ArticlePDF Available

Abstract and Figures

Aerosol acidity largely regulates the chemistry of atmospheric particles, and resolving the drivers of aerosol pH is key to understanding their environmental effects. We find that an individual buffering agent can adopt different buffer pH values in aerosols and that aerosol pH levels in populated continental regions are widely buffered by the conjugate acid-base pair NH4+/NH3 (ammonium/ammonia). We propose a multiphase buffer theory to explain these large shifts of buffer pH, and we show that aerosol water content and mass concentration play a more important role in determining aerosol pH in ammonia-buffered regions than variations in particle chemical composition. Our results imply that aerosol pH and atmospheric multiphase chemistry are strongly affected by the pervasive human influence on ammonia emissions and the nitrogen cycle in the Anthropocene.
Content may be subject to copyright.
ATMOSPHERIC AEROSOLS
Multiphase buffer theory explains contrasts
in atmospheric aerosol acidity
Guangjie Zheng
1
, Hang Su
2
*, Siwen Wang
2
, Meinrat O. Andreae
2,3,4
, Ulrich Pöschl
2
, Yafang Cheng
1
*
Aerosol acidity largely regulates the chemistry of atmospheric particles, and resolving the drivers
of aerosol pH is key to understanding their environmental effects. We find that an individual buffering
agent can adopt different buffer pH values in aerosols and that aerosol pH levels in populated
continental regions are widely buffered by the conjugate acid-base pair NH
4+
/NH
3
(ammonium/
ammonia). We propose a multiphase buffer theory to explain these large shifts of buffer pH, and we
show that aerosol water content and mass concentration play a more important role in determining
aerosol pH in ammonia-buffered regions than variations in particle chemical composition. Our results
imply that aerosol pH and atmospheric multiphase chemistry are strongly affected by the pervasive
human influence on ammonia emissions and the nitrogen cycle in the Anthropocene.
Aerosol acidity has attracted increasing
interest in atmospheric research because
it influences the thermodynamics of gas-
particle partitioning and the chemical
kinetics of the formation and transfor-
mation of air particulate matter (18). Under-
standing the temporal and spatial variations
of aerosol pH in the atmosphere is crucial
for accurate predictions of the properties of
atmosphericaerosolsandtheireffectsonhealth,
ecosystems, and climate (912). In marine envi-
ronments, the uptake of acidic gases like SO
2
,
H
2
SO
4
,andHNO
3
may rapidly consume the
alkalinity and reduce the pH of sea salt aerosols
(13,14). For continental air masses in the south-
eastern United States, Weber et al. (15)have
suggested that aerosol pH is buffered in the
range of ~0 to 2 because of the interaction of
aqueous (NH
4
)
2
SO
4
-NH
4
HSO
4
with gaseous
NH
3
. However, their later studies have at-
tributed the elevated pH levels in northern
China (~3 to 6) (8,1619) mainly to changes in
particle chemical compositionsi.e., a shift from
sulfate- to nitrate-dominated aerosols (12,19)
whereas Cheng et al. (8) have highlighted the
role of ammonia and alkaline aerosol compo-
nents from natural and anthropogenic emissions
in understanding aerosol pH in this region.
Despite these advances, it is still unclear
how aerosol pH is buffered in other continen-
tal regions, such as northern China, compared
with the southeastern United States. To answer
this question, we first performed numerical
model calculations with the state-of-the-art
thermodynamic model ISORROPIA (20)to
examine the response of pH in aerosols upon
the addition of sulfuric acid under different
conditions that are characteristic of the south-
eastern United States (15,21), the North China
Plain (8,22), northern India (23), and western
Europe (24)[tableS1and(25)]. For reference,
we also calculated the response of an aqueous
solution of Na
2
SO
4
. As shown in fig. S1, the
Na
2
SO
4
solution exhibits the expected steep
decrease of pH upon acid addition. For aerosol
systems, however, the pH does not show a
substantial decrease until the added amount
of acid (H
+
equivalent) reaches ~20% of the
initial amount of anions in the aqueous par-
ticles (molar ratio), which indicates an ad-
ditional buffering effect. To further investigate
this phenomenon, we focus on the scenarios
for the southeastern United States (SE-US)
and for the North China Plain (NCP), which
have been intensively investigated and dis-
cussed in earlier studies. As indicated in
table S1, the SE-US scenario is characterized
by relatively low aerosol concentration, low
aerosol water content (AWC), and high tem-
perature, as observed under clean-air summer
conditions in the southeastern United States.
By contrast, the NCP scenario is characterized
by the high aerosol concentration, high AWC,
and low temperature observed during extreme
winter haze events in the Beijing region.
In aqueous solutions, the pH of different
buffer systems is usually determined by the pK
a
(where K
a
is the acid dissociation constant)
of the buffering agents (26). Accordingly, the
different pH buffer levels in fig. S1 would sug-
gest different buffering agents corresponding
to different particle chemical compositions.
To identify the most relevant buffering agents,
key controlling parameters, we introduce the
concept of a multiphase buffering capacity
that describes the resistance to pH changes
upon input of acids or bases in an aerosol
multiphase system in analogy to the traditional
buffering capacity of bulk aqueous solutions.
The buffering capacity bis defined as the ratio
between the amount of acid or base added
to the system (n
acid
or n
base
,inmolesper
kilogram) and the corresponding pH change
in the aqueous phase of the system, or b=
dn
acid
/dpH = dn
base
/dpH. The larger the
buffering capacity b, the less the pH will change
upon the addition of acids or bases.
Figure 1A shows the buffering capacities for
the SE-US and NCP aerosol scenarios and for
bulk aqueous solutions of the individual buf-
fering agents (i.e., conjugate acid-base pairs
NH
4
+
/NH
3
, HSO
4
/SO
4
2
, and HNO
3
/NO
3
)
as derived from numerical simulations of the
gas-liquid and acid-base equilibria [see mate-
rials and methods, section M1; results of the
northern India and western Europe scenarios
are in fig. S2; and results of organic buffers are
in the supplementary text, section S7 (25)]. In
both aerosol scenarios, the largest buffering
capacity is obtained for the acid-base pair
NH
4
+
/NH
3
followed by HSO
4
/SO
4
2
and HNO
3
/
NO
3
. The peak buffer pH value (defined as the
pH corresponding to the highest local maxi-
mum of b) for the SE-US scenario is ~0.7, and
the peak buffer pH value for the NCP scenario
is ~4.5. Thus, the buffer pH ranges (i.e., peak
buffer pH ± 1) (26,27) closely match the aerosol
pH ranges previously reported for the southeast-
ern United States and for Beijing, respectively.
This indicates that the conjugate acid-base
pair NH
4
+
/NH
3
is the main buffering agent in
both the SE-US and NCP aerosol scenarios.
