Sunteți pe pagina 1din 11

Water Air Soil Pollut

DOI 10.1007/s11270-008-9826-5

The Odd–Even Behaviour of Dicarboxylic Acids Solubility


in the Atmospheric Aerosols
Mohd Zul Helmi Rozaini & Peter Brimblecombe

Received: 25 January 2008 / Accepted: 6 August 2008


# Springer Science + Business Media B.V. 2008

Abstract The solubility and the enthalpy of dicarbox- 1 Introduction


ylic acids have been determined in water at intervals
between 278.5 and 543.5 K. At 298.15 K, the values 1.1 Dicarboxlic Acids Distributions
derived were: ΔsolHm (m=1.33 mol kg−1)=29.80 kJ in the Atmosphere
mol−1for oxalic acid; ΔsolHm (m=16.03 mol kg−1)=
12.82 kJ mol−1 for malonic acid; ΔsolHm (m=0.75 Low molecular weight dicarboxylic acids (DCAs) are
mol kg−1)=28.20 kJ mol−1 for succinic acid; ΔsolHm ubiquitous and important components of the organic
(m=8.77 mol kg−1)=48.01 kJ mol−1 and ΔsolHm (m= fraction of atmospheric particles. They have been
0.17 mol kg−1)=40.30 kJ mol−1 for glutaric and detected in a variety of environments in many regions
adipic acid respectively. The solubility value exhibits in the world, including urban (Kawamura and Kaplan
a prominent odd–even effect with respect to terms 1987; Kawamura and Watanabe 2004; Limbeck and
with even number of carbon atoms with the odd Puxbaum 1999; Rohrl and Lammel 2001; Sempere and
carbon numbers showing much higher solubility. Kawamura 2003), rural (Kawamura and Kaplan 1987;
Observations made in the atmospheres suggest that Khwaja et al. 1995; Narukawa et al. 1999; Yu et al.
this odd–even effect may have implications for the 2005) and remote marine atmospheres (Kawamura and
relative abundance of these acids in aerosols. Sakagushi 1999).
Whatever the region; urban and continental, or
Keywords Oxalic acid . Malonic acid . Succinic acid . remote marine (see Table 1), oxalic acid (C2: HOOC-
Glutaric acid . Adipic acid . Solubility . COOH) is found to be the most abundant diacid
Water soluble aerosol followed by succinic (C4: HOOC(CH2)2COOH) and/or
malonic (C3: HOOCCH2COOH) acid with concentra-
tions of several hundreds of nanograms per cubic meter
in urban and continental regions (Kawamura and
M. Z. H. Rozaini (*) : P. Brimblecombe Ikushima 1993; Kawamura and Kaplan 1987) to a
School of Environmental Science, few tens of nanograms per cubic meter in remote marine
University of East Anglia,
boundary layer (Kawamura and Sakagushi 1999;
Norwich, Norfolk NR4 7TJ, UK
e-mail: mohd.rozaini@uea.ac.uk Sempere and Kawamura 2003). In Europe, the most
continuous study of diacids was conducted for over 1
M. Z. H. Rozaini year by Limbeck et al. (2005) at Vienna, Austria.
Department of Chemical Sciences,
Although available data on diacids are more sparse at
Faculty of Science and Technology,
University Malaysia Terengganu, mid-latitudes in Europe, they tend to show that oxalic
21030 Kuala Terengganu, Malaysia acid levels at nonurban or rural sites are not consider-
Water Air Soil Pollut

Table 1 Summary of aerosol dicarboxylate concentration (ng m-3) in urban and continental (a); remote marine (b) locations

