Intermolecular Hydrogen-Bonded Interactions of Oxalic Acid Conformers with Sulfuric Acid and AmmoniaClick to copy article linkArticle link copied!
- Eduardo da Silva CarvalhoEduardo da Silva CarvalhoDepartment of Materials Physics, Federal University of Amazonas, Manaus, AM 69080-900, BrazilMore by Eduardo da Silva Carvalho
- Angsula GhoshAngsula GhoshDepartment of Materials Physics, Federal University of Amazonas, Manaus, AM 69080-900, BrazilMore by Angsula Ghosh
- Puspitapallab Chaudhuri*Puspitapallab Chaudhuri*E-mail: [email protected]Department of Materials Physics, Federal University of Amazonas, Manaus, AM 69080-900, BrazilMore by Puspitapallab Chaudhuri
Abstract
Oxalic acid is one of the simplest naturally occurring dicarboxylic acids that is abundantly found in the atmosphere, and it has several stable structural conformers. Hydrogen-bonded interactions of oxalic acid with other atmospheric molecules are important, as they might influence the chemical composition of the atmosphere, thereby impacting atmospheric chemistry and environmental processes. In this work, we used density functional calculations with the M06–2X/6-311++G(3df,3pd) model to examine the interaction of five oxalic acid conformers with sulfuric acid and ammonia─two widely recognized atmospheric nucleation precursor molecules─with the aim of observing the hydrogen-bonding characteristics of the conformers individually. An extensive and systematic quantum-chemical calculation has been conducted to analyze the structural, thermodynamical, electrical, and spectroscopic characteristics of several binary and ternary clusters mediated by five oxalic acid conformers. Our analysis of the electronic-binding energies and free energy changes associated with the formation of the clusters at ambient temperature reveals that multiple conformations of oxalic acid have the potential to engage in stable cluster formation in the atmosphere. In fact, the highest energy oxalic acid conformer exhibits the lowest bonding free energy in most cases. According to our calculations, clusters of oxalic acid with sulfuric acid demonstrate greater thermodynamic stability, a higher probability of formation, and more intense light scattering compared to clusters with ammonia. Furthermore, the analysis of successive cluster formation reveals that clusters formed between sulfuric and oxalic acids are more likely to grow spontaneously than those formed between ammonia and oxalic acid.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
1. Introduction
2. Computational Methods
3. Results and Discussion
3.1. Isolated Molecule of Oxalic Acid
cTc | cTt | tTt | tCt | cCt | |
---|---|---|---|---|---|
μ (D) | 0.00 | 3.15 | 0.00 | 2.98 | 4.89 |
α (a.u.) | 37.36 | 37.84 | 38.13 | 38.16 | 37.99 |
Δα (a.u.) | 20.37 | 19.72 | 19.53 | 19.60 | 19.72 |
A (GHz) | 5.878 | 6.027 | 6.149 | 6.120 | 6.000 |
B (GHz) | 3.855 | 3.708 | 3.611 | 3.619 | 3.672 |
C (GHz) | 2.328 | 2.296 | 2.275 | 2.274 | 2.278 |
σ (a.u.) | 0.044 | 0.041 | 0.039 | 0.039 | 0.040 |
(a.u.) | 68188.7 | 67140.3 | 68103.4 | 68198.4 | 69983.7 |
Experimental values of the rotational constants (GHz): A = 5.951, B = 3.684, C = 2.276. (61).
3.2. Clusters of Oxalic Acid and Ammonia, (OA)(AM)n (n= 1,2)
RO–N (Å) | R(O)H···N (Å) | ∠O─H···N (degrees) | ΔRO–H (Å) | vO–H (cm–1) | ΔvO–H (cm–1) | ||
---|---|---|---|---|---|---|---|
(OA)(AM) | |||||||
(cTc)(AM) | 2.638 | 1.619 | 171.1 | 0.054 | 2692 | –1035 | |
(cTt)(AM)-1 | 2.659 | 1.649 | 170.7 | 0.052 | 2786 | –1034 | |
(cTt)(AM)-2 | 2.685 | 1.690 | 167.4 | 0.040 | 2974 | –799 | |
(tTt)(AM) | 2.681 | 1.682 | 168.6 | 0.045 | 2904 | –928 | |
(tCt))(AM) | 2.678 | 1.678 | 168.2 | 0.046 | 2917 | –904 | |
(cCt)(AM)-1 | 2.641 | 1.632 | 168.4 | 0.056 | 2753 | –1071 | |
(cCt)(AM)-2 | 2.696 | 1.719 | 164.1 | 0.037 | 3063 | –767 | |
(OA)(AM)2 | |||||||
(cTc)(AM)2-1 | 2.678 | 1.678 | 168.2 | 0.042 | 2894b | –833 | |
(cTc)(AM)2-2 | 2.562 | 1.503 | 172.1 | 0.094 | 2121 | –1607 | |
(cTt)(AM)2-1 | C | 2.722 | 1.740 | 166.4 | 0.034 | 3101 | –672 |
T | 2.692 | 1.693 | 170.3 | 0.041 | 2977 | –843 | |
(cTt)(AM)2-2 | 2.596 | 1.545 | 177.7 | 0.084 | 2287 | –1532 | |
(tTt)(AM)2-1 | 2.704 | 1.714 | 167.5 | 0.039 | 2997c | –835 | |
(tTt)(AM)2-2 | 2.623 | 1.586 | 178.2 | 0.071 | 2495 | –1337 | |
(tCt)(AM)2-1 | 2.701 | 1.710 | 167.3 | 0.040 | 2985d | –836 | |
(tCt)(AM)2-2 | 2.622 | 1.567 | 178.7 | 0.072 | 2478 | –1343 | |
(cCt)(AM)2-1 | 2.584 | 1.526 | 179.8 | 0.092 | 2167 | –1657 | |
(cCt)(AM)2-2 | C | 2.727 | 1.763 | 161.8 | 0.030 | 3173 | –657 |
T | 2.686 | 1.690 | 167.7 | 0.044 | 2921 | –903 |
The labels “C” and “T” have been indicated in Figure 4.
Asymmetric stretching mode of the two O–H groups with intensity of 4171 km/mol. The calculated value of the symmetric stretching mode of the same bonds is 2892 cm–1 with an intensity of just 5.2 km/mol.
Asymmetric stretching mode of the two O–H groups with intensity of 3785 km/mol. The calculated value of the symmetric stretching mode of the same bonds is 3023 cm–1 with a negligible intensity.
Asymmetric stretching mode of the two O–H groups with intensity of 3831 km/mol. The calculated value of the symmetric stretching mode of the same bonds is 3023 cm–1 with an intensity of just 6.6 km/mol.