This finding raises the question of how the
same buffering agent can stabilize the aerosol
pH at verydifferent levels. As shown in Fig. 1A,
in bulk aqueous solution, the peak buffer pH
of NH
4
+
/NH
3
is ~9.2, but in the NCP and SE-US
aerosol scenarios, it shifts to much lower values
of ~4.5 and ~0.7, respectively. By contrast, the
peak buffer pH of the conjugate acid-base pair
HNO
3
/NO
3
shifts in the opposite direction
from ~1.5 in the bulk aqueous solution to
higher values of ~0.2 and ~3.8 in the NCP and
SE-US scenarios, respectively. The conjugate
acid-base pair HSO
4
/SO
4
2
, on the other hand,
exhibits similar peak buffer pH values of ~2
in all three scenarios (Fig. 1A). These differ-
ences and shifts of peak buffer pH reflect spe-
cial features of the aerosol multiphase buffer
system that go beyond the traditional buffer
theory for bulk aqueous solutions, and they
highlight the need for a mechanistic under-
standing of the multiphase buffering mech-
anism in atmospheric aerosols.
To elucidate the underlying mechanisms and
key parameters, we have developed a multi-
phase buffer theory and derived an analytical
expression for the buffering capacity of a buf-
fering agent X (conjugate acid-base pair) in an
aerosol multiphase buffer system as detailed
in section S1 (25)
b¼dnbase
dpH ¼2:303 Kw
½Hþþ½Hþþ
X
i
Ka;i½Hþ
ðKa;iþ½HþÞ2½Xitot
ð1Þ
RESEARCH
Zheng et al., Science 369, 13741377 (2020) 11 September 2020 1of4
1
Minerva Research Group, Max Planck Institute for
Chemistry, Mainz 55128, Germany.
2
Multiphase Chemistry
Department, Max Planck Institute for Chemistry, Mainz
55128, Germany.
3
Scripps Institution of Oceanography,
University of California, San Diego, La Jolla, CA 92093, USA.
4
Department of Geology and Geophysics, King Saud
University, 11451 Riyadh, Saudi Arabia.
*Corresponding author. Email: yafang.cheng@mpic.de (Y.C.);
h.su@mpic.de (H.S.)
on September 16, 2020 http://science.sciencemag.org/Downloaded from
Here, K
w
is the water dissociation constant,
[X
i
]
tot
* represents the total equivalent mol-
ality of the buffering agent X
i
, including the
gas phase and aqueous phase of both conju-
gate acid-base speciese.g., the sum of NH
3
(g),
NH
3
(aq), and NH
4
+
(aq) for the buffering agent
NH
4
+
/NH
3
.K
a,i
* is an effective acid dissociation
constant of the buffering agent X
i
and can be
expressed by
Ka;BOH ¼
½HþðaqÞ½BOHðaqÞ þ ½BOHðgÞ
½BþðaqÞ
¼Ka;BOH 1þrw
HiRT AWC

ð2AÞ
Ka;HA ¼½HþðaqÞ½AðaqÞ
½HAðaqÞþ½HA ðgÞ
¼Ka;HA =1þrw
HiRT AWC

ð2BÞ
for volatile base BOH and volatile acid HA that
dissociate in the form
BþðaqÞþH2OHþðaqÞþBOHðaqÞð2CÞ
HAðaqÞHþðaqÞþAðaqÞð2DÞ
As shown in Eq. 2, the effective dissociation
constant K
a,i
* depends on the classical disso-
ciation constant K
a,i
as well as on the Henrys
law coefficient H
i
(gas-particle partitioning
constant) (in moles per liter per atmosphere)
and on the AWC (in micrograms per cubic
meter)i.e., the amount of liquid water in
the aerosol multiphase system. Here, r
w
is
the liquid water density (~10
12
mgm
3
), Ris the
gas constant (8.205 ×10
2
atm L mol
1
K
1
), and
Tis the absolute temperature (in kelvin). Note
that gas concentrations in square brackets are
expressed in units of equivalent molality (in
moles per kilogram of water) [see section M1
(25)]. The expression of Eqs. 1 and 2 in the
other unit system can be found in section S2.
By solving Eq. 1, we can find a local maxi-
mum of bat pH = pK
a,i
*; i.e., the peak buffer
pH of the agent X
i
is determined by K
a,i
*.
Therefore, a single buffering agent can have
its peak buffering capacity at very different
pH values in an aerosol multiphase buffer sys-
tem. According to Eq. 2A, for the buffering
agent NH
4
+
/NH
3
(volatile base), increasing
AWC results in a reduced K
a
*andincreased
pK
a
*. Thus, the traditional alkaline buffering
agent NH
4
+
/NH
3
effectively becomes an acidic
buffering agent (pK
a
* < 7) in multiphase sys-
tems (Fig. 1A). For volatile acid buffering agents
(HNO
3
/NO
3
), the AWC has the opposite effect
on pK
a
* (Fig. 1A). Moreover, the shift of pK
a
*
upon changes to the AWC is inversely pro-
portional to the partitioning coefficient H
i
.
Thus, the volatile buffering agents HNO
3
/NO
3
and NH
4
+
/NH
3
(low H
i
)exhibitlargeshifts,
whereas the peak buffer pH of HSO
4
/SO
4
2
hardly changes (high H
i
) (table S2). As shown
in Fig. 1B, the effective dissociation constant
K
a
* converges with the standard acid-base
dissociation constant of the buffering agent
for high values of AWC (Eq. 2), and the multi-
phasebuffer theory converges with the con-
ventional buffer theory in solution chemistry
[see sections S1 to S4 (25)]. Note that ac-
tivity coefficients must be considered in the
calculation of nonideal systems [see section
S3 (25)].
Figure 2 further explains the thermodynam-
ics that causes the shift of buffer pH in a multi-
phase system. The conventional bulk buffer
solutions (e.g., NH
4
+
/NH
3
), assuming no ex-
change with a gas phase, achieve their largest
resistance to pH change when the molality of
NH
4
+
(aq) is equal to that of NH
3
(aq) (Eq. 3A) (26)
pH ¼pKa;NH3þlog10
½NH3ðaqÞ
½NH4þðaqÞ ð3AÞ
where
Ka;NH3¼Kw
Kb;NH3
¼½HþðaqÞ½NH3ðaq Þ
½NH4þðaqÞ ð3BÞ
and Kb;NH3is the base dissociation constant of
NH
3
(table S2).
For gas-liquid multiphase systems, this equi-
librium is extended to the gas phase, and pH
becomes a function of K
a
* and the ratio of total
NH
3
in both gas and aqueous phase to NH
4
+
in
the aqueous phase [Eq. 4A; section S1 (25)].