References Location Oxalic Malonic Succinic Glutaric Adipic

(a)
(Grosjean et al. 1978) New York – 3.6 21.8 17.2 13.2
(Grosjean et al. 1978) New York – 3.9 24.9 23.2 11.6
(Kawamura and Kaplan 1987) West LA 6.38 1.58 1.96 0.6 2.22
(Kawamura and Kaplan 1987) West LA 2.12 0.4 0.66 0.22 0.94
(Kawamura and Kaplan 1987) West LA 8.13 0.72 2.34 0.66 3.31
(Kawamura and Kaplan 1987) West LA 8.65 1.45 2.37 0.74 0.49
(Kawamura and Kaplan 1987) Downtown LA 6.21 0.71 1.19 0.52 0.1
(Kawamura and Kaplan 1987) Downtown LA 6.6 0.76 1.84 0.52 0.2
(Kawamura and Kaplan 1987) Downtown LA 8.31 1.22 2.13 0.83 0.63
(Sempere and Kawamura 1994) Tokyo 29.65 6.69 13.18 3.72 6.66
(Sempere and Kawamura 1994) Tokyo 58.89 20.29 28.82 7.54 6.79
(Sempere and Kawamura 1994) Tokyo 330 141.3 161.1 4.15 2.91
(Limbeck and Puxbaum 1999) South Africa 193 142 58 8.8 7.9
(Limbeck and Puxbaum 1999) Sonblick observatory 153 22 14 2.7 4.4
(Limbeck and Puxbaum 1999) Vienna 340 244 117 26 117
(Kawamura and Watanabe 2004) Tokyo 357 71.4 73.4 23.1 25.8
(Kawamura and Watanabe 2004) Tokyo 157 44 41 11 13
(Kawamura and Watanabe 2004) Tokyo 186 40.5 47.4 18.2 14.2
(Rohrl and Lammel 2001) Helsinki – – – – –
(Ho et al. 2006) Hong Kong (road) 478 89.1 71.88 20 10.7
(Ho et al. 2006) Hong Kong (road) 268 47.6 33 6.95 12.7
(Hsieh et al. 2007) Tainan, Taiwan 574 65.8 101 43 13.2
(Hsieh et al. 2007) Tainan, Taiwan 432 34.2 87.9 10.3 8.8
(Limbeck et al. 2005) Vienna, Austria 99.6 34 37 7.7 3.3
(Limbeck et al. 2005) Vienna, Austria 66.2 38.6 30.8 6.6 3.2
(Limbeck et al. 2005) Vienna, Austria 63.1 21.5 31.2 5.6 2.5
(Limbeck et al. 2005) Mt Rax, Austria 34.5 9.1 16.4 2.3 0.8
(Limbeck et al. 2005) Mt Rax, Austria 26.4 6.9 14.9 2.3 4.3
(Limbeck et al. 2005) Mt Rax, Austria 32.6 16.4 22.4 3 1.7
(Decesari et al. 2006) Rondonia, Brazil 194.7 73.1 123.5 23.5 14.5
(Decesari et al. 2006) Rondonia, Brazil 793.3 56.8 210.2 32.1 12.6
(Decesari et al. 2006) Rondonia, Brazil 937.9 128.5 423.9 34.7 21.2
(Decesari et al. 2006) Rondonia, Brazil 1260 476.5 667.2 121.1 97.4
(Wang et al. 2006) Hong Kong (Tunnel) 505 69.4 85.2 20.9 26.4
(Wang et al. 2006) Hong Kong (Tunnel) 221 34.5 32.7 14.7 13.5
(Wang et al. 2006) Hong Kong (Tunnel) 234 42 51.4 17.1 24.7
(Wang et al. 2006) Hong Kong (Tunnel) 312 59.7 62.9 16.7 15.5
(Wang et al. 2006) Hong Kong (Tunnel) 633 59.3 95.1 30.3 25.9
(b)
(Kawamura and Kaplan 1987) Green House LA 1.31 0.3 0.29 0.04 0.1
(Kawamura and Kaplan 1987) Green House LA 2.83 0.14 0.86 0 0.22
(Kawamura and Sakagushi 1999) North Pacific 44.7 23.2 19.5 2.57 3.08
(Kawamura and Sakagushi 1999) North Pacific 8.73 2.18 2.16 0.61 1.26
(Kawamura and Sakagushi 1999) North Pacific 10.6 1.98 2.22 0.23 2.12
(Kawamura and Sakagushi 1999) North Pacific 28.6 12.8 13 1.84 1.34
(Kawamura and Sakagushi 1999) North Pacific 667 189 93 20.1 4.9
(Kawamura and Sakagushi 1999) North Pacific 190 38.6 16.7 10.2 2.76
(Kawamura and Sakagushi 1999) North Pacific 88.5 34.5 21.6 4.72 6.04
(Kawamura and Sakagushi 1999) North Pacific 24.9 5.66 10.1 1.87 1.67
(Kawamura and Sakagushi 1999) North Pacific 10 2.12 1.52 0.32 0.43
(Kawamura and Sakagushi 1999) North Pacific 18.3 3.45 4.02 0.62 0.46
Water Air Soil Pollut

Table 1 (continued)

References Location Oxalic Malonic Succinic Glutaric Adipic

(Kawamura and Sakagushi 1999) North Pacific 25.5 5.93 2.99 0.65 0.4
Kawamura (1996) Antarctic 1.59 0.13 0.63 0.31 0.49
Kawamura (1996) Antarctic 3.12 0.38 5.77 0.58 0.85
Kawamura (1996) Antarctic 3.26 0.52 1.18 0.34 0.33
Kawamura (1996) Antarctic 10.29 2.69 61.53 2.26 1.81
Narukawa et al. (1999) Indonesia 2200 800.3 1090 310 350
Narukawa et al. (1999) Indonesia 225 18.4 123 30 40
Khwaja et al. (1995) Semi urban site NY 308 84 55 12 89
Khwaja et al. (1995) Semi urban site NY 245 92 106 16.3 101
Khwaja et al. (1995) Semi urban site NY 118 165 107 15 40
Khwaja et al. (1995) Semi urban site NY 58 81 129 20 21
Khwaja et al. (1995) Semi urban site NY 298 96 90 23 31
Khwaja et al. (1995) Semi urban site NY 1 43 0.5 39 20
Khwaja et al. (1995) Semi urban site NY 360 88 167 46 50
Sempere and Kawamura (2003) Western Pacific 428.5 78.6 33.4 7.6 7.2
Rohrl and Lammel (2001) Rural (I) – – 14 – –
Rohrl and Lammel (2001) Rural (II) – – 8.8 – –
Rohrl and Lammel (2001) Rural (III) – – 18 – –
Kawamura et al. 2007 Canadian arctic 9.89 2.74 2.16 0.54 0.51
Kawamura et al. 2007 Canadian arctic 8.3 2.87 1.44 0.37 0.26
Kawamura et al. 2007 Canadian arctic 5.26 1.67 1.08 0.22 0.27
Narukawa et al. 2002 Arctic, Alert 23.5 5.03 3.21 1.21 0.54
Narukawa et al. 2002 Arctic, Alert 40.09 11.6 15.67 2.16 0.55
Mochida et al. 2007 North Pacific, ACE 600 110 52 8.9 2