ΔEB | ED(OA) | ED(tot) | ΔG | RPF | ΔGMC | Keq | |
---|---|---|---|---|---|---|---|
(OA)(AM) | |||||||
(cTc)(AM) | –11.89 | 4.32 | 4.39 | –3.65 | 100.00 | –3.65 | 4.8 × 102 |
(cTt)(AM)-1 | –12.76 | 3.16 | 3.22 | –4.51 | 99.79 | –4.51 | 2.1 × 103 |
(cTt)(AM)-2 | –9.73 | 1.84 | 1.88 | –0.87 | 0.21 | ||
(tTt)(AM) | –11.72 | 1.48 | 1.53 | –3.39 | 100.00 | –3.39 | 3.1 × 102 |
(tCt)(AM) | –11.81 | 1.50 | 1.55 | –2.71 | 100.00 | –2.71 | 9.8 × 101 |
(cCt)(AM)-1 | –13.57 | 2.06 | 2.11 | –5.03 | 99.68 | –5.03 | 4.9 × 103 |
(cCt)(AM)-2 | –8.90 | 2.60 | 2.63 | –1.63 | 0.32 | ||
(OA)(AM)2 | |||||||
(cTc)(AM)2-1 | –21.33 | 5.91 | 6.05 | –5.86 | 99.92 | –5.86 | 2.0 × 104 |
(cTc)(AM)2-2 | –18.13 | 8.00 | 8.18 | –1.66 | 0.08 | ||
(cTt)(AM)2-1 | –20.26 | 3.53 | 3.65 | –3.68 | 89.15 | –3.75 | 5.7 × 102 |
(cTt)(AM)2-2 | –19.37 | 4.65 | 4.84 | –2.43 | 10.85 | ||
(tTt)(AM)2-1 | –22.42 | 2.19 | 2.31 | –5.75 | 99.92 | –5.75 | 1.7 × 104 |
(tTt)(AM)2-2 | –18.33 | 3.59 | 3.76 | –1.52 | 0.08 | ||
(tCt)(AM)2-1 | –22.70 | 2.34 | 2.46 | –5.43 | 99.88 | –5.43 | 9.7 × 103 |
(tCt)(AM)2-2 | –18.74 | 3.56 | 3.73 | –1.44 | 0.12 | ||
(cCt)(AM)2-1 | –21.21 | 5.12 | 5.20 | –4.51 | 48.04 | –4.95 | 4.3 × 103 |
(cCt)(AM)2-2 | –20.83 | 3.30 | 3.40 | –4.56 | 51.96 |
Final channel (ternary cluster) | Initial channel (binary cluster + AM) | ΔGS (kcal/mol) |
---|---|---|
(cTc)(AM)2-1 | (cTc)(AM) + AM | –2.21 |
(cTc)(AM)2-2 | 1.99 | |
(cTt)(AM)2-1 | (cTt)(AM)-1 + AM | –2.81 |
(cTt)(AM)-2 + AM | 0.83 | |
(cTt)(AM)2-2 | (cTt)(AM)-1 + AM | –1.56 |
(cTt)(AM)-2 + AM | 2.08 | |
(tTt)(AM)2-1 | (tTt)(AM) + AM | –2.36 |
(tTt)(AM)2-2 | 1.87 | |
(tCt)(AM)2-1 | (tCt)(AM) + AM | –2.72 |
(tCt)(AM)2-2 | 1.27 | |
(cCt)(AM)2-1 | (cCt)(AM)-1 + AM | 0.52 |
(cCt)(AM)-2 + AM | –2.88 | |
(cCt)(AM)2-2 | (cCt)(AM)-1 + AM | 0.47 |
(cCt)(AM)-2 + AM | –2.93 |
3.3. Clusters of Oxalic Acid and Sulfuric Acid, (OA)(SA)n (n = 1,2)
ΔEB | ED(OA) | ED(tot) | ΔG | RPF | ΔGMC | Keq | |
---|---|---|---|---|---|---|---|
(cTc)(SA)-1 | –13.69 | 1.76 | 2.70 | –3.39 | 99.10 | ||
(cTc)(SA)-2 | –11.54 | 1.09 | 1.96 | –0.58 | 0.88 | –3.39 | 3.1 × 102 |
(cTc)(SA)-3 | –7.91 | 0.34 | 0.62 | 1.67 | 0.02 | ||
(cTt)(SA)-1 | –16.10 | 1.39 | 3.42 | –5.30 | 99.70 | ||
(cTt)(SA)-2 | –11.50 | 1.62 | 2.27 | –1.84 | 0.29 | –5.30 | 7.8 × 103 |
(cTt)(SA)-3 | –11.21 | 1.55 | 3.99 | 0.24 | 0.01 | ||
(tTt)(SA)-1 | –17.28 | 1.48 | 3.94 | –6.69 | 99.99 | ||
(tTt)(SA)-2 | –12.97 | 0.66 | 1.61 | –1.30 | 0.01 | –6.69 | 8.1 × 104 |
(tTt)(SA)-3 | –10.61 | 1.05 | 1.98 | –0.26 | 0.00 | ||
(tCt)(SA)-1 | –17.61 | 1.53 | 4.13 | –6.26 | 99.99 | ||
(tCt)(SA)-2 | –12.83 | 0.49 | 1.17 | –0.41 | 0.01 | –6.26 | 3.9 × 104 |
(tCt)(SA)-3 | –9.58 | 0.61 | 0.95 | 2.41 | 0.00 | ||
(cCt)(SA)-1 | –18.03 | 1.74 | 4.40 | –7.04 | 100.00 | ||
(cCt)(SA)-2 | –9.68 | 0.35 | 0.81 | 0.13 | 0.00 | –7.04 | 1.5 × 105 |
(cCt)(SA)-3 | –7.76 | 4.25 | 5.07 | 2.95 | 0.00 |
ΔEB | ED(OA) | ED(T) | ΔG | RPF | ΔGMC | Keq | |
---|---|---|---|---|---|---|---|
(cTc)(SA)2-1 | –32.84 | 1.92 | 12.57 | –9.65 | 96.61 | ||
(cTc)(SA)2-2 | –28.32 | 2.95 | 5.02 | –7.67 | 3.38 | ||
(cTc)(SA)2-3 | –27.62 | 2.19 | 7.23 | –4.20 | 0.01 | –9.67 | 1.3 × 107 |
(cTc)(SA)2-4 | –25.46 | 2.76 | 4.79 | –2.92 | 0.00 | ||
(cTc)(SA)2-5 | –22.42 | 2.00 | 7.45 | –0.32 | 0.00 | ||
(cTt)(SA)2-1 | –30.38 | 1.98 | 5.12 | –7.01 | 60.15 | ||
(cTt)(SA)2-2 | –30.15 | 1.91 | 8.58 | –6.66 | 33.10 | ||
(cTt)(SA)2-3 | –28.78 | 1.71 | 6.41 | –5.26 | 3.15 | –7.31 | 2.3 × 105 |
(cTt)(SA)2-4 | –29.38 | 10.39 | 23.55 | –5.28 | 3.23 | ||
(cTt)(SA)2-5 | –25.01 | 1.64 | 4.01 | –4.00 | 0.37 | ||
