ACES), Stockholm University,

9

Stockholm, Sweden.

School of Earth, Environmental and Atmospheric Science, University of Manchester, Manchester,

10

*Corresponding author: Thomas J. Bannan. 4.30 Simon Building, University of Manchester, Oxford

11

Road, Manchester, M13 9PL, Phone: 0161 306 6587, Fax: 0161 306 3951, Email:

12

[email protected]

13

TOC Art

14 15

Abstract

16

Phenolic and nitro-aromatic compounds are extremely toxic components of atmospheric aerosol

17

that are currently not well understood. In this paper, solid and sub-cooled liquid state saturation

18

vapour pressures of phenolic and nitro-aromatic compounds are measured using Knudsen Effusion

19

Mass Spectrometry (KEMS) over a range of temperatures (298–318 K). Vapour pressure estimation

20

methods, assessed in this study, do not replicate the observed dependency on the relative positions

21

of functional groups. With a few exceptions, the estimates are biased towards predicting saturation

22

vapour pressures that are too high, by 5 to 6 orders of magnitude in some cases. Basic partitioning

ACS Paragon Plus Environment

Environmental Science & Technology

23

theory comparisons indicate that overestimation of vapour pressures in such cases would cause us

24

to expect these compounds to be present in the gas state, whereas measurements in this study

25

suggest these phenolic and nitro-aromatic will partition into the condensed state for a wide range of

26

ambient conditions if absorptive partitioning plays a dominant role. Whilst these techniques might

27

have both structural and parametric uncertainties, the new data presented here should support

28

studies trying to ascertain the role of nitrogen containing organics on aerosol growth and human

29

health impacts.

30

Introduction

31

Aerosol particles are an important component of the Earth’s atmosphere, affecting visibility, human

32

health and climate on regional and global scales (Bilde et al., 2015). Mechanistic models attempt to

33

predict the evolving composition and microphysics of aerosol populations through the use of

34

fundamental properties of pure components and mixtures. A critical parameter in this regard is the

35

pure component equilibrium vapour pressure (here referred to as saturation vapour pressure) of the

36

constituent molecules. Given the chemical complexity of the organic fraction of atmospheric aerosol,

37

with potentially many hundreds of thousands of individual species (Jenkin et al., 2003, Aumont et al.,

38

2005) and the scarcity of existing experimental data, the saturation vapour pressure of a large

39

fraction of atmospherically relevant compounds must be estimated rather than comprehensively

40

measured. There are a number of estimation methods available (e.g. Joback et al., 1987; Stein and

41

Brown 1994; Myrdal and Yalkowsky, 1997), but these suffer from inherent bias due to the lack of

42

data for heavily oxygenated, low volatility multifunctional compounds in the fitting dataset, which

43

are the sort of molecules expected to be present in atmospheric aerosols (Bilde et al., 2015). This

44

has been recognised as a problem in predicting total mass and speciation (Barley and McFiggans,

45

2010, Booth et al., 2010, O’Meara et al., 2014) and there have been efforts to develop improved

46

estimation methods more suited to atmospherically relevant compounds (e.g. EVAPORATION

47

(Compernolle et al., 2011)). However, such methods still remain largely invalidated across broad

48

ranges in absolute saturation vapour pressure and functionality. The most recent comparison

49

between methods was carried out by O’Meara et al (2014), using a database of 90 compounds

50

covering a vapour pressure range of 60 Pa to 1.34x10-5 Pa at 298 K. Whilst a first order problem

51

might be the propagated uncertainty in total particulate mass, it is important to evaluate any

52

heterogeneity in predicting the concentration of certain chemical classes. Developing accurate

53

predictive methods is challenging due to the influence of additional groups not necessarily being

54

linear (intramolecular bonding, shielding or steric hindrance), and the spread between data from

55

different instruments (Bilde et al., 2015).