ably different from those at urban sites (Limbeck and nuclei (Kerminen et al. 2000) and may contribute to
Puxbaum 1999; Rohrl and Lammel 2001). the acidity of precipitation (Chebbi and Carlier 1996).
Motor exhausts have been proposed to be primary With a low vapour pressure and high water solubility,
sources of oxalic, malonic, succinic, and glutaric (C5: diacids have an influence on the chemical and physical
HOOC(CH2)3COOH) acids (Grosjean et al. 1978; properties of aerosol (Lightstone et al. 2000). Conse-
Kawamura and Kaplan 1987). Some of these diacids quently, they may have direct and indirect effects on
are also emitted by wood burning, particularly malonic the earth’s radiation balance by scattering incoming
acid (pine wood) and succinic acid (oak wood; Rogge solar radiation, which counteracts the global warming
et al. 1991, 1993). Until now no direct source of malic caused by the increase of greenhouse gasses.
(hydroxysuccinic: hC4: HOOCCH2CHOHCOOH) and As an initial step towards a better understanding on
tartaric (dihydroxysuccinic: dhC4: HOOC(CHOH)2- the physical chemistry of these multiphase com-
COOH) acids has been identified for aerosols. pounds, laboratory studies into the solubility of the
Glutaric, succinic, and adipic (C6: HOOC(CH2)4 acids have been undertaken. It is also particularly
COOH) acids have been identified in laboratory studies noticeable that there is odd and even effect of carbon
(Hatakeyama et al. 1985) as secondary organic aerosol atom number on dicarboxylic acid solubility, which
products of the reaction of O3 with cyclohexene, a may have an influence on the phase and chemical
symmetrical alkene molecule similar to monoterpenes properties of atmospheric aerosols.
emitted by the biosphere. Hatakeyama et al. (1985)
also suggested that malonic and oxalic acids are also 1.2 Solubility
produced in the cyclohexene-ozone system.
The DCAs have received much attention because The dissolution of compounds in aqueous aerosols
of their roles in affecting the global climate by affects their water activity and thus the related to the
involving in the production of cloud condensation relative humidity at which the aerosol equilibrates
Water Air Soil Pollut

with the atmosphere. Highly soluble compounds can respectively) and the substances were used without
form very concentrated solutions and can remain so further purification.
even at low relative humidity. It is also possible for
aerosol droplets to persist as supersaturated systems. 2.2 Solubility Determination
The low molecular weight DCAs and dicarboxylates
have a strong affinity for moisture so will absorb The simple apparatus used to determine solubility is
relatively large amounts of water from the atmosphere shown in Fig. 1. A two-hole rubber stopper fitted
to form a liquid solution The relative humidity at snugly into a test tube (25×150 mm) of borosilicate
which such deliquescence takes place is also depen- glass. The thermometer was inserted carefully through
dent on the solubility of the compound. There are one hole of the rubber stopper so that the thermometer
surprisingly few measurements of the solubility of tip was near the bottom of the test tube. A glass rod
DCAs especially in electrolytes of relevance to the was inserted through other hole, so that the solution
atmosphere (such as sodium chloride and ammonium could be agitated. The large borosilicate beaker (2 L)
sulphate solutions), so we have undertaken a range of contained water and sat on a hot plate. The solid acids
solubility measurements for low molecular weight and were weighed accurately into the test tube and
DCAs. distilled water (initially about 2.00 mL) was dis-
Solubility describes the composition of solutions pensed from the burette into the test tube and
and in the case of aqueous solutions of soluble recorded to 0.01 mL.
organic materials it represents the amount material The solution in the tube was heated (to about 80°C)
that will dissolve in a liquid. The aqueous solubility and once the acid dissolved was allowed to cool, while
of a solute is often expressed as the number of moles being agitated, to determine at which temperature the
of solid dissolved in a certain amount of water, here a. acid begins to crystallize. The temperature was noted
kilogram, such that the units are referred as molality, and a small amount of water was added from the burette
m (mol kg−1). Although volumetric measurements are and recorded. The solution was heated again until the
popular in the laboratory for ease of preparation (i.e. acid dissolved and cooled once more to record a new
molar units, or moles of solute per litre of solution), point of crystal formation. This was repeated many
they are inconvenient in terms of physical representa- times such that the solubility (ever decreasing) at
tions. The concentration changes with temperature various temperatures could be determined and it is these
and at high concentrations the solute solvent ratio is that appear as our data tabulated within Figs. 2, 3, 4, 5
not proportional to the expressed concentration. Here and 6. Solubility is plotted as the logarithm of the
we report the variation of aqueous solubility with solubility (ln m vs 1/T) as this is preferable from a
temperature for the 2–6 carbon acids. physical chemistry perspective as slope of the curve is
related to the molar enthalpy of dissolution.