(tTt)(SA)2-1 | –34.25 | 2.96 | 7.56 | –11.82 | 99.43 | ||
(tTt)(SA)2-2 | –32.15 | 3.28 | 11.42 | –7.70 | 0.10 | ||
(tTt)(SA)2-3 | –31.90 | 2.17 | 9.81 | –8.56 | 0.40 | –11.83 | 4.8 × 108 |
(tTt)(SA)2-4 | –30.68 | 1.89 | 5.09 | –7.11 | 0.03 | ||
(tTt)(SA)2-5 | –30.63 | 1.25 | 10.04 | –7.19 | 0.04 | ||
(tCt)(SA)2-1 | –35.76 | 3.29 | 13.57 | –10.08 | 6.53 | ||
(tCt)(SA)2-2 | –34.71 | 2.90 | 7.59 | –11.65 | 92.27 | ||
(tCt)(SA)2-3 | –33.36 | 1.63 | 17.43 | –8.91 | 0.91 | –11.70 | 3.9 × 108 |
(tCt)(SA)2-4 | –32.54 | 2.26 | 10.05 | –8.25 | 0.30 | ||
(tCt)(SA)2-5 | –29.83 | 1.71 | 7.28 | –5.22 | 0.00 | ||
(cCt)(SA)2-1 | –37.70 | 4.36 | 13.19 | –13.31 | 99.41 | ||
(cCt)(SA)2-2 | –33.70 | 1.81 | 17.50 | –10.16 | 0.49 | ||
(cCt)(SA)2-3 | –32.98 | 2.40 | 9.77 | –9.24 | 0.10 | –13.31 | 5.9 × 109 |
(cCt)(SA)2-4 | –29.80 | 2.92 | 8.67 | –5.83 | 0.00 | ||
(cCt)(SA)2-5 | –28.38 | 5.78 | 11.18 | –4.57 | 0.00 |
3.4. Atmospheric Relevance of the Binding Free Energies
AM-containing cluster | %PF | [C] | SA-containing cluster | %PF | [C] |
---|---|---|---|---|---|
(cTc)(AM) | 4.77 × 10–5 | 2.39 × 105 | (cTc)(SA) | 6.24 × 10–8 | 3.12 × 102 |
(cTt)(AM) | 2.04 × 10–4 | 1.02 × 106 | (cTt)(SA) | 1.57 × 10–6 | 7.87 × 103 |
(tTt)(AM) | 3.08 × 10–5 | 1.54 × 105 | (tTt)(SA) | 1.65 × 10–5 | 8.25 × 104 |
(tCt)(AM) | 9.75 × 10–6 | 4.87 × 104 | (tCt)(SA) | 7.98 × 10–6 | 3.99 × 104 |
(cCt)(AM) | 4.92 × 10–4 | 2.46 × 106 | (cCt)(SA) | 2.98 × 10–5 | 1.49 × 105 |
(cTc)(AM)2 | 2.00 × 10–12 | 9.99 × 10–3 | (cTc)(SA)2 | 5.15 × 10–15 | 2.58 × 10–5 |
(cTt)(AM)2 | 5.85 × 10–14 | 2.83 × 10–4 | (cTt)(SA)2 | 9.55 × 10–17 | 4.77 × 10–7 |
(tTt)(AM)2 | 1.66 × 10–12 | 8.30 × 10–3 | (tTt)(SA)2 | 1.98 × 10–13 | 9.91 × 10–4 |
(tCt)(AM)2 | 9.66 × 10–13 | 4.83 × 10–3 | (tCt)(SA)2 | 1.59 × 10–13 | 7.96 × 10–4 |
(cCt)(AM)2 | 4.29 × 10–13 | 2.15 × 10–3 | (cCt)(SA)2 | 2.42 × 10–12 | 1.21 × 10–2 |
3.5. Interaction with Solar Radiation
4. Conclusions
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.4c06290.
Containing the optimized bond lengths and bond angles of different oxalic acid conformers; the Cartesian coordinates of the optimized geometries of binary and ternary clusters of OA with AM and SA; relevant structural parameters related to hydrogen bond formation in (OA)(SA) and (OA)(SA)2 clusters, supported by the ternary cluster images (Figure S1); successive binding free energies (ΔGS) for the formation of various (OA)(SA)2 ternary clusters and Boltzman-averaged values of Rayleigh scattering intensities of the OA conformers and their binary and ternatry clusters with AM and SA (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
The authors acknowledge financial support from the Brazilian funding agencies Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES)─finance code 001, Fundação de Amparo à Pesquisa do Estado do Amazonas (FAPEAM), and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq).
References
This article references 84 other publications.
- 1Huang, X.-F.; Hu, M.; He, L.-Y.; Tang, X.-Y. Chemical characterization of water-soluble organic acids in PM in Beijing, China. Atmos. Environ. 2005, 39, 2819– 2827, DOI: 10.1016/j.atmosenv.2004.08.038Google ScholarThere is no corresponding record for this reference.
- 2Wagner, R.; Möhler, O.; Saathoff, H.; Schnaiter, M.; Leisner, T. High variability of the heterogeneous ice nucleation potential of oxalic acid dihydrate and sodium oxalate. Atmos. Chem. Phys. 2010, 10, 7617– 7641, DOI: 10.5194/acp-10-7617-2010Google ScholarThere is no corresponding record for this reference.