Accordingly, the largest resistance to pH change
under given [NH
3
]
tot
*([NH
3
]
tot
*=[NH
3
(aq)] +
[NH
3
(g)] + [NH
4
+
(aq)]) is achieved when the
molality of NH
4
+
(aq) is equal to the sum of
NH
3
(aq) and NH
3
(g). Note that Eq. 3 still holds
for the aqueous phase in the multiphase system
pH ¼pKa;NH3
þlog10
½NH3ðaqÞþ½NH3ðgÞ
½NH4þðaqÞ
ð4AÞ
where
Ka;NH3
¼
½HþðaqÞ½NH3ðaqÞþ½NH3ðgÞ
½NH4þðaqÞ
¼Ka;NH31þrw
HNH3RT AWC

ð4BÞ
Figure 2 shows the conditions where the
peak buffer pH values are achieved in different
systemsi.e., the same height of NH
4
+
and
NH
3
in each panel represents their same molar
numbers in each system. Compared with bulk
Zheng et al., Science 369, 13741377 (2020) 11 September 2020 2of4
Fig. 1. Buffering capacity
for aerosol multiphase
systems compared with
bulk aqueous solution.
(A) Buffer capacities (b) for
the SE-US and NCP aerosol
scenarios and for bulk
aqueous solution of indi-
vidual buffering agents
(solid lines). The overall
buffering capacity (black
dashed lines) is obtained
by adding the individual
buffer agent contributions
to the solvent background
of water [fig. S3 and sec-
tion S5 (25)]. The
composition of the bulk
solution is assumed to have
the same aqueous phase
molality as in the SE-US
scenario. (B) Dependence
of the peak buffer capacity
(pK
a
*) of NH
4+
/NH
3
on
aerosol water content
(AWC) and temperature.
RESEARCH |REPORT
on September 16, 2020 http://science.sciencemag.org/Downloaded from
solution, a fraction of NH
3
partitions to the gas
phase in the NCP scenario, which results in less
NH
3
(aq)andareduced[NH
3
(aq)]/[NH
4
+
(aq)]
ratio, which leads to a lower pH in the aqueous
phase according to Eq. 3. Further reduction of
AWC in the SE-US scenario will push more NH
3
to
the gas phase and further reduce the aerosol pH.
Figure 3 compares the contribution of in-
dividual factors in explaining the difference in
aerosol pH between the NCP (~5.4) and SE-US
(~0.7) scenarios in fig. S1 [sections M2, S3, S5,
and S6 (25)]. The AWC appears to be the most
important factor, contributing 2.2 units of pH
change (DpH), followed by T,whichcontributes
another 1.6 units of DpH. Although earlier
studies have hypothesized that the marked
observed pH difference is caused by a transi-
tion in particle chemical composition from a
sulfate- to a nitrate-dominated regime (12,19),
our results show that the change of chemical
composition only plays a minor role. Differ-
ent AWCs [mainly caused by different aerosol
concentrations at a given relative humidity
(RH)] and Tvalues can already explain a shift
of ~4 units of aerosol pH. The difference in
chemical composition contributes ~0.7 pH units
in total, with ~0.5 from the difference in total
NH
3
fraction and ~0.1 and ~0.1 from the dif-
ference in the fraction of NO
3
and nonvolatile
cations (NVCs), respectively. Overall, different
AWCs and Tvalues are the main drivers of the
pH difference between the NCP and SE-US
scenarios, whereas the higher fraction of total
NH
3
,NVCs,andNO
3
in the NCP further en-
larges the difference.
In Fig. 4, we performed global model sim-
ulations to identify the buffered regions and
used both simulation and observational data to
further compare the roles of AWC and chemical
compositions in determining the variabilities
of aerosol pH [sections M3, S3, and S6 and
table S3 (25)]. As shown in Fig. 4A, ~40% of
continental surface areas (not including Ant-
arctica) and 71% of urban populated areas
were buffered by the NH
4
+
/NH
3
agent with
aerosol pH values mostly within the buffer range
[pK
a,NH3
*±1(26,27)]. In these regions, without
knowing the temporal and spatial variability
of particle chemical composition, variations
in AWC alone explain almost 70% (R
2
= 0.66,
simulation; where R
2
is the coefficient of
determination) and 80% (R
2
=0.77,observa-
tion)ofthevariationofaerosolpH,assuming
an NH
4
+
/NH
3
buffered system (Fig. 4B). On
the other hand, when a constant AWC is as-
sumed, distinct variations of aerosol acidity with
particle chemical composition were observed,
but they only played a secondary role (R
2
=
0.22 and 0.26 for simulation and observation,
respectively; Fig. 4C). We also found a reverse
role for AWC and composition in regions that
are not buffered by NH
4
+
/NH
3
,wherechem-
ical composition differences alone explain >90%
of the variations of aerosol pH (fig. S5). Overall,
the buffering effect of ammonia suppresses the
influence of compositional differences, making
aerosol water content the primary determinant
of aerosol pH.
The multiphase buffering of aerosols and
the key role of AWC in determining the peak
buffer pH (pK
a
*) have implications for atmo-
spheric research and air pollution control.
Drivers of historical trends in aerosol pH can
now be better understood and quantified [sec-
tion S8 (25)]. In populated continental regions
with high anthropogenic emissions and atmo-
spheric concentrations of ammonia (28), aero-
sol pH is likely controlled by the buffering pair
NH
4
+
/NH
3
and can thus be approximated on
the basis of aerosol mass concentration and
Zheng et al., Science 369, 13741377 (2020) 11 September 2020 3of4
Fig. 2. Schematic diagram of buffer pH transition from aerosol multiphase systems to bulk aqueous
solution for NH
4+
/NH
3
.AWC concentrations are not drawn to scale, for illustration purposes. Diagrams for
a generic volatile acid and base are shown in fig. S4.
Fig. 3. Fractional contribution of individual drivers to the aerosol pH difference between SE-US and
NCP scenarios. Red and blue lines mark the corresponding values in the SE-US and NCP scenarios,
respectively [see table S1 for detailed scenario information (25)].
RESEARCH |REPORT
on September 16, 2020 http://science.sciencemag.org/Downloaded from
water content (Eq. 4). This opens up possibili-
ties to reconstruct long-term trends and large-
scale spatial distributions of aerosol pH. Other
buffering agents, such as HSO
4
/SO
4
2
,HCl/Cl
,
or HCO
3
/CO
3
2
,arelikelytocontrolaerosol
pH over the oceans (13,14,29,30), but the buf-
fering effects of NH
4
+
/NH
3
may extend over
ammonia-rich coastal and downwind regions.
Thus, the notable human influence on ammo-
nia emissions and the global nitrogen cycle in
the Anthropocene substantially affects aerosol
pH and atmospheric multiphase chemistry on
a global scale.