2 Experimental
3 Results and Discussions
2.1 Reagents
3.1 Solubility of Dicarboxylic Acids
The reagents were used throughout this work stated
by the molecular weight, Mw were: oxalic acid, Mw = Solubility is a quantitative means of describing the
90.636 g mol−1; malonic acid, Mw =104.06 g mol−1; composition of solutions. Solutions may be formed by
succinic acid, Mw =118.19 g mol−1; glutaric acid, Mw = mixing various combinations of liquids, gases, and
132.11 g mol−1; adipic acid, Mw =146.14 g mol−1; solids. In these experiments only the dissolution of a
sodium chloride; Mw =58.44 g mol−1; ammonium solid in a liquid will be considered. Thus, solubility as
nitrate, Mw =80.04 g mol−1; ammonium sulphate, Mw = used to describe a solid dissolved in a liquid refers to
132.14 g mol−1 and water, Mw =18.0153 g mol−1. There the quantitative composition of the solution. Solid
were obtained from Aldrich with analytical grade acids were dissolved and concentration at various
reagents purities >99% (with the exception of glutaric saturation temperatures expressed as mol per kg of
and adipic acid with the purity ~98 and >97% water. A solubility curve was drawn by plotting the
Water Air Soil Pollut
Fig. 1 The apparatus used
to determine solubility

mole per kilogram water on y-axis against saturation mental conditions could explain this small disagree-
temperature on x axis. The dependence of solubility ment, but it is clear that saturated oxalic acid forms
as function of temperature T can be expressed at approximately molal solutions in the atmosphere. At
constant pressure, by the relation: 348.15 K, the solubility of oxalic acid we determined
the solubility as 7.23 mol kg−1 and there are no
fd ln m=d ð1=T Þgð1 þ d ln g=d ln mÞ ¼ Δsol Hm =R ð1Þ
solubility measurements beyond this temperature.
Where ΔsolHm is the molar enthalpy of solution, R Using the equation above for oxalic acid the molar
is the gas constant and m and γ is the solute and its enthalpy (i.e. neglecting non-ideality) of solution at
activity coefficient. 298.15 K is ΔsolHm(m=1.33 mol kg−1)=29.8 kJ mol−1
The solubility values of oxalic acid (1.33 mol kg−1) (see Fig. 2).
from this study were similar to those of Apelblat The solubility of malonic acid (Fig. 3) was in
(1989) and Marcoli et al. (2004) at 298.15 K; with reasonable agreement with literature values from
almost same slope over the range of 293.15–338.15 K. 278.15–323.15 K, but at higher temperatures (up to
The results Stephen and Stephen found a slightly 343.15 K) there is a slight difference. As a compar-
lower solubility of 1.10 mol kg−1at 298.15. Experi- ison, at 298.15 K, our solubility measurement for
Water Air Soil Pollut
Fig. 2 Solubility of oxalic
acid in the water. Key: This
study (square), Apelblat
(1986) (circle), Stephens and
Stephens (1986) (triangle)
and Marcolli et al. (2004)
(diamond).m=mol kg−1

malonic acid is 16.02 mol kg−1 in satisfactory molar enthalpy of solution at infinite dilution by
agreement with Apelblat (1986) and also from the Apelblat (1986) which is (m=15.99 mol kg−1)=
older ones from Cahn (1899) cited by Seidell and 11.20 kJ mol−1.
Linke (1952) of 15.99 and 15.92 mol kg−1 respec- Results of our solubility determination for succinic
tively. However, there is a slight disagreement with acid were compared with the values measured over the
the measurements of Marcolli et al. (2004) and period 1874 to 1919 as compiled by Seidell and Linke
(Stephens and Stephens 1986), which are almost (1952). This acid is notably less soluble than oxalic
half the value at 8.15 and 6.71 mol kg−1 respectively. and malonic acid. In fact, bulk solution measurements
Studies of bulk solutions by Peng and Chan (2001) are restricted to molalities of less than 0.7 mol kg−1. At
suggest solubility that is reasonably consistent with 298.15 K, the solubility of succinic acid in our study
our measurement, so we have discarded the lower shows almost the same values with others in the
values. The molar enthalpy of solution for malonic literature at 0.748 mol kg−1. There was apparent
acid at 298.15 K is ΔsolHm (m=16.02 mol kg−1)= problem in the data of Apelblat and Manzurola (1990)
12.82 kJ mol −1 . This value is slightly higher where we can see an obvious decrease in solubility at
than that based on the calorimetric values from the 333.15 K which seems rather unusual (see Fig. 4). We