- 3Xu, Y.; Nadykto, A. B.; Yu, F.; Jiang, L.; Wang, W. Formation and properties of hydrogen-bonded complexes of common organic oxalic acid with atmospheric nucleation precursors. J. Mol. Struct.: THEOCHEM 2010, 951, 28– 33, DOI: 10.1016/j.theochem.2010.04.004Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXmtlKlt7o%253D&md5=76c544813ec5a3150586367e7398b9c1Formation and properties of hydrogen-bonded complexes of common organic oxalic acid with atmospheric nucleation precursorsXu, Yisheng; Nadykto, Alexey B.; Yu, Fangqun; Jiang, Lei; Wang, WeiJournal of Molecular Structure: THEOCHEM (2010), 951 (1-3), 28-33CODEN: THEODJ; ISSN:0166-1280. (Elsevier B.V.)Nucleation in the Earth's atm. plays a vital role in the aerosol radiative forcing assocd. directly with the prodn. of ultrafine particles responsible for adverse public health impacts and global climate changes. The interaction of the common org. oxalic acid with the atm. nucleation precursors and trace ionic species has been investigated using the D. Functional Theory (DFT). A comprehensive study of the hydrogen-bonded complexes of oxalic acid with neutral monomers and ionic clusters of the sulfuric acid, water and ammonia, the key atm. nucleation precursors, has been carried out. Bonding of the oxalic acid with the above-mentioned neutrals was found to be moderately weak, and thus, the oxalic acid cannot stabilize the binary sulfuric acid-water and ternary sulfuric acid-ammonia clusters. The interaction of the oxalic acid with pos. ionic species is, in contrast, strong and significantly enhances stability of the clusters formed over the ions. The hydration of pos. charged (C2H2O4)(H3O+), (C2H2O4)(NH4+) and (C2H2O4)(H3O+)(H2SO4) clusters is also strong and close to that of the sulfuric acid, the key atm. nucleation precursor. These considerations lead to the logical conclusion that oxalic acid can catalyze the prodn. of pos. charged pre-nucleation clusters and, thus, its role in nucleation of pos. ions in the Earth's troposphere should be studied in further details.
- 4Xu, W.; Zhang, R. Theoretical investigation of interaction of dicarboxylic acids with common aerosol nucleation precursors. J. Phys. Chem. A 2012, 116, 4539– 4550, DOI: 10.1021/jp301964uGoogle Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XkvFOnu7c%253D&md5=f83fc23ed930da24fdb0fd0719292623Theoretical Investigation of Interaction of Dicarboxylic Acids with Common Aerosol Nucleation PrecursorsXu, Wen; Zhang, RenyiJournal of Physical Chemistry A (2012), 116 (18), 4539-4550CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Dicarboxylic acids are important products from photooxidn. of volatile org. compds. and are believed to play an important role in the formation and growth of atm. secondary org. aerosols. In this paper, the interaction of five dicarboxylic acids, i.e., oxalic acid (C2H2O4), malonic acid (C3H4O4), maleic acid (C4H4O4), phthalic acid (C8H6O4), and succinic acid (C4H6O4), with sulfuric acid and ammonia has been studied, employing quantum chem. calcns., quantum theory of atoms in mols. (QTAIM), and the natural bond orbital (NBO) anal. methods. Several levels of quantum chem. calcns. are considered, including coupled-cluster theory with single and double excitations with perturbative corrections for the triple excitations (CCSD(T)) and two d. functionals, B3LYP and PW91PW91. The free energies of formation of the heterodimer and heterotrimer clusters suggest that dicarboxylic acids can contribute to the aerosol nucleation process by binding to sulfuric acid and ammonia. In particular, the formation energies and structures of the heterotrimer clusters show that dicarboxylic acids enhance nucleation in two directions, in contrast to monocarboxylic acids.
- 5Weber, K. H.; Morales, F. J.; Tao, F. Theoretical study on the structure and stabilities of molecular clusters of oxalic acid with water. J. Phys. Chem. A 2012, 116, 11601– 11617, DOI: 10.1021/jp308499fGoogle Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsFCrurrJ&md5=5f1a7fac54947abdbade8bb7b9ba4b5aTheoretical Study on the Structure and Stabilities of Molecular Clusters of Oxalic Acid with WaterWeber, Kevin H.; Morales, Francisco J.; Tao, Fu-MingJournal of Physical Chemistry A (2012), 116 (47), 11601-11617CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The importance of aerosols to humankind is well-known, playing an integral role in detg. Earth's climate and influencing human health. Despite this fact, much remains unknown about the initial events of nucleation. In this work, the mol. properties of common org. atm. pollutant oxalic acid and its gas phase interactions with water have been thoroughly examd. Local min. single-point energies for the monomer conformations were calcd. at the B3LYP and MP2 level of theory with both 6-311++G(d,p) and aug-cc-pVDZ basis sets and are compared with previous works. Optimized geometries, relative energies, and free energy changes for the stable clusters of oxalic acid conformers with up to six waters were then obtained from B3LYP calcns. with 6-31+G(d) and 6-311++G(d,p) basis sets. Initially, cooperative binding is predicted to be the most important factor in nucleation, but as the clusters grow, dipole cancellations are found to play a pivotal role. The clusters of oxalic acid hydrated purely with water tend to produce extremely stable and neutral core systems. Free energies of formation and atm. implications are discussed.
- 6Weber, K. H.; Liu, Q.; Tao, F.-M. Theoretical study on stable small clusters of oxalic acid with ammonia and water. J. Phys. Chem. A 2014, 118, 1451– 1468, DOI: 10.1021/jp4128226Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXht12lsrs%253D&md5=d06cb1431aa9546ee0ce4c56d193d850Theoretical Study on Stable Small Clusters of Oxalic Acid with Ammonia and WaterWeber, Kevin H.; Liu, Qian; Tao, Fu-MingJournal of Physical Chemistry A (2014), 118 (8), 1451-1468CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Thermodynamically stable small clusters of oxalic acid (CO2H)2, ammonia (NH3), and water (H2O) are studied through quantum chem. calcns. The (CO2H)2-NH3 core system with up to three waters of hydration was examd. by B3LYP d. functional theory and MP2 MO theory with the aug-cc-pVDZ basis set. The (CO2H)2-NH3 core complexes are obsd. to hydrogen bond strongly and should be found in appreciably significant concns. in the atm. Subsequent hydration of the (CO2H)2-NH3 core, however, is found to be somewhat prohibitive under ambient conditions. Relative populations of the examd. clusters are predicted and the binding patterns detailed. Atm. implications related to new particle formations are discussed.