REFERENCES AND NOTES
1. J. H. Seinfeld, S. N. Pandis, Atmospheric Chemistry and
Physics: From Air Pollution to Climate Change (Wiley, 2016).
2. D. J. Jacob, J. W. Munger, J. M. Waldman, M. R. Hoffmann,
J. Geophys. Res. Atmos. 91, 10731088 (1986).
3. D. J. Jacob, J. M. Waldman, J. W. Munger, M. R. Hoffmann,
J. Geophys. Res. Atmos. 91, 10891096 (1986).
4. A. R. Ravishankara, Science 276, 10581065 (1997).
5. M. Jang, N. M. Czoschke, S. Lee, R. M. Kamens, Science 298,
814817 (2002).
6. S. Gao et al., Environ. Sci. Technol. 38, 65826589 (2004).
7. J. D. Surratt et al., Environ. Sci. Technol. 41,53635369 (2007).
8. Y. Cheng et al., Sci. Adv. 2, e1601530 (2016).
9. D. W. Dockery et al., Environ. Health Perspect. 104, 500505
(1996).
10. W. Li et al., Sci. Adv. 3, e1601749 (2017).
11. E. Harris et al., Science 340, 727730 (2013).
12. H. O. T. Pye et al., Atmos. Chem. Phys. 20, 48094888 (2020).
13. W. C. Keene, A. A. P. Pszenny, J. R. Maben, R. Sander, Geophys.
Res. Lett. 29, 1101 (2002).
14. B. Alexander et al., J. Geophys. Res. Atmos. 110, D10307
(2005).
15. R. J. Weber, H. Guo, A. G. Russell, A. Nenes, Nat. Geosci. 9,
282285 (2016).
16. M. Liu et al., Geophys. Res. Lett. 44, 52135221 (2017).
17. G. Shi et al., Environ. Sci. Technol. 51, 42894296 (2017).
18. S. Song et al., Atmos. Chem. Phys. 18, 74237438 (2018).
19. H. Guo, R. J. Weber, A. Nenes, Sci. Rep. 7, 12109 (2017).
20. C. Fountoukis, A. Nenes, Atmos. Chem. Phys. 7, 46394659
(2007).
21. H. Guo et al., Atmos. Chem. Phys. 15, 52115228 (2015).
22. G. J. Zh eng et al., Atmos. Chem. Phys. 15, 29692983
(2015).
23. S. N. Behera, M. Sharma, Sci. Total Environ. 408, 35693575
(2010).
24. L. Bencs et al., J. Environ. Monit. 10,11481157 (2008).
25. See supplementary materials.
26. P. Atkins, L. Jones, Chemical Principles: The Quest for Insight
(Macmillan, 2007).
27. H. N. Po, N. Senozan, J. Chem. Educ. 78, 1499 (2001).
28. M. Van Damme et al., Nature 564,99103 (2018).
29. W. C. Keene, D. L. Savoie, Geophys. Res. Lett. 25, 21812184
(1998).
30. D. J. Gaspar, A. Laskin, W. Wang, S. W. Hunt,
B. J. Finlayson-Pitts, Appl. Surf. Sci. 231-232, 520523 (2004).
ACKNO WLED GME NTS
Funding: This study is support by the Max Planck Society (MPG).
Y.C. thanks the Minerva Program of MPG. Author contributions:
Y.C. and H.S. conceived the theory and led the study. G.Z., Y.C.,
and H.S. performed the research. S.W. performed the GEOS-Chem
simulation. M.O.A. commented on the manuscript. H.S., Y.C., G.Z.,
and U.P. wrote the manuscript with inputs from all coauthors.
Competing interests: The authors declare no competing interests.
Data and materials availability: All data used in the analysis are
provided in the supplementary materials.
SUPPLEMENTARY MATERIALS
science.sciencemag.org/content/369/6509/1374/suppl/DC1
Materials and Methods
Supplementary Text
Figs. S1 to S15
Table S1 to S4
References (3197)
Data S1
29 November 2019; accepted 21 July 2020
10.1126/science.aba3719
Zheng et al., Science 369, 13741377 (2020) 11 September 2020 4of4
Fig. 4. Drivers of aerosol pH diversity in ammonia-buffered regions. (A) Global distribution of continental
surface regions buffered by NH
4+
/NH
3
. The color coding shows the maximum buffer capacity by NH
4+
/NH
3
(inmolespercubicmeterofair).(B) Correlation of aerosol pH modeled by ISORROPIA with the predicted pH
derived using constant buffering agent and multiphase buffer theory. Sim., simulation; Obs., observation.
(C) Correlation of aerosol pH modeled by ISORROPIA with the predicted pH by ISORROPIA using constant
AWC but variable compositions. Black circles and gray dots represent analysis based on model simulations and
observations, respectively (see section M3 and table S3). Note that the observations are based on individual case
studies and thus show a wider range of aerosol pH than the annual average simulation results.
RESEARCH |REPORT
on September 16, 2020 http://science.sciencemag.org/Downloaded from
Multiphase buffer theory explains contrasts in atmospheric aerosol acidity
Guangjie Zheng, Hang Su, Siwen Wang, Meinrat O. Andreae, Ulrich Pöschl and Yafang Cheng
DOI: 10.1126/science.aba3719
(6509), 1374-1377.369Science
, this issue p. 1374Science
important influence of ammonia emissions in the Anthropocene.
important role for water content in determining pH in ammonia-buffered regions. Their conclusions underscore the
considered how buffering capacity in a multiphase aerosol system differs from bulk solution and found anet al.Zheng
is their acidity, so understanding what determines aerosol pH is fundamental for determining their environmental effects.
Aerosols exert a primary influence on atmospheric chemistry. One of the main controls on their internal chemistry
A multiphasic effect
ARTICLE TOOLS http://science.sciencemag.org/content/369/6509/1374
MATERIALS
SUPPLEMENTARY http://science.sciencemag.org/content/suppl/2020/09/09/369.6509.1374.DC1
REFERENCES http://science.sciencemag.org/content/369/6509/1374#BIBL
This article cites 94 articles, 7 of which you can access for free
PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions
Terms of ServiceUse of this article is subject to the
is a registered trademark of AAAS.ScienceScience, 1200 New York Avenue NW, Washington, DC 20005. The title
(print ISSN 0036-8075; online ISSN 1095-9203) is published by the American Association for the Advancement ofScience
Science. No claim to original U.S. Government Works
Copyright © 2020 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of
on September 16, 2020 http://science.sciencemag.org/Downloaded from
... These results were generally consistent with the pH estimates of cloud water and PM 2.5 . The thermodynamic equilibrium of SO 4 2− , NO 3 − , and NH 4 + (SNA) plays a dominant role in aerosol acidity (Hennigan et al., 2015;Weber et al., 2016;Zheng et al., 2020). Thus, the SNA of particulate matter and that of cloud water in the same period were compared. ...