Fig. 3 Solubility of malonic


acid in the water. Key: This
study (square), Apelblat
(1986) (circle), Stephens and
Stephens (1986) (triangle)
and Marcolli et al. (2004)
(diamond). m=mol kg−1
Water Air Soil Pollut
Fig. 4 Solubility of succinic
acid in the water. Key: This
study (square), Apelblat
(1986) (circle), Stephens and
Stephens (1986) (triangle)
and Marcolli et al. (2004)
(diamond). m=mol kg−1

have discarded this anomalous value and determined (m=8.768 mol kg−1)=48.01 kJ mol−1 can be com-
the enthalpy from their data (28.29 kJ mol−1) which is pared with the calometric result of Apelblat (1989):
similar to our own 28.20 kJ mol−1. Our measured solu- ΔsolHm (m=8.956 mol kg−1)=51.01 kJ mol−1.
bility for succinic acid at 298.15 K is 0.748 mol kg−1, Results of our solubility determination for adipic
much in agreement with 0.747 from Apelblat and acid is 0.171 mol kg−1 were comparable with other
Manzurola (1990) as well. measured values: 0.169 mol kg−1 (Apelblat 1986),
Measurements of glutaric acid solubility are 0.173 mol kg−1 (Attane and Doumani 1949) and
available from Attane and Doumani (1949) along 0.172 mol kg−1 (Marcolli et al. 2004) as in Fig. 6.
with determinations perform by Apelblat (1989). As However, a measurement performed for adipic acid is
can be seen at Fig. 5, the agreement is satisfactory expressed as concentrations (Davies and Griffiths
only for temperatures up to 308.15 K. Above this 1953) but without knowledge of the densities of
point and up to 343.15 K there is an unexpected saturated solution cannot be used for comparison,
scatter of experimental points. However at 298.15 K, nonetheless, from estimated the values of 0.180, it is
we have a significant solubility values which ours is still in a good agreement with those presented here.
8.768 and 8.956 mol kg−1 by Apelblat (1989). Our Our value for the molar enthalpy of adipic acid is
molar enthalpy of glutaric acid at 298.15 K is ΔsolHm ΔsolHm (m=0.171 mol kg−1)=40.30 kJ mol−1 and can

Fig. 5 Solubility of glutaric


acid in the water. Key: This
study (square), Apelblat
(1989) (circle), Stephens and
Stephens (1986) (triangle)
and Marcolli et al. (2004)
(diamond). m=mol kg−1
Water Air Soil Pollut
Fig. 6 Solubility of adipic
acid in the water. Key: This
study (square), Apelblat
(1986) (circle), Stephens and
Stephens (1986) (triangle)
and Marcolli et al. (2004)
(diamond). m=mol kg−1

be compared with the calorimetric result of Apelblat negative charges on carbon atoms (Falk and Nelson
(1986): ΔsolHm (m=0.169 mol kg−1)=42.58 kJ mol−1. 1910). Although these ideas now seem discordant with
our views on covalent bonding (Burrows 1992), they
3.2 Odd–Even Effect perhaps relate to resonance within covalence bonding
(Muller and Carl 2006). Larsson (1966) suggested that
A series of organic compounds with a chain of carbon packing in the planes containing the end groups
atoms often show an affect of physical property explained the effect. An increase in packing density
dependent on whether there is an odd or even number will increases the intermolecular interactions, notably
of carbon atoms. The aqueous solubility of DCAs is Van der Waals’s forces. These will stabilize the solid
highly dependent on the number of carbon atoms with phase and the particular alternation of the properties is
acids with odd numbers of carbon atoms being much
more soluble than those with even numbers.
Oxalic, adipic and succinic (even carbon numbers)
have much lower solubility than malonic and glutaric
acids, which are exceptionally soluble. The solubility (in
mole per kilogram) of the dicarboxylic acids is
presented as function of carbon chain length up to C8
in Fig. 7. A number of previous authors have observed
this particularly noticeable odd–even effect with
dicarboxylic acid solubility (e.g. Burrows 1992).
Cappa et al. (2007) and Larsson (1966) approached
this from an atmospheric science perspective in relation
the vapour pressure of (C4–C10) dicarboxylic acids.
The odd–even effect is often observed with longer
chain compounds particularly in terms of the triple
point of the mono-carboxylic acids and the Kraft
temperature which relates to surfactant properties. Fig. 7 Solubility of the dicarboxylic acids in the water at
Early theoretical explanations suggested that odd– 298.15 K. Also included are data for pimelic (square), suberic
(triangle) and azelaic acid (diamond) from Apelblat (1989) and
even differences arose from altering strong and weak
Apelblat and Manzurola (1990). Inset (i) shows the water
bonds along the carbon chain and others modified this activity of the saturated dicarboxylic acids from the thermody-
by introducing the idea of alternating positive and namic model of Pitzer (1991)
Water Air Soil Pollut