- 7Peng, X.-Q.; Liu, Y.-R.; Huang, T.; Jiang, S.; Huang, W. Interaction of gas phase oxalic acid with ammonia and its atmospheric implications. Phys. Chem. Chem. Phys. 2015, 17, 9552– 9563, DOI: 10.1039/C5CP00027KGoogle Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXjtlWgtbc%253D&md5=eb6efa7d64ab2c0d30464a57cae4b0a6Interaction of gas phase oxalic acid with ammonia and its atmospheric implicationsPeng, Xiu-Qiu; Liu, Yi-Rong; Huang, Teng; Jiang, Shuai; Huang, WeiPhysical Chemistry Chemical Physics (2015), 17 (14), 9552-9563CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Oxalic acid is believed to play an important role in the formation and growth of atm. org. aerosols. However, as a common org. acid, the understanding of the larger clusters formed by gas phase oxalic acid with multiple ammonia mols. is incomplete. In this work, the structural characteristics and thermodn. of oxalic acid clusters with up to six ammonia mols. have been investigated at the PW91PW91/6-311++G(3df,3pd) level of theory. We found that oxalic acid forms relatively stable clusters with ammonia mols., and that ionization events play a key role. The analyses of the thermodn. and atm. relevance indicate that the heterodimer (H2C2O4)(NH3) shows an obvious relative concn. in the atm., and thus likely participates in new particle formation. However, with increasing no. of ammonia mols., the concn. of clusters decreases gradually. Addnl., clusters of oxalic acid with ammonia mols. are predicted to form favorably in low temp. conditions and show high Rayleigh scattering intensities.
- 8Miao, S.-K.; Jiang, S.; Chen, J.; Ma, Y.; Zhu, Y.-P.; Wen, Y. Hydration of a sulfuric acid-oxalic acid complex: acid dissociation and its atmospheric implication. RSC Adv. 2015, 5, 48638– 48646, DOI: 10.1039/C5RA06116DGoogle Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXptFClu7c%253D&md5=867367bf4afb469541b384519cc4ea02Hydration of a sulfuric acid-oxalic acid complex: acid dissociation and its atmospheric implicationMiao, Shou-Kui; Jiang, Shuai; Chen, Jiao; Ma, Yan; Zhu, Yu-Peng; Wen, Yang; Zhang, Miao-Miao; Huang, WeiRSC Advances (2015), 5 (60), 48638-48646CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Oxalic acid (OA), one of the most common org. acids in the Earth's atm., is expected to enhance the nucleation and growth of nanoparticles contg. sulfuric acid (SA) and water (W); however, the details about the hydration of OA-SA are poorly understood, esp. for the larger clusters with more water mols. We have investigated the structural characteristics and thermodn. of these clusters using d. functional theory at the PW91PW91/6-311++G(3df,3pd) level. The favorable free energies of formation and obvious concns. of the OA-SA-Wn (n = 0-6) clusters at 298.15 K predict that oxalic acid can contribute to the aerosol nucleation process by binding to sulfuric acid and water until n = 6. There is strong temp. dependence for the complexes formation, and the energy order of these complexes is altered from 100 to 400 K, regardless of different cluster sizes or different isomers within the same cluster size. The lower temp. and higher relative humidity promote the formation of hydrates. Addnl., the investigation of acid dissocn. predicts that several acid-dissocd. models could coexist in the atm., specifically when more water mols. are present. Fewer waters may be needed to cause the acid dissocn., as the relative acidity of the cluster increases, which plays a key role in forming relatively stable hydrated clusters of OA-SA. Finally, the Rayleigh scattering properties of OA-SA-Wn (n = 0-6) have been systematically investigated for the first time to further discuss its atm. implication.
- 9Pavuluri, C. M.; Kawamura, K.; Mihalopoulos, N.; Swaminathan, T. Laboratory photochemical processing of aqueous aerosols: Formation and degradation of dicarboxylic acids, oxocarboxylic acids and α- dicarbonyls. Atmos. Chem. Phys. 2015, 15, 7999– 8012, DOI: 10.5194/acp-15-7999-2015Google ScholarThere is no corresponding record for this reference.
- 10Wang, G.; Cheng, C.; Meng, J.; Huang, Y.; Li, J.; Ren, Y. Field observation on secondary organic aerosols during Asian dust storm periods: Formation mechanism of oxalic acid and related compounds on dust surface. Atmos. Environ. 2015, 113, 169– 176, DOI: 10.1016/j.atmosenv.2015.05.013Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXosVKqur0%253D&md5=41d96574d82a771f739b414cde6dcf2bField observation on secondary organic aerosols during Asian dust storm periods: Formation mechanism of oxalic acid and related compounds on dust surfaceWang, Gehui; Cheng, Chunlei; Meng, Jingjing; Huang, Yao; Li, Jianjun; Ren, YanqinAtmospheric Environment (2015), 113 (), 169-176CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)Chem. evolution of East Asian dust during transpacific transport has been given much attention for inorg. species such as sulfate, nitrate and ammonium. However, the role of org. species during the transport has almost entirely been ignored. To understand the formation mechanism of secondary org. aerosols (SOA) on dust surfaces, this study investigated the concns. and compns. of dicarboxylic acids, keto-carboxylic acids, α-dicarbonyls and inorg. ions in size-segregated aerosols (9-stages) collected in Xi'an, central China during the two dust storm episodes in the springs of 2009 and 2011 and compared with those in nondust storm periods. During the events the ambient particulate dicarboxylic acids were 932-2240 ng m-3, which are comparable and even higher than those in nondust periods. Mol. compns. of the above SOA are similar to those in nondust periods with oxalic acid being the leading species. In the presence of the dust storms, all the above mentioned SOA species in Xi'an were predominantly enriched on the coarse particles (>2.1 μm), and oxalic acid well correlated with NO3- (R2 = 0.72, p < 0.001) rather than SO42-. This phenomenon differs greatly from the SOA in any other nondust period that is usually characterized by an enrichment of oxalic acid in fine mode and a strong correlation of oxalic acid with SO42-. We propose a formation pathway to explain these observations, in which nitric acid and/or nitrogen oxides react with dust to produce Ca(NO3)2 and form a liq. phase on the surface of dust aerosols via water vapor-absorption of Ca(NO3)2, followed by a partitioning of the gas-phase water-sol. org. precursors (e.g.,glyoxal and methylglyoxal) into the aq.-phase and a subsequent oxidn. into oxalic acid. To the best of our knowledge, we found for the first time the enrichment of glyoxal and methylglyoxal on dust surface. Our data suggest an important role of nitrate in the heterogeneous formation process of SOA on the surface of dust.