... Whiteface was more than four times that of NO 3 − (Khwaja et al., 1995). NH 4 + is the most abundant cation in alpine regions (Deguillaume et al., 2014;Herckes et al., 2002;Marinoni et al., 2004;Michna et al., 2015;Plessow et al., 2001;van Pinxteren et al., 2016), and the results from previous research have confirmed the key buffering effect of NH 4 + / NH 3 (ammonium/ammonia) on the acidity of atmospheric aerosols and clouds (Zheng et al., 2020). The abundance of NH 3 in the atmosphere also varies by one or two orders of magnitude in different regions of the world. ...
... Mechanistic studies have revealed that AWC was a primary contributor to pH shifts. Zheng et al. (2020) proposed a multiphase buffer theory suggesting that AWC could considerably regulate the peak buffer pH of the individual buffering agent (i.e., conjugate acid-base pairs NH4 + /NH3, HSO4 − /SO4 2− , and HNO3/NO3 − ). The distribution of AWC was characterized similarly to a skewed log-normal distribution, with noticeable differences between its arithmetic mean (53.3 μg m −3 ), median (6.8 μg m −3 ), and mode (0.5 μg m −3 ). ...
... As the RH escalated to around 50%, the SNA proportion experienced a rapid ascent, concurrently with a precipitous decline in the dust proportion, which better compatibly explained the decrease in aerosol pH. The relatively 180 stable pH w * variation at from 50% to 90% RH could be explained by the multiphase buffering theory (Zheng et al., 2020;Zheng et al., 2022). The theoretical equation derived from the multiphase buffering theory (see Text S2) suggested, when the aerosol pH was predominantly moderated by the buffering of the conjugate acid-base pair NH3/NH4 + , that aerosol pH could be simplified as a function of p NH 3 (partial pressure of gaseous NH3), NH 4 + aq (molality of NH4 + in aerosol water), and γ NH 4 + (aq) (activity coefficient of NH 4 + aq ). ...
Preprint
Full-text available
Aerosol pH is commonly used to characterize the acidity of aqueous aerosols and is of significant scientific interest due to its close relationship with atmospheric processes. Estimation of ambient aerosol pH usually relies on the thermodynamic modeling approach. In the existing chemical transport model and field observation studies, the temporal resolution of the input chemical and meteorological data into thermodynamic models varied substantially ranging from less than an hour to a year because of the inconsistency in the resolution of the original data and the aggregation of time-series data in some studies. Furthermore, the average value of aerosol pH has been represented by diverse metrics of central tendency in existing studies. This study attempts to evaluate the potential discrepancies in the calculated average aerosol pH arising from differences in both averaging metrics and temporal resolutions based on the ISORROPIA-II thermodynamic model and the example datasets prepared by the GEOS-Chem chemical transport model simulation. Overall, we find that the variation in the temporal resolutions of input data may lead to a change of up to more than two units in the average pH, and that the averaging metrics calculated based on the pH value of individual samples may be about two units higher than the averaging metrics calculated based on the activity of hydrogen ions. Accordingly, we recommend that the chosen averaging metrics and temporal resolutions should be stated clearly in future studies to ensure comparability of the average aerosol pH between models and/or observations.
... Campbell et al. (2021) investigated the associations between more-oxidized and less-oxidized oxygenated OA (MO-OOA and LO-OOA) with OP in Beijing but their contributions to OP were not resolved. On the other hand, several studies have suggested that the sources, formation mechanisms, and aging processes of OA and fine PM in the NCP are distinct from those in the United States and Europe Cheng et al., 2016;Xu et al., 2017;Zheng et al., 2020). For example, J. Wang et al. (2021) suggested that the aqueous production of SOA from fossil fuel emissions is a dominant mechanism for SOA formation in Beijing during winter haze episodes, while the photochemical processing is the major pathway for SOA formation in the United States and Europe (Wood et al., 2010;Q. ...
Article
Full-text available
The oxidative potential (OP) of ambient particulate matter (PM) is a common metric for estimating PM toxicity and linking PM exposure to adverse health effects. Organic aerosol (OA), a dominant fraction of ambient PM worldwide, may significantly contribute to PM toxicity. Here, we investigated the source‐based OA components contributing to the OP of PM in the urban (Beijing, summer and winter) and rural (Gucheng, winter) environments of the North China Plain (NCP). Various OA components as identified by the aerosol mass spectrometer/aerosol chemical speciation monitor (AMS/ACSM), transition metals, and black carbon were compared with the OP of PM measured by dithiothreitol assays. The results consistently demonstrate the importance of OA as a contributor to PM's OP in both urban and rural NCP environments. Higher intrinsic OP was observed in winter Beijing than in summer, possibly due to OA being predominantly from anthropogenic sources in winter. Furthermore, different OA components were found to drive the response of OP in the two environments. More‐oxidized oxygenated OA (MO‐OOA), cooking OA, and oxidized primary OA (during winter) are the OA contributors to OP in the urban environment, with a dominant contribution from MO‐OOA. In contrast, biomass burning OA (BBOA) and OOA play a major role in the OP in the rural environment, with BBOA making the largest contribution. Overall, this work highlights the significance of OA in determining PM's OP and calls for more work to reveal the sources and characteristics of OA components contributing to OP across different regions.
Article
Full-text available
Secondary inorganic aerosols play an important role in air pollution and climate change, and their formation modulates the atmospheric deposition of reactive nitrogen (including oxidized and reduced nitrogen), thus impacting the nitrogen cycle. Large-scale and long-term analyses of secondary inorganic aerosol formation based on model simulations have substantial uncertainties. Here we improve constraints on secondary inorganic aerosol formation using decade-long in situ observations of aerosol composition and gaseous precursors from multiple monitoring networks across the United States. We reveal a shift in the secondary inorganic aerosol formation regime in the rural United States between 2011 and 2020, making rural areas less sensitive to changes in ammonia concentrations and shortening the effective atmospheric lifetime of reduced forms of reactive nitrogen. This leads to potential increases in reactive nitrogen deposition near ammonia emission hotspots, with ecosystem impacts warranting further investigation. Ammonia (NH3), a critical but not directly regulated precursor of fine particulate matter in the United States, has been increasingly scrutinized to improve air quality. Our findings, however, show that controlling NH3 became significantly less effective for mitigating fine particulate matter in the rural United States. We highlight the need for more collocated aerosol and precursor observations for better characterization of secondary inorganic aerosols formation in urban areas.