driven by differences at the interfaces between the In some cases succinic acid proved to be dominant
chain ends for even and odd members. among the DCAs detected, for example in the South-
The large differences in solubility between subse- eastern Aerosol and Visibility Study (SEAVS). Here it is
quent homologues means that atmospheric aerosols of argued that high humidity allowed succinic acid to be
malonic and glutaric acid would remain liquid to present in the aerosol and thus protected from photo-
rather low relative humidity (~60% and 85% respec- oxidation (Yu et al. 2005). The SEAVS also gave
tively), while oxalic, succinic and adipic would importance to the C6 adipic acid reinforcing a view
readily crystallize. that DCAs with even carbon number predominate.
It remote continental areas it is possible that primary
3.3 Atmospheric Implications emissions from forest fires, soil or vegetation contributed
to the production of oxalic acid. Oxalic acid is also
Measurements of aerosol concentrations of DCAs formed in plants through incomplete oxidation of
from the literature are listed in Table 1 including carbohydrates, e.g., by fungi (Aspergillus niger) or
information on the nature of the site. We chose to bacteria (acetobacter) and in the animal kingdom
divide the data into two groups: (1) urban and through carbohydrate metabolism via tricarboxylic acid
continental measurements and (2) remote marine cycle (Clarke 1986). Mammalian urine also contains a
measurements in Fig. 8. Not unexpectedly, the small amount of oxalate. Moreover, some of the authors
reported urban concentrations are higher than those believe that oxalic acid can also be produced from the
from the remote marine sites. The distributions of photo-oxidation of unsaturated fatty acids emitted from
dicarboxylic acids with respect to carbon chain seawater (Kawamura and Sakagushi 1999).
length, shows the predominance of oxalic acid (see The odd–even pattern in composition is known from
Fig. 8). This is typically reported from urban areas petroleum geology as the carbon preference index
such as Tokyo, Los Angeles, Johannesburg, Christ- (CPI). This index is related to Σ (odd homologues)/Σ
church and Vienna and even in non-urban locations. (even homologues) for high molecular weight n-alkanes
In urban locations the primary emission from vehicles traces their biological origin. This has also been used in
(incomplete combustion products of fuel or fuel the atmosphere where a CPI in the order of 3 (i.e. odd
additives) are a potential source of oxalic acid, but carbon numbers dominate) shows recent biological
we have to imagine other sources for the non-urban origin of alkanes and oxygenated organic compounds
location as here oxalic acid still predominates. from leaf waxes etc. (Simoneit et al. 1990). The signal is
Succinic acid is usually at the second highest not always clear for carboxylic acids.
concentrations in both locations followed by malonic, It may well be that the dominance of even carbon
adipic and glutaric acid. numbers among the smaller DCAs found in aerosols
reflects a physical process driven by the large solubility
differences over the series rather than an initial biological
signature. An odd–even dependence with respect to total
carbon number is observed in the vapour pressure of 4–
12 carbon DCAs, with higher pressures fund in the odd
members (Cappa et al. 2007). However, this implies that
larger DCAs with an even number of carbon atoms
would partition more effectively onto aerosol particles.
The reason for the predominance of even carbon
numbers among the low molecular weight DCAs in
aerosol is not clear. It could also be affected by the
presence of aqueous electrolytes and other aerosol
components. Nevertheless, the strong variation in
their solubility has to be considered as a potential
driver of this variation and reminds us of potential
Fig. 8 Comparison of average dicarboxylic acids concentra- importance a relatively simple physical interaction in
tions in urban and continental and remote marine locations moderating aerosol composition.
Water Air Soil Pollut