- 11Kawamura, K.; Bikkina, S. A review of dicarboxylic acids and related compounds in atmospheric424 aerosols: Molecular distributions, sources and transformation. Atmos. Res. 2016, 170, 140– 160, DOI: 10.1016/j.atmosres.2015.11.018Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXitVSqt7vF&md5=fcba54277b4abd2ddc9092cefe8141f6A review of dicarboxylic acids and related compounds in atmospheric aerosols: Molecular distributions, sources and transformationKawamura, Kimitaka; Bikkina, SrinivasAtmospheric Research (2016), 170 (), 140-160CODEN: ATREEW; ISSN:0169-8095. (Elsevier B.V.)This review aims to update our understanding on mol. distributions of water-sol. dicarboxylic acids and related compds. in atm. aerosols with a focus on their geog. variability, size distribution, sources and formation pathways. In general, mol. distributions of diacids in aerosols from the continental sites and over the open ocean waters are often characterized by the predominance of oxalic acid (C2) followed by malonic acid (C3) and/or succinic acid (C4), while those sampled over the polar regions often follow the order of C4 ≥ C2 and C3. The most abundant and ubiquitous diacid is oxalic acid, which is principally formed via atm. oxidn. of its higher homologues of long chain diacids and other pollution-derived org. precursors (e.g., olefins and arom. hydrocarbons). However, its occurrence in marine aerosols is mainly due to the transport from continental outflows (e.g., East Asian outflow during winter/spring to the North Pacific) and/or governed by photochem./aq. phase oxidn. of biogenic unsatd. fatty acids (e.g., oleic acid) and isoprene emitted from the productive open ocean waters. The long-range atm. transport of pollutants from mid latitudes to the Arctic in dark winter facilitates to accumulate the reactants prior to their intense photochem. oxidn. during springtime polar sunrise. Furthermore, the relative abundances of C2 in total diacid mass showed similar temporal trends with downward solar irradn. and ambient temps., suggesting the significance of atm. photochem. oxidn. processing. Compd.-specific isotopic analyses of oxalic acid showed the highest δ13C among diacids whereas azelaic acid showed the lowest value, corroborating the significance of atm. aging of oxalic acid. On the other hand, other diacids gave intermediate values between these two diacids, suggesting that aging of oxalic acid is assocd. with 13C enrichment.
- 12Peng, X.-Q.; Huang, T.; Miao, S.-K.; Chen, J.; Wen, H.; Feng, Y.-J. Hydration of oxalic acid–ammonia complex: atmospheric implication and Rayleigh-scattering properties. RSC Adv. 2016, 6, 46582– 47593, DOI: 10.1039/C6RA03164AGoogle ScholarThere is no corresponding record for this reference.
- 13Zhao, H.; Zhang, Q.; Du, L. Hydrogen bonding in cyclic complexes of carboxylic acid–sulfuric acid and their atmospheric implications. RSC Adv. 2016, 6, 71733– 71743, DOI: 10.1039/C6RA16782AGoogle ScholarThere is no corresponding record for this reference.
- 14Elm, J.; Myllys, N.; Kurtén, T. What is required for highly oxidized molecules to form clusters with sulfuric acid?. J. Phys. Chem. A 2017, 121, 4578– 4587, DOI: 10.1021/acs.jpca.7b03759Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXos1amtbY%253D&md5=bca0321a8b08702791e619e491c323daWhat Is Required for Highly Oxidized Molecules To Form Clusters with Sulfuric Acid?Elm, Jonas; Myllys, Nanna; Kurten, TheoJournal of Physical Chemistry A (2017), 121 (23), 4578-4587CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)We have studied the specific requirements of a given neutral org. mol. to act as a stabilizer in sulfuric acid induced new particle formation. Based on an anal. of the reaction Gibbs free energies between simple functional groups and sulfuric acid, carboxylic acid groups are identified to show the strongest hydrogen bonding interaction with sulfuric acid. The free energy assocd. with the hydrogen bonding between sulfuric acid and 14 different carboxylic acids of atm. relevance reveal that the binding strength is very dependent on the ability of sulfuric acid to form an addnl. hydrogen bond via its vacant S-OH group to a γ-carbonyl group in the org. mol. Extending the anal. to monoterpene oxidn. products and further to large dimer esters, we identify the following necessary criteria for a given org. oxidn. product to efficiently stabilize sulfuric acid clustering: (1) weak or no intramol. hydrogen bonds in the isolated monomer; (2) more than two carboxylic acid groups. As a proof of concept we show that these requirements correspond to the docking of a sulfuric acid mol. between two non-interacting carboxylic acid groups in the org. mol. These findings suggests that, for a given org. oxidn. product to participate in the initial steps in new particle formation involving sulfuric acid, very distinct mol. features are required.
- 15Hu, Y.-C.; Zhang, X.-H.; Li, Q.-S.; Zhang, Y.-H.; Li, Z.-S. Effect of water on the structure and stability of hydrogen-bonded oxalic acid dimer. ChemPhyschem 2017, 18, 3375– 3383, DOI: 10.1002/cphc.201700950Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhsleiu7vK&md5=192a1c69aaf0b74c11242298629f69c3Effect of Water on the Structure and Stability of Hydrogen-Bonded Oxalic Acid DimerHu, Yuan-Chun; Zhang, Xiu-Hui; Li, Quan-Song; Zhang, Yun-Hong; Li, Ze-ShengChemPhysChem (2017), 18 (23), 3375-3383CODEN: CPCHFT; ISSN:1439-4235. (Wiley-VCH Verlag GmbH & Co. KGaA)As the simplest and most abundant dicarboxylic acid in the atm., oxalic acid (OA) not only plays a key role in aerosol nucleation, but also acts as a prototypical compd. for the investigation of intra- and intermol. hydrogen-bonding interactions. A systematic theor. study on the hydrated OA dimers performed by using DFT at the M06-2X/6-311++G(3df, 2p) level is discussed herein. The properties of hydrogen bonds in clusters are inspected through topol. anal. by using atoms in mols. (AIM) theory. The most stable OA dimer involves a cyclic structure with two intermol. hydrogen bonds. Calcns. show that one H2O has a slight effect on the hydrogen bonds, whereas two water mols. weaken and three water mols. break the two intermol. hydrogen bonds between OAs. Furthermore, there are no hydrogen-bond interactions between OAs in almost all stable clusters as the no. of H2O mols. increases to four and five. Addnl., ionization and isomerization of OA through water-assisted proton-transfer phenomena are obsd. in tetra- and pentahydrates. This work provides new insights into the conversion of anhyd. OA into hydrated clusters that are helpful for further understanding the atm. nucleation process and nature of hydrogen bond.