ACS Paragon Plus Environment

Page 2 of 19

Page 3 of 19

Environmental Science & Technology

56

With this in mind, this paper adds to the existing body of work in which we measure the saturated

57

vapour pressure of two related classes: naphthalene phenolic and nitro-aromatic compounds.

58

Oxygenated aromatics (acids and phenols) have been measured in atmospheric aerosols (Hemming

59

and Seinfeld 2001; Kawamura and Kaplan 1987; Yassineet al., 2009). Phenolic compounds are useful

60

tracers for anthropogenic emissions as they are produced from a variety of anthropogenic volatile

61

organic compound (VOC) precursors such as benzene, toluene and xylene (Tremp et al., 1993;

62

Berndt and Böge, 2006) and a major component of biomass burning (Yee et al., 2013). For resolving

63

discrepancies between instruments and models they are also useful for probing intramolecular

64

effects due to the relative ease of obtaining samples with phenol groups in different configurations.

65

Nitro-aromatic compounds have proved to be useful tracers for anthropogenic emissions (Grosjean,

66

1992) and biomass burning (Yee et al., 2013), whilst studies quantifying the overall role of nitrogen-

67

containing organic compounds on aerosol growth could benefit from more refined pure component

68

vapour pressures (Smith et al 2008; Duporte et al 2016). With regards to potential impacts, many of

69

these compounds are noted to be highly toxic (Kovacic and Somanathan, 2014). Relating

70

toxicological response to certain classes is not the focus of this study, yet growing evidence suggests

71

specific particle components might be more important for certain diseases (Cassee et al., 2013,

72

Atkinson et al., 2010) whilst the majority of correlating studies rely on existing epidemiological

73

techniques (Stanek et al., 2011). Even if mechanistic models struggle to predict adequate mass due

74

to missing process phenomena (e.g.McVay et al., 2016), resolving expected partitioning in measured

75

samples is important. More generally the data presented should lead to a more thorough evaluation,

76

and potentially improvement, of vapour pressure predictive techniques as there is a significant lack

77

of data for functionalised nitrogen containing organic compounds in the datasets used for fitting

78

model parameters (e.g. Nannoolal et al., 2008).

79

In the following text we first briefly present the experimental and modelling methodology before

80

discussing measured vapour pressure from a Knudsen Effusion Mass Spectrometer (KEMS) (Booth et

81

al., 2012) and comparisons with a range of estimation methods. We then discuss the need for

82

measurements from complimentary techniques to improve datasets used for refitting estimation

83

methods.

84

2.Experimental

85

The following compounds were purchased from Sigma-Aldrich and used with no further preparation:

86

1-Naphthol,

87

dihydroxynaphthalene, 2,7-dihydroxynaphthalene, 1,5-dihydroxynaphthalene, 1-nitronaphthalene,

88

1,3-dinitronaphthalene, 2-methyl-1-nitronaphthalene, ortho-amino-benzoic acid, meta-amino-

2-

Naphthol,

1,3-dihydroxynaphthalene,

2,3-dihydroxynaphthalene,

ACS Paragon Plus Environment

1,7-

Environmental Science & Technology

Page 4 of 19

89

benzoic acid, para-amino-benzoic acid, meta-nitrophenol, para-nitrophenol, ortho-nitroaniline,

90

meta-nitroaniline, para-nitroaniline and Glutarimide, all of which are solid at room temperature.

91

2.1 The Knudsen Effusion Mass Spectrometry (KEMS)

92

The KEMS system is identical to that used in previous studies (e.g. Booth et al., 2009) and an

93

overview of the measurement procedure is repeated here: to calibrate, a sample of known vapour

94

pressure is placed in the temperature controlled Knudsen cell, in this case para-anisic acid (VP at

95

298k = 5.24x10-4 Pa, (Colomina et al., 1978; Booth et al., 2012). The linearity of the one calibration

96

point for calibrating measurements over a wide range of vapour pressures has been tested

97

exhaustively for measurements at close to limits of detection of the instrument (1 × 10-8 Pa) and for

98

higher vapour measurements such as benzophenone. This was most recently validated through

99

comparison with other instruments, in work that will be presented at a later date, which presents

100

measurements agreeing well across techniques over a vapour pressure range of 1x10-2-1x10-7 (Pa).