Article
Full-text available
Adsorption and desorption of gases on liquid or solid substrates are involved in multiphase processes and heterogeneous chemical reactions. The desorption energy (Edes0), which depends on the intermolecular forces between adsorbate and substrate, determines the residence time of chemical species at interfaces. We show how Edes0 and temperature influence the net uptake or release of gas species, the rates of surface–bulk exchange and surface or bulk reactions, and the equilibration timescales of gas–particle partitioning. Using literature data, we derive a parameterization to estimate Edes0 for a wide range of chemical species based on the molecular mass, polarizability, and oxygen-to-carbon ratio of the desorbing species independent of substrate-specific properties, which is possible because of the dominant role of the desorbing species' properties. Correlations between Edes0 and the enthalpies of vaporization and solvation are rooted in molecular interactions. The relation between Edes0 and desorption kinetics reflects the key role of interfacial exchange in multiphase processes. For small molecules and semi-volatile organics (VOC, IVOC, SVOC), Edes0 values around 10–100 kJ mol−1 correspond to desorption lifetimes around nanoseconds to days at room temperature. Even higher values up to years are obtained at low temperatures and for low volatile organic compounds (LVOC, ELVOC/ULVOC) relevant for secondary organic aerosols (SOA). Implications are discussed for SOA formation, gas–particle partitioning, organic phase changes, and indoor surface chemistry. We expect these insights to advance the mechanistic and kinetic understanding of multiphase processes in atmospheric and environmental physical chemistry, aerosol science, materials science, and chemical engineering.
Article
Full-text available
The light absorbing component of organic aerosols, brown carbon (BrC), directly affects climate and can play a role in the oxidative aging of organic aerosols. Recent estimates suggest that globally BrC may have a warming potential that is approximately 20% that of black carbon, and photochemistry from BrC compounds can increase or transform aqueous SOA. Photobleaching of BrC is estimated to occur with a timescale of hours to days, a range that complicates estimates of the effects of BrC on climate and aerosol chemistry. The chemical environment (e.g. pH, ionic strength, and non-BrC organic content) of aqueous aerosols can also affect the reactivity of BrC, potentially altering absorption spectra and reactions of excited states formed upon irradiation. A range of solar illumination sources have been used in studying the photochemistry of BrC compounds, making direct comparisons between results difficult. Higher energy, single wavelength studies (e.g. 308 nm) show much larger quantum yields than broadband studies, indicating wavelength dependent quantum yields for a wide range of atmospherically relevant substituted aromatics. In this work we investigate the wavelength dependence of the quantum yield for loss of a prototypical BrC compound found in wildfire emissions, vanillin, using several narrow band UV-LEDs that span the 295–400 nm range. These wavelength dependent quantum yields will allow a more direct comparison between photochemical experiments with laboratory irradiation sources and actual actinic fluxes. Vanillin photochemical loss rates are concentration-dependent due to direct reaction between triplet excited state and ground state vanillin molecules. The quantum yield for photochemical loss of vanillin can be approximated by a Gaussian decay from 295 nm to ∼365 nm. This function is used to directly calculate the solar zenith angle (SZA) dependence for photochemical loss. Computational results show the presence of two π → π* transitions responsible for the observed UV-vis spectrum and that the rate of intersystem crossing has a wavelength dependence remarkably similar to that of the quantum yield for loss. A strong kinetic salt effect is observed, showing a doubling of the loss rate at high ionic strength.
Article
Full-text available
Acidity is an essential characteristic of aerosol particles that influences chemical and physical properties. Aerosol particles in tropical Asia, such as Singapore, have been reported as highly acidic due to intense anthropogenic sulfur emissions. At the same time, the region has also been experiencing wildfire over tropical peatlands, which is recognized as an intense source of NH3. Here, we investigated the role of NH3 from wildfire on aerosol particles in Singapore by employing the aerosol mass spectrometric technique. The observation was conducted both during wildfire haze (2015 October and 2019 September–October) and non‐haze (2018 October and 2019 April) periods. The observation result demonstrated that inorganic ionic species in Singapore were neutralized by NH4⁺ during the haze periods. Namely, the degree of neutralization of aerosol particles (i.e., measured NH4⁺ concentration/predicted NH4⁺ concentration by assuming that NH4⁺ fully neutralized SO4²⁻, NO3⁻, and Cl⁻) was lower than 0.77during the non‐haze periods. On the other hand, the corresponding values were higher than 0.93 during the haze periods. In addition, NO3⁻ concentration during the daytime of the haze period in 2015 was higher than that in other observation periods. A thermodynamic model calculation suggested that the regime shifts from the “NH3 sensitive region” to the “NH3 and HNO3 sensitive region” or “HNO3 sensitive region” might have occurred during the haze period. In the future, continuous monitoring of both gas‐ and particle‐phase inorganic chemical species will need to be conducted to investigate the impact of wildfire haze on atmospheric chemical processes in more detail.
Article
Full-text available
Wet processes, including aqueous-phase chemistry, wet scavenging, and wet surface uptake during dry deposition, are important for global modeling of aerosols and aerosol precursors. In this study, we improve the treatments of these wet processes in the Goddard Earth Observing System with chemistry (GEOS-Chem) v12.6.0, including pH calculations for cloud, rain, and wet surfaces, the fraction of cloud available for aqueous-phase chemistry, rainout efficiencies for various types of clouds, empirical washout by rain and snow, and wet surface uptake during dry deposition. We compare simulated surface mass concentrations of aerosols and aerosol precursors with surface monitoring networks over the United States, European, Asian, and Arctic regions, and show that model results with updated wet processes agree better with measurements for most species. With the implementation of these updates, normalized mean biases (NMBs) of surface nitric acid, nitrate, and ammonium are reduced from 78 %, 126 %, and 45 % to 0.9 %, 15 %, and 4.1 % over the US sites, from 107 %, 127 %, and 90 % to −0.7 %, 4.2 %, and 16 % over European sites, and from 121 %, 269 %, and 167 % to −21 %, 37 %, and 86 % over Asian remote region sites. Comparison with surface measured SO2, sulfate, and black carbon at four Arctic sites indicated that those species simulated with the updated wet processes match well with observations except for a large underestimate of black carbon at one of the sites. We also compare our model simulation with aircraft measurement of nitric acid and aerosols during the Atmospheric Tomography Mission (ATom)-1 and ATom-2 periods and found a significant improvement of modeling skill of nitric acid, sulfate, and ammonium in the Northern Hemisphere during wintertime. The NMBs of these species are reduced from 163 %, 78 %, and 217 % to −13 %, −1 %, and 10 %, respectively. The investigation of impacts of updated wet process treatments on surface mass concentrations indicated that the updated wet processes have strong impacts on the global means of nitric acid, sulfate, nitrate, and ammonium and relative small impacts on the global means of sulfur dioxide, dust, sea salt, black carbon, and organic carbon.