4 Conclusion Decesari, S., Fuzzy, S., & Monica, E. (2006). Characterization


of the organic composition of aerosols from Rondonia,
Brazil, during the LBA-SMOCC 2002 experiment and its
The smaller dicarboxylic acids (2–6 carbon atoms) are representation through model compounds. Atmospheric
soluble compounds commonly found in the atmo- Chemistry and Physics, 6, 375–402.
spheric aerosol. Measurements show that the acids Falk, K. G., & Nelson, J. M. (1910). The electron conception of
with odd numbers of carbon atoms (malonic and valence. Journal of the American Chemical Society, 32,
1637–1654. doi:10.1021/ja01930a012.
glutaric) have much higher solubility than those even Grosjean, D., Van Cauwenberghe, K., & Schmid, J. P. (1978).
with even numbers (oxalic, succinic and adipic). Identification of C3-C10 aliphatic dicarboxylic acids in
Although odd–even effects on physical properties airborne particulate matter. Environmental Science &
have been observed before, they are typically small. Technology, 12, 313–317. doi:10.1021/es60139a005.
Hatakeyama, S., Tanonaka, T., Weng, J., Bandow, H., Takagi,
The extraordinarily large variations in solubility from H., & Akimoto, H. (1985). Ozone-cyclohexene reaction in
one DCA homologue to the next may well leave its air: Quantitative analyses of particulate products and the
imprint on aerosols. Observations from aerosol reaction mechanism. Environmental Science & Technolo-
composition illustrate an odd even pattern in DCA gy, 19, 935–942. doi:10.1021/es00140a008.
Ho, K. F., Lee, S. C., Cao, J. J., Kawamura, K., Watanabe, T.,
abundance give some support to this possibility. Cheng, Y., & Chow, J. C. (2006). Dicarboxylic acids,
ketocarboxylic acids and dicarbonyls in the urban roadside
Acknowledgement The corresponding author would like to area of Hong Kong. Atmospheric Environment, 40, 3030–
thank the Malaysian Government for financially supporting this 3040. doi:10.1016/j.atmosenv.2005.11.069.
research under a scholarship. Hsieh, L.-Y., Kuo, S.-C., Chen, C.-L., & Tsai, Y. I. (2007).
Origin of low-molecular-weight dicarboxylic acids and
their concentration and size distribution variation in
suburban aerosol. Atmospheric Environment, 41, 6648–
References 6661. doi:10.1016/j.atmosenv.2007.04.014.
Kawamura, K., & Ikushima, K. (1993). Seasonal changes in the
Apelblat, A. (1986). Enthalpy of solution of oxalic, succinic, distribution of dicarboxylic acids in the urban atmosphere.
adipic, maleic, malic, tartaric, and citric acids, oxalic acid Environmental Science & Technology, 27, 2227–2235.
dihydrate, and citric acid monohydrate in water at 298.15 doi:10.1021/es00047a033.
K. The Journal of Chemical Thermodynamics, 18, 351– Kawamura, K., & Kaplan, I. R. (1987). Dicarboxylic acids
357. doi:10.1016/0021-9614(86)90080-7. generated by thermal alteration of kerogen and humic
Apelblat, A. (1989). Solubility of ascorbic, 2-furancarboxylic, acids. Geochimica et Cosmochimica Acta, 51, 3201–3207.
glutaric, pimelic, salicyclic, and o-phtalic acids in water at doi:10.1016/0016-7037(87)90128-1.
298.15 K. The Journal of Chemical Thermodynamics, 21, Kawamura, K., Kasukabe, H., and Barrie, L. A. (1996). Source
1005–1008. doi:10.1016/0021-9614(89)90161-4. and reaction pathways of dicarboxylic acids, ketoacids and
Apelblat, A., & Manzurola, E. (1990). Solubility of suberic, dicarbonyls in arctic aerosols: One year observations.
azelaic, levulinic, glycolic and diglycolic acids in water from Atmospheric Environment, 30, 1709–1722.
278.15 K–361.35 K. The Journal of Chemical Thermody- Kawamura, K., Narukawa, M., Li, S.-M., & Barrie, L. A.
namics, 22, 289–292. doi:10.1016/0021-9614(90)90201-Z. (2007). Size distributions of dicarboxylic acids and
Attane, E. C., & Doumani, T. F. (1949). Industrial & Engineering inorganic ions in atmospheric aerosols collected during
Chemistry, 41, 2015. doi:10.1021/ie50477a043. polar sunrise in the Canadian high Arctic. Journal of
Burrows, H. D. (1992). Studying odd–even effects and Geophysical Research D: Atmospheres, 112, 10–18.
solubility behaviour using 1,w-dicarboxylic acids. Journal Kawamura, K., & Sakagushi, F. (1999). Molecular distribution of
of Chemical Education, 69, 69. water soluble dicarboxylic acids in marine aerosols over the
Cappa, C. D., Lovejoy, E. R., & Ravishankara, A. R. (2007). pacific ocean including tropics. Journal of Geophysical
Determination of evaporation rates and vapor pressures of Research, 104, 3501–3509. doi:10.1029/1998JD100041.
very low volatility compounds: A study of the C4-C10 and Kawamura, K., & Watanabe, T. (2004). Determination of stable
C12 dicarboxylic acids. The Journal of Physical Chemis- carbon isotopic compositions of low molecular weight
try A, 111, 3099–3109. doi:10.1021/jp068686q. dicarboxylic acids and ketocarboxylic acids in atmospher-
Chebbi, A., & Carlier, P. (1996). Carboxylic Acids in the ic aerosol and snow samples. Analytical Chemistry, 76,
troposphere, occurrence, sources and sinks: A review. 5762–5768. doi:10.1021/ac049491m.
Atmospheric Environment, 30, 4233–4249. doi:10.1016/ Kerminen, V.-M., Kulmala, M., & Lehtinen, K. E. J. (2000).
1352-2310(96)00102-1. Low molecular weight dicarboxylic acids in an urban and
Clarke, H. T. (1986). Organic synthesis pp. 421–425. New rural atmosphere. Journal of Aerosol Science, 31, 349–
York: Wiley. 362. doi:10.1016/S0021-8502(99)00063-4.
Davies, M., & Griffiths, D. M. L. (1953). The solubilities of Khwaja, H. A., Brudnoy, S., & Husain, L. (1995). Chemical
dicarboxylic acids in benzene and aqueous solutions. characterization of three summer cloud episodes at
Transactions of the Faraday Society, 49, 1405–1410. whiteface mountain. Chemosphere, 31, 3357–3381.
doi:10.1039/tf9534901405. doi:10.1016/0045-6535(95)00187-D.
Water Air Soil Pollut