101

The cell has a chamfered effusing orifice with a size ≤1/10 the mean free path of the gas molecules

102

in the cell. This ensures the orifice does not significantly disturb the thermodynamic equilibrium of

103

the samples in the cell (Hilpert, 2001). The resulting molecular beam is ionised by 70 eV electron

104

impact, then sampled by the mass spectrometer. After correcting for the ionization cross section of

105

the calibration compound, this produces a signal proportional to the vapour pressure.

106

After this calibration a sample of unknown vapour pressure can be measured. During sample

107

change, the first chamber with the Knudsen cell is isolated via the gate valve and vented to air

108

allowing the ioniser filament and secondary electron multiplier to be left on. The unknown vapour

109

pressures can be determined from the intensity of the mass spectrometer signal of the compound in

110

question. If the Knudsen number, the ratio of the mean free path of molecules to the size of the

111

effusion orifice, is high enough, then effusing gas does not significantly disturb the equilibrium in the

112

cell (Booth et al., 2009, Hilpert 1991; 2001). This allows the steady state vapour pressure, as

113

measured by the KEMS, to be as close as possible to the equilibrium vapour pressure.

114

Once the solid state vapour pressure, P (Pa), has been determined at a number of different

115

temperatures, the Clausius-Clapeyeron equation is used to derive enthalpies and entropies of

116

sublimation as described in Booth et al., (2009);

117

ln ܲ =

௱ுೞೠ್ ோ்

+

௱ௌೞೠ್ ோ

(1)

118

where T is the temperature (K), R is the ideal gas constant (J mol−1 K−1) and ΔHsub (J mol-1) and ΔSsub (J

119

mol-1 K-1) are the enthalpies and entropies of sublimation respectively. P (Pa) was obtained over a

ACS Paragon Plus Environment

Page 5 of 19

Environmental Science & Technology

120

range of 20 K in this work, starting at 298 K. The reported solid state vapour pressures at 298.5K

121

(P298) are calculated from the linear fit of ln P vs. 1/T used in the Clausius-Clapeyon equation.

122

2.2 Differential Scanning Calorimetry (DSC)

123

As noted in the proceeding section, since we measure the saturation vapour pressure above the

124

solid state, we need to convert this to sub-cooled reference state for comparison with predictive

125

techniques and use within dynamic aerosol models. Following previous studies (e.g. Booth et al.,

126

2010), the information needed to perform this conversion, melting points (Tm) and Enthalpies of

127

Fusion (ΔHfus), were measured using a TA instruments Q200 Differential Scanning Calorimeter (DSC).

128

The procedure is repeated here: heat flow and temperature were calibrated using an indium

129

reference, and heat capacity using sapphire reference. A heating rate of 10 K min-1 was used. 5 – 10

130

mg of sample was measured using a microbalance, then pressed into a hermetically sealed

131

aluminium DSC pan. A purge gas of N2 was used with a flow rate of 30 ml min-1. The reference was

132

an empty sealed pan of the same type. Data processing was performed using the `Universal Analysis'

133

software supplied with the instrument. Δcp,sl was estimated using Δcp,sl = ΔSfus (Mauger et al., 1972;

134

Grant et al., 1984).

135

3. Theory

136

3.1 Sub-cooled correction

137

The sub-cooled vapour pressure is derived from the value measured above the solid state using the

138

Prausnitz equation (Prausnitz et al., 1986):

139

ln

Pl ∆H fus  Tm  ∆c p , sl  Tm  ∆c p , sl Tm = ln  − 1 −  − 1 + Ps RTm  T R T R T  

(2)

140

where: P is the saturation vapour pressure (Pa) with the subscript ‘s’ referring to the solid and ‘l’ to

141

the sub-cooled liquid phase, ΔHfus is the enthalpy of fusion (J mol-1), Δcp,sl denotes the best estimate

142

of the underlying change in heat capacity between the liquid and solid state at the melting point (J

143

mol-1 K-1), T is the temperature (K), Tm is the melting point (K) (which is commonly used instead of

144

the triple point,(Tt) (Prausnitz et al., 1986).