Article
Full-text available
Acidity, defined as pH, is a central component of aqueous chemistry. In the atmosphere, the acidity of condensed phases (aerosol particles, cloud water, and fog droplets) governs the phase partitioning of semivolatile gases such as HNO3, NH3, HCl, and organic acids and bases as well as chemical reaction rates. It has implications for the atmospheric lifetime of pollutants, deposition, and human health. Despite its fundamental role in atmospheric processes, only recently has this field seen a growth in the number of studies on particle acidity. Even with this growth, many fine-particle pH estimates must be based on thermodynamic model calculations since no operational techniques exist for direct measurements. Current information indicates acidic fine particles are ubiquitous, but observationally constrained pH estimates are limited in spatial and temporal coverage. Clouds and fogs are also generally acidic, but to a lesser degree than particles, and have a range of pH that is quite sensitive to anthropogenic emissions of sulfur and nitrogen oxides, as well as ambient ammonia. Historical measurements indicate that cloud and fog droplet pH has changed in recent decades in response to controls on anthropogenic emissions, while the limited trend data for aerosol particles indicate acidity may be relatively constant due to the semivolatile nature of the key acids and bases and buffering in particles. This paper reviews and synthesizes the current state of knowledge on the acidity of atmospheric condensed phases, specifically particles and cloud droplets. It includes recommendations for estimating acidity and pH, standard nomenclature, a synthesis of current pH estimates based on observations, and new model calculations on the local and global scale.
Article
Full-text available
Abstract Atmospheric NH3 plays a vital role not only in the environmental ecosystem but also in atmosphere chemistry. To further understand the effects of NH3 on the formation of haze pollution in Beijing, ambient NH3 and related species were measured and simulated at high resolutions during the wintertime Air Pollution and Human Health‐Beijing (APHH‐Beijing) campaign in 2016. We found that the total NHx (gaseous NH3+particle NH4+) was mostly in excess of the SO42−‐NO3−‐NH4+‐water equilibrium system during our campaign. This NHx excess made medium aerosol acidity, with the median pH value being 3.6 and 4.5 for polluted and nonpolluted conditions, respectively, and enhanced the formation of particle phase nitrate. Our analysis suggests that NH4NO3 is the most important factor driving the increasing of aerosol water content with NO3− controlling the prior pollution stage and NH4+ the most polluted stage. Increased formation of NH4NO3 under excess NHx, especially during the nighttime, may trigger the decreasing of aerosol deliquescence relative humidity even down to less than 50% and hence lead to hygroscopic growth even under RH conditions lower than 50% and the wet aerosol particles become better medium for rapid heterogeneous reactions. A further increase of RH promotes the positive feedback “aerosol water content‐heterogeneous reactions” and ultimately leads to the formation of severe haze. Modeling results by Nested Air Quality Prediction Monitor System (NAQPMS) show the control of 20% NH3 emission may affect 5–11% of particulate matter PM2.5 formation under current emissions conditions in the North China Plain.
Article
Full-text available
Aerosol pH is difficult to measure directly but can be calculated if the chemical composition is known with sufficient accuracy and precision to calculate the aerosol water content and the H+ concentration through the equilibrium among acids and their conjugate bases. In practical terms, simultaneous measurements of at least one semi-volatile constituent, e.g. NH3 or HNO3, are required to provide a constraint on the calculation of pH. Long-term records of aerosol pH are scarce due to the limited monitoring of NH3 in conjunction with PM2.5. In this study, 10-year (2007–2016) records of pH of PM2.5 at six eastern Canadian sites were calculated using the E-AIM II model with the input of gaseous NH3, gaseous HNO3 and major water-soluble inorganic ions in PM2.5 provided by Canada's National Air Pollution Surveillance (NAPS) Program. Clear seasonal cycles of aerosol pH were found with lower pH (∼2) in summer and higher pH (∼3) in winter consistently across all six sites, while the day-to-day variations of aerosol pH were higher in winter compared to summer. Tests of the sensitivity of aerosol pH to meteorological parameters demonstrate that the changes in ambient temperature largely drive the seasonal cycle of aerosol pH. The sensitivity of pH to chemical composition shows that pH has different responses to the changes in chemical composition in different seasons. During summertime, aerosol pH was mainly determined by temperature with limited impact from changes in NHx or sulfate concentrations. However, in wintertime, both meteorological parameters and chemical composition contribute to the variations in aerosol pH, resulting in the larger variation during wintertime. This study reveals that the sensitivity of aerosol pH to chemical composition is distinctly different under different meteorological conditions and needs to be carefully examined for any particular region.
Article
Full-text available
Aerosol pH is a useful diagnostic of aerosol chemistry for formation of secondary aerosol and has been hypothesized to be a key factor in specific chemical reaction routes producing sulfate and nitrate. In this study, we measured hourly concentrations of water‐soluble ions in particulate matter with an aerodynamic diameter less than 2.5 μm, along with gaseous pollutants in Tianjin, China, from 4 to 31 January 2015. The following source contributions to water‐soluble ions were estimated by positive matrix factorization: secondary sulfate (13%), secondary nitrate (44%), coal (14%), vehicle (16%), and dust (13%). ISORROPIA‐II was used to investigate the complex relationships among aerosol pH, ammonia, and secondary aerosol formation. The estimated hourly aerosol pH varied from −0.3 to 7.7, with an average of 4.9 (±0.78); the median value was 4.89, and the interquartile range was 0.72. During less polluted conditions, aerosol pH ranged from less than 0 to about 7; during heavily polluted conditions, pH was close to 5 (3.9–7.9) despite large amounts of sulfate. Sufficient ammonia/ammonium was present to balance high sulfate and nitrate formation. NH4⁺/NH3 (g) helped stabilize pH while nonvolatile cations contributed less to decreasing aerosol acidity. High acidy (pH < 3), light pollution (total water soluble ions < 30 μg/m³), and low water content (less than 5 μg/m³) were more correlated with higher rates of sulfate formation than nitrate formation in the winter.
Article
Full-text available
Aerosol acidity plays a key role in secondary aerosol formation. The high-temporal-resolution PM2.5 pH and size-resolved aerosol pH in Beijing were calculated with ISORROPIA II. In 2016–2017, the mean PM2.5 pH (at relative humidity (RH) > 30 %) over four seasons was 4.5±0.7 (winter) > 4.4±1.2 (spring) > 4.3±0.8 (autumn) > 3.8±1.2 (summer), showing moderate acidity. In coarse-mode aerosols, Ca²⁺ played an important role in aerosol pH. Under heavily polluted conditions, more secondary ions accumulated in the coarse mode, leading to the acidity of the coarse-mode aerosols shifting from neutral to weakly acidic. Sensitivity tests also demonstrated the significant contribution of crustal ions to PM2.5 pH. In the North China Plain (NCP), the common driving factors affecting PM2.5 pH variation in all four seasons were SO42-, TNH3 (total ammonium (gas + aerosol)), and temperature, while unique factors were Ca²⁺ in spring and RH in summer. The decreasing SO42- and increasing NO3- mass fractions in PM2.5 as well as excessive NH3 in the atmosphere in the NCP in recent years are the reasons why aerosol acidity in China is lower than that in Europe and the United States. The nonlinear relationship between PM2.5 pH and TNH3 indicated that although NH3 in the NCP was abundant, the PM2.5 pH was still acidic because of the thermodynamic equilibrium between NH4+ and NH3. To reduce nitrate by controlling ammonia, the amount of ammonia must be greatly reduced below excessive quantities.