- 16Xu, J.; Finlayson-Pitts, B. J.; Gerber, R. B. Proton transfer in mixed clusters of methanesulfonic acid, methylamine, and oxalic acid: implications for atmospheric particle formation. J. Phys. Chem. A 2017, 121, 2377– 2385, DOI: 10.1021/acs.jpca.7b01223Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXktV2kurc%253D&md5=63eae62077f46c6a94d7802ebab7b8f1Proton Transfer in Mixed Clusters of Methanesulfonic Acid, Methylamine, and Oxalic Acid: Implications for Atmospheric Particle FormationXu, Jing; Finlayson-Pitts, Barbara J.; Gerber, R. BennyJournal of Physical Chemistry A (2017), 121 (12), 2377-2385CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Understanding the properties of atm. particles made of several components is a very challenging problem. In this paper, we perform quantum chem. calcns. for small multicomponent clusters of atm. relevance that incorporate methanesulfonic acid (MSA), methylamine (MA), oxalic acid (OxA), and water (H2O). Potential correlations between theor. predictions of proton transfer in the small clusters and findings of recent expts. on formation of particles of detectable sizes (>2 nm) from the same components are studied. It is proposed that proton transfer from the acid to the amine in the 1:1 clusters correlates with expts. on particle formation in systems, such as MSA-MA and MSA-MA-OxA. In the case of OxA + MA, which has been obsd. to give few particles, proton transfer does not occur for the 1:1 cluster but does for the 2:2 cluster. Adding H2O to OxA-MA promotes the occurrence of proton transfer, and corresponding particles are slightly enhanced. The partial charge on the MA component increases by adding OxA or H2O to MSA-MA, which is correlated with enhanced particle formation compared to MSA-MA alone. Ab initio mol. dynamics simulations show that proton transfer at room temp. (T = 298 K) and high temp. (T = 500 K) is little affected compared with the equil. structure (T = 0 K). These results suggest that small cluster calcns. may be useful in predicting the formation of multicomponent particles in the atm.
- 17Chen, J.; Jiang, S.; Liu, Y.-R.; Huang, T.; Wang, C.-Y.; Miao, S.-K. Interaction of oxalic acid with dimethylamine and its atmospheric implications. RSC Adv. 2017, 7, 6374– 6388, DOI: 10.1039/C6RA27945GGoogle Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXht1amur4%253D&md5=9706a306eb06cfcc2afeadce68da99fcInteraction of oxalic acid with dimethylamine and its atmospheric implicationsChen, Jiao; Jiang, Shuai; Liu, Yi-Rong; Huang, Teng; Wang, Chun-Yu; Miao, Shou-Kui; Wang, Zhong-Quan; Zhang, Yang; Huang, WeiRSC Advances (2017), 7 (11), 6374-6388CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Oxalic acid and dimethylamine are the most common org. acid and base in the atm., and are recognized as significant precursor species in atm. new particle formation. However, the interaction between oxalic acid and dimethylamine in the presence of hydration is not yet understood. In this study, the most stable geometric structures and thermodn. of (C2H2O4)m(CH3NHCH3)(H2O)n (m = 1-2, n = 0-4) clusters are investigated using M06-2X coupled with the 6-311+G(2d,p) basis set. A high level explicitly cor. CCSD(T)-F12/VDZ-F12 method is utilized to benchmark the d. functional theory (DFT) methods. Hydration promotes proton transfer from oxalic acid to dimethylamine for (C2H2O4)(CH3NHCH3)(H2O)n (n = 0-4) clusters, while proton transfer from oxalic acid to dimethylamine occurs without hydration for (C2H2O4)2(CH3NHCH3)(H2O)n (n = 0-4) clusters. With regards to the isomer distribution at the potential energy surface, temp. seems not to be an important parameter, since almost all of the global min. for the investigated size range dominate within the investigated temp. range, except for in the (C2H2O4)m(CH3NHCH3)(H2O)2 clusters. Under atm. conditions, the peak hydration distribution shifts from unhydrated clusters to trihydrates for the (C2H2O4)(CH3NHCH3)(H2O)n (n = 0-4) clusters, while for the (C2H2O4)2(CH3NHCH3)(H2O)n (n = 0-4) clusters, unhydrated clusters clearly dominate the cluster distribution, irresp. of whether the humidity is low or high. Finally, the formation free energies obtained from quantum calcns. are used to calc. the evapn. rates. We find that evapn. of dimethylamine is preferred compared to oxalic acid for the (C2H2O4)(CH3NHCH3)(H2O)n clusters, while the results are reversed for the (C2H2O4)2(CH3NHCH3)(H2O)n clusters.
- 18Arquero, K. D.; Xu, J.; Gerber, B.; Finlayson-Pitts, B. J. Particle formation and growth from oxalic acid, methanesulfonic acid, trimethylamine and water: a combined experimental and theoretical study. Phys. Chem. Chem. Phys. 2017, 19, 28286– 28301, DOI: 10.1039/C7CP04468BGoogle Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhs1Crt7fI&md5=d4b35c04200187aab197360823bd56b0Particle formation and growth from oxalic acid, methanesulfonic acid, trimethylamine and water: a combined experimental and theoretical studyArquero, Kristine D.; Xu, Jing; Gerber, R. Benny; Finlayson-Pitts, Barbara J.Physical Chemistry Chemical Physics (2017), 19 (41), 28286-28301CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Atm. particles affect visibility, health, and climate but the mechanisms of their formation from initial clusters and their growth to detectable particles remain largely unknown. Previous studies showed methanesulfonic acid (MSA) reactions with NH3 and amines form particles, a process enhanced by water. Results of a combined exptl./theor. study of oxalic acid (OxA) effect on particle formation and growth from the MSA reaction with trimethylamine (TMA) in the presence/absence of water are reported. Gas phase reactants were mixed in an aerosol flow reactor at 1 atm pressure and 294° K. Particle no. concns. and size distributions were measured as a function of reaction time from 0.8-12 s. The interaction of OxA with TMA with/without water did not lead to significant particle formation. When OxA was present during the MSA reaction with TMA, there was little change (≤2 times more) in particle no. concn., but particles were larger compared vs. the base case (MSA with TMA alone). However, the presence of water with MSA and TMA overwhelmed the OxA effect so that OxA has no significant impact on particle no. concn. or size. Exptl. results suggested MSA hydrate is important for particle formation and growth in the four component OxA/MSA/TMA/water system. Results were compared to earlier studies of the OxA effect on the MSA/methylamine reaction and interpreted based on theor. calcd. properties of small clusters of the components.