145

3.2 Vapour pressure predictive techniques

146

The predictive techniques used here are the same as those in a recent evaluation paper by O’Meara

147

et al., (2014) with the exception of the EVAPORATION method which is currently only applicable to

148

aliphatic compounds. These methods are all based on group contributions where each group has a

ACS Paragon Plus Environment

Environmental Science & Technology

149

set of values associated with it for calculating the appropriate property obtained by fitting to data.

150

The accuracy of a given method is therefore very sensitive to the choice of data used to fit it. This is

151

especially true for groups where data is limited, for which this study is relevant.

152

Nine methods are evaluated in total. Four of the group contribution vapour pressure methods

153

extrapolate from a boiling point down to the temperature of interest. These four methods are;

154

Nannoolal et al (2008), Moller et al (2008), Myrdal and Yalkowsky (1997) and Lee-Kesler (Reid et al.,

155

1987). The two boiling point estimation methods that are then used are Nannoolal et al., (2004) and

156

Stein and Brown (1994), for a total of 8 methods. The method of Joback (1987) is excluded due to its

157

known biases (Barley and McFiggans, 2010). The ninth method is SIMPOL (Pankow and Asher, 2008)

158

that directly estimates the vapour pressure without requiring a boiling point value.

159

4. Results

160

4.1 Solid state vapour pressures

161

Solid state vapour pressures directly measured in the KEMS are reported in Table 1. Measurements

162

were made from 298 to 318 +/- 274 K and the Clausius-Clayperon equation was used to derive

163

enthalpies and entropies of sublimation, as noted in section 1. 1,3 and 2,3 dihydroxnapthalene

164

were measured up to 333 K to produce enough vapour, thus signal, to detect with the KEMS, and the

165

P298 was extrapolated from these measurements. However, 1,7/2,7/1,5-dihydroxynaphthalene could

166

not be detected even at elevated temperatures, from which we infer that the vapour pressure is less

167

than the current reported minimum detectable pressure in the KEMS which corresponds to 1 × 10-8

168

Pa as determined by Booth et al., (2009). Some general observations regarding functional group

169

positioning on the solid-state vapour pressure can be inferred from table 1. 1,3-

170

dihydroxynaphthalene is within the current reported detectable limit (P298 of 9 × 10-6 Pa) whilst 2,3-

171

dihydroxynaphthalene has a volatility two orders of magnitude higher than the compound with the

172

functional group at the 1,3 position. One possible explanation for this is intramolecular bonding,

173

where adjacent groups on a molecule can hydrogen bond with each other thus lowering the

174

apparent polarity of the molecule and raising the apparent volatility relative to a molecule with non-

175

adjacent groups. There is no discernable pattern for the positional effects (ortho-meta-para

176

isomerism) in the amino-benzoic acids, nitrophenols and nitroanilines. Previous work on different

177

ortho-meta-para substitutions of benzoic acid showed that in molecules with different functional

178

groups the isomer effects were the result of a complex interplay of steric, resonance effect and

179

polarity (Booth et al., 2012). As noted in Bilde et al (2015), measurements between different

ACS Paragon Plus Environment

Page 6 of 19

Page 7 of 19

Environmental Science & Technology

180

instruments can often diverge; resolving the true effect of functional group positioning is still not

181

clear and might benefit from quantum scale modelling (Schroeder et al., 2016, Kurten et al., 2016).

182 183

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 19

184

Table 1. Solid state vapour pressure at 298 K, enthalpies and entropies of sublimation. Estimated

185

errors on the vapour pressures are