Article
Full-text available
Through its important role in the formation of particulate matter, atmospheric ammonia affects air quality and has implications for human health and life expectancy1,2. Excess ammonia in the environment also contributes to the acidification and eutrophication of ecosystems3–5 and to climate change⁶. Anthropogenic emissions dominate natural ones and mostly originate from agricultural, domestic and industrial activities⁷. However, the total ammonia budget and the attribution of emissions to specific sources remain highly uncertain across different spatial scales7–9. Here we identify, categorize and quantify the world’s ammonia emission hotspots using a high-resolution map of atmospheric ammonia obtained from almost a decade of daily IASI satellite observations. We report 248 hotspots with diameters smaller than 50 kilometres, which we associate with either a single point source or a cluster of agricultural and industrial point sources—with the exception of one hotspot, which can be traced back to a natural source. The state-of-the-art EDGAR emission inventory¹⁰ mostly agrees with satellite-derived emission fluxes within a factor of three for larger regions. However, it does not adequately represent the majority of point sources that we identified and underestimates the emissions of two-thirds of them by at least one order of magnitude. Industrial emitters in particular are often found to be displaced or missing. Our results suggest that it is necessary to completely revisit the emission inventories of anthropogenic ammonia sources and to account for the rapid evolution of such sources over time. This will lead to better health and environmental impact assessments of atmospheric ammonia and the implementation of suitable nitrogen management strategies.
Article
Full-text available
The implementation of stringent emission regulations has resulted in the decline of anthropogenic pollutants including sulfur dioxide (SO2), nitrogen oxides (NOx), and carbon monoxide (CO). In contrast, ammonia (NH3) emissions are largely unregulated, with emissions projected to increase in the future. We present real-time aerosol and gas measurements from a field study conducted in an agriculturally intensive region in the southeastern US during the fall of 2016 to investigate how NH3 affects particle acidity and secondary organic aerosol (SOA) formation via the gas–particle partitioning of semi-volatile organic acids. Particle water and pH were determined using the ISORROPIA II thermodynamic model and validated by comparing predicted inorganic HNO3-NO3⁻ and NH3-NH4⁺ gas–particle partitioning ratios with measured values. Our results showed that despite the high NH3 concentrations (average 8.1±5.2ppb), PM1 was highly acidic with pH values ranging from 0.9 to 3.8, and an average pH of 2.2±0.6. PM1 pH varied by approximately 1.4 units diurnally. Formic and acetic acids were the most abundant gas-phase organic acids, and oxalate was the most abundant particle-phase water-soluble organic acid anion. Measured particle-phase water-soluble organic acids were on average 6% of the total non-refractory PM1 organic aerosol mass. The measured molar fraction of oxalic acid in the particle phase (i.e., particle-phase oxalic acid molar concentration divided by the total oxalic acid molar concentration) ranged between 47% and 90% for a PM1 pH of 1.2 to 3.4. The measured oxalic acid gas–particle partitioning ratios were in good agreement with their corresponding thermodynamic predictions, calculated based on oxalic acid's physicochemical properties, ambient temperature, particle water, and pH. In contrast, gas–particle partitioning ratios of formic and acetic acids were not well predicted for reasons currently unknown. For this study, higher NH3 concentrations relative to what has been measured in the region in previous studies had minor effects on PM1 organic acids and their influence on the overall organic aerosol and PM1 mass concentrations.
Article
The pH value is an important parameter of atmospheric aerosol. It affects the concentration of conjugate acid-base pair through the acid-base equilibrium and thus determines the gas–particle partitioning of acids or bases with volatility. Our recent report shows that there is a substitution of weak base for strong base in the aerosols of internally mixed water-soluble organic acid salt/ammonium sulfate. However, the acidity effect on the substitution process still remains ambiguous. In this work, the aerosols generated from sodium succinate/ammonium sulfate solutions with different pHs were studied in detail by using ATR-FTIR technique. The effects of relative humidity (RH) and acidity (pH) on the composition evolution, hygroscopic property and phase change were monitored. At a constant RH for a given pH, there were continuous depletions of NH4⁺, COO⁻ and water content accompanying occurrence of (CH2COOH)2 at initial stage, and then followed by Na2SO4 efflorescence, and at last participation of (CH2COOH)2. Lower RH was conductive to faster chemical composition evolution and resultant Na2SO4 crystallization. Higher pH promoted the composition evolution process and solid phase formation process. The consumptions of COO⁻ and NH4⁺ increased with increasing pH, showing that the dissolution of NH4⁺ to release H⁺ in aerosols and NH3 to gas phase led to water loss, in turn, Na2SO4 and succinic acid efflorescence. Water loss was more sensitive to Na2SO4 efflorescence than succinic acid. When a RH cycle was experienced, Efflorescence RHs of sodium succinate/ammonium sulfate aerosols almost kept unchanged and deliquscence RHs increased obviously with pH, while no deliquescence for pH 7.62. To our knowledge, it is the first time to investigate the pH effect on chemical process about composition evolution.
Article
In this work, we utilize a rich set of simulated and ground-based observational data in Tianjin, China, to examine and compare the differences in aerosol acidity and composition predicted by three popular thermodynamic equilibrium models: ISORROPIA II, the Extended Aerosol Inorganics Model vision IV (E-AIM IV), and the Aerosol Inorganic-Organic Mixtures Functional groups Activity Coefficients model (AIOMFAC). The species used to estimate aerosol acidity for both simulated and ambient data were NH4+, Na+, SO42−, NO3−, and Cl−. For simulated data, there is good agreement between ISORROPIA II and E-AIM IV predicted acidity in the forward and metastable mode, resulting from the hydrogen ion activity coefficient (γ_((H^+))) and the molality (m_((H^+))) showing opposite trends. While almost all other inorganic species concentrations are found to be similar among the three models, such is not the case for the bisulfate ion (HSO4-), which is linked to m_((H^+)). We find that differences in predicted bisulfate between the three models primarily result from differences in the treatment of the 〖HSO〗_4^- □(↔) H^++〖SO〗_4^(2-) reaction for highly acidic conditions. This difference in bisulfate is responsible for much of the difference in estimated pH for the ambient data (average pH of 3.5 for ISORROPIA II and 3.0 for E-AIM IV).