Larsson, K. (1966). Journal of the American Oil Chemists’ Rogge, W. F., Hildemann, L. M., Mazurek, M. A., Cass, G. R., &
Society, 43, 559–562. doi:10.1007/BF02641187. Simoneit, B. R. T. (1991). Sources of fine organic aerosol. 1.
Lightstone, J. M., Onasch, T. B., Imre, D., & Oatis, S. (2000). Charbroilers and meat cooking operations. Environmental
Deliquescence, efflorescence, and water activity in ammo- Science & Technology, 25, 1112–1125. doi:10.1021/
nium nitrate and mixed ammonium nitrate/succinic acid es00018a015.
microparticles. The Journal of Physical Chemistry A, 104, Rogge, W. F., Mazurek, M. A., Hildemann, L. M., Cass, G. R., &
9337–9346. doi:10.1021/jp002137h. Simoneit, B. R. T. (1993). Quantification of urban organic
Limbeck, A., Kraxner, Y., & Puxbaum, H. (2005). Gas to particle aerosols at a molecular level: Identification, abundance and
distribution of low molecular weight dicarboxylic acids at two seasonal variation. Atmospheric Environment—Part A
different sites in central Europe (Austria). Journal of Aerosol General Topics, 27A(8), 1309–1330.
Science, 36, 991–1005. doi:10.1016/j.jaerosci.2004.11.013. Rohrl, A., & Lammel, G. (2001). Low-molecular weight
Limbeck, A., & Puxbaum, H. (1999). Organic acids in continental dicarboxylic acids and glyoxylic acid: seasonal and air
background aerosols. Atmospheric Environment, 33, 1847– mass characteristics. Environmental Science & Technolo-
1852. doi:10.1016/S1352-2310(98)00347-1. gy, 35, 95–101. doi:10.1021/es0000448.
Manzurola, E., & Apelblat, A. (1985). The Journal of Chemical Seidell, A., & Linke, W. F. (1952). Solubilities of inorganic and
Thermodynamics, 17, 579. doi:10.1016/0021-9614(85) organic compounds. New York: Van Nostrand.
90057-6. Sempere, R., & Kawamura, K. (1994). Comparative distribu-
Marcolli, C., Luo, B. P., Peter, T., & Wienhold, F. G. (2004). tion of dicarboxylics acids and related polar compounds in
Internal mixing of the organic aerosol by gas phase snow, rain and aerosols from urban atmosphere. Atmo-
diffusion of semivolatile organic compounds. Atmospheric spheric Environment, 28, 449–459. doi:10.1016/1352-
Chemistry and Physics, 4, 2593–2599. 2310(94)90123-6.
Mochida, M., Umemoto, N., Kawamura, K., Lim, H. J., & Sempere, R., & Kawamura, K. (2003). Trans-hemispheric
Turpin, B. J. (2007). Bimodal size distributions of various contribution of C2-C10 a,w-dicarboxylic acids, and related
organic acids and fatty acids in the marine atmosphere: polar compounds to water-soluble organic carbon in the
Influence of anthropogenic aerosols, Asian dusts, and sea western Pacific aerosols in relation to photochemical
spray off the coast of East Asia. Journal of Geophysical oxidation reactions. Global Biogeochemical Cycles, 17,
Research D: Atmospheres, 112(15), DI5209. 38–1. doi:10.1029/2002GB001980.
Muller, U., & Carl, R. (2006). Fundamentals of preparative Simoneit, B. R. T., Cardoso, J. N., & Robinson, N. (1990). An
organic chemistry. New York: Springer. assessment of the origin and composition of higher
Narukawa, M., Kawamura, K., Li, S.-M., & Bottenheim, J. W. molecular weight organic matter in aerosols over Amazonia.
(2002). Dicarboxylic acids in the Arctic aerosols and snow- Chemosphere, 21, 1285–1301. doi:10.1016/0045-6535(90)
packs collected during ALERT 2000. Atmospheric Environ- 90145-J.
ment, 36, 2491–2499. doi:10.1016/S1352-2310(02)00126-7. Stephens, H., & Stephens, T. (1986). Solubilities of inorganic
Narukawa, M., Kawamura, K., Takeuchi, N., & Nakajima, T. and organic compounds. Oxford: Pergamon.
(1999). Distribution of dicarboxylic acids and carbon Wang, H., Kawamura, K., Ho, K. F., & Lee, S. C. (2006). Low
isotopic compositions in aerosols from 1997 Indonesian molecular weight dicarboxylic acids, ketoacids, and dicarbon-
forest fires. Geophysical Research Letters, 26, 3101–3104. yls in the fine particles from a roadway tunnel: Possible
doi:10.1029/1999GL010810. secondary production from the precursors. Environmental
Peng, C., & Chan, C. K. (2001). The water cycles of water-soluble Science & Technology, 40, 6255–6260. doi:10.1021/
organic salts of atmospheric. Atmospheric Environment, 35, es060732c.
1183–1192. doi:10.1016/S1352-2310(00)00426-X. Yu, L., Shulman, M., Kopperud, R., & Hildemann, L. (2005).
Pitzer, K. S. (1991). Ion interaction approach: Theory and data Characterization of organic compounds collected during
correlation. In K. S. Pitzer (Ed.). Activity coefficient in southeastern aerosol and visibility study: Water-soluble
electrolyte solution (pp.75–153). Boca raton, Florida; CRC organic species. Environmental Science & Technology, 39,
Press 707–715. doi:10.1021/es0489700.

S-ar putea să vă placă și