- 19Hong, Y.; Liu, Y.-R.; Wen, H.; Miao, S.-K.; Huang, T.; Peng, X.-Q. Interaction of oxalic acid with methylamine and its atmospheric implications. RSC Adv. 2018, 8, 7225– 7234, DOI: 10.1039/C7RA13670FGoogle Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXislSmu7k%253D&md5=e17fd3798381e4fd8f7627d7e7be524dInteraction of oxalic acid with methylamine and its atmospheric implicationsHong, Yu; Liu, Yi-Rong; Wen, Hui; Miao, Shou-Kui; Huang, Teng; Peng, Xiu-Qiu; Jiang, Shuai; Feng, Ya-Juan; Huang, WeiRSC Advances (2018), 8 (13), 7225-7234CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Oxalic acid, which is one of the most common dicarboxylic acids, is expected to be an important component of atm. aerosols. However, the contribution of oxalic acid to the generation of new particles is still poorly understood. In this study, the structural characteristics and thermodn. of (C2H2O4)(CH3NH2)n (n = 1-4) were investigated at the PW91PW91/6-311++G(3df,3pd) level of theory. We found that clusters formed by oxalic acid and methylamine are relatively stable, and the more the atoms participating in the formation of a ring-like structure, the more stable is the cluster. In addn., via the anal. of atm. relevance, it can be revealed that clusters of (C2H2O4)(CH3NH2)n (n = 1-4) have a noteworthy concn. in the atm., which indicates that these clusters could be participating in new particle formation. Moreover, by comparison with (H2C2O4)(NH3)n (n = 1-6) species, it can be seen that oxalic acid is more readily bound to methylamine than to ammonia, which promotes nucleation or new particle formation. Finally, the Rayleigh scattering properties of clusters of (C2H2O4)(CH3NH2)n (n = 1-4) were investigated for the first time to det. their atm. implications.
- 20Boreddy, S. K. R.; Kawamura, K. Investigation on the hygroscopicity of oxalic acid and atmospherically relevant oxalate salts under sub- and supersaturated conditions. Environ. Sci.: Processes Impacts 2018, 20, 1069– 1080, DOI: 10.1039/C8EM00053KGoogle ScholarThere is no corresponding record for this reference.
- 21Lv, G.; Sun, X.; Zhang, C.; Li, M. Understanding the catalytic role of oxalic acid in SO3 hydration to form H2SO4 in the atmosphere. Atmos. Chem. Phys. 2019, 19, 2833– 2844, DOI: 10.5194/acp-19-2833-2019Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXosVahur4%253D&md5=44add63e2ab82d65cb6d2837a2b3d449Understanding the catalytic role of oxalic acid in SO3 hydration to form H2SO4 in the atmosphereLv, Guochun; Sun, Xiaomin; Zhang, Chenxi; Li, MeiAtmospheric Chemistry and Physics (2019), 19 (5), 2833-2844CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)The hydration of SO3 plays an important role in atm. sulfuric acid formation. Some atm. species can be involved in and facilitate the reaction. In this work, using quantum chem. calcns., we show that oxalic acid, the most common dicarboxylic acid in the atm., can effectively catalyze the hydration of SO3. The energy barrier of the SO3 hydration reaction catalyzed by oxalic acid (cTt, tTt, tCt and cCt conformers) is a little higher or less than 1 kcal mol-1, which is lower than the energy barrier of 5.17 kcal mol-1 for water-catalyzed SO3 hydration. Compared with the rates of the SO3 hydration reaction catalyzed by oxalic acid and water, it can be found that in the upper troposphere the OA-catalyzed SO3 hydration can play an important role in promoting SO3 hydration. It leads us to conclude that the involvement of oxalic acid in SO3 hydration to form H2SO4 is significant in the atm.
- 22Deshmukh, D. K.; Kawamura, K.; Gupta, T.; Haque, M. M.; Zhang, Y.-L.; Singh, D. K.; Tsai, Y. I. High Loadings of Water-soluble Oxalic Acid and Related Compounds in PM2.5 Aerosols in Eastern Central India: Influence of Biomass Burning and Photochemical Processing. Aerosol Air Qual. Res. 2019, 19, 2625– 2644, DOI: 10.4209/aaqr.2019.10.0543Google ScholarThere is no corresponding record for this reference.
- 23Zhao, Y.; Liu, Y.-R.; Jiang, S.; Huang, T.; Wang, Z.-H.; Xu, C.-X. Volatile organic compounds enhancing sulfuric acid-based ternary homogeneous nucleation: the important role of synergistic effect. Atmos. Environ. 2020, 233, 117609, DOI: 10.1016/j.atmosenv.2020.117609Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhtVWhs7bI&md5=3e257ecf2eb85da3ef364866615feff2Volatile organic compounds enhancing sulfuric acid-based ternary homogeneous nucleation: The important role of synergistic effectZhao, Yu; Liu, Yi-Rong; Jiang, Shuai; Huang, Teng; Wang, Zi-Hang; Xu, Cai-Xin; Feng, Ya-Juan; Huang, WeiAtmospheric Environment (2020), 233 (), 117609CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)New particle formation (NPF) is an important source of atm. aerosols. Sulfuric acid (SA) and water (W) are recognized as essential participating substances in the nucleation of the atm. In addn., as one of the most common org. acids, oxalic acid (OA) can improve NPF when amine such as methylamine (MA) is present. However, exploring the properties of atm. particles made up of different components is challenging and the role of volatile org. compds. in SA-based ternary homogeneous nucleation is still lacking in research. In this work, the structures and energies of (SA)x(OA)y(MA)z(W)m (0 ≤ x, y, z ≤ 3, 0 ≤ m ≤ 1) are investigated. The results indicate that it is accessible for SA to form clusters with OA and MA mols. through hydrogen bonds and proton transfer interactions. The anal. of non-covalent interactions and proton transfer reveals that ternary nucleation systems have stronger hydrogen bonds and more proton transfers than binary systems to stabilize the clusters. In terms of the thermodn. properties, the Gibbs free energies of the clusters will decrease as the addn. of SA, OA or MA mols. indicating that the synergistic effect of these three substances may be of potential in forming the initial cluster and subsequent growth processes. Moreover, the evapn. rates of clusters show that the synergistic effect of ternary clusters leads to a decrease in evapn. rate which may promote atm. nucleation.
- 24Chen, J. Theoretical analysis of sulfuric acid–dimethylamine–oxalic acid–water clusters and implications for atmospheric cluster formation. RSC Adv. 2022, 12, 22425– 22434, DOI: 10.1039/D2RA03492AGoogle ScholarThere is no corresponding record for this reference.
- 25Palmieri, F.; Estoppey, A.; House, G. L.; Lohberger, A.; Bindschedler, S.; Chain, P. S. G.; Junier, P. Oxalic acid, a molecule at the crossroads of bacterial-fungal interactions. Adv. Appl. Microbiol. 2019, 106, 49– 77, DOI: 10.1016/bs.aambs.2018.10.001Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BB3cfmvFamtA%253D%253D&md5=7de8e8c2bf98155b660ecaa870f1ae77Oxalic acid, a molecule at the crossroads of bacterial-fungal interactionsPalmieri Fabio; Estoppey Aislinn; Lohberger Andrea; Bindschedler Saskia; House Geoffrey L; Chain Patrick S G; Junier PilarAdvances in applied microbiology (2019), 106 (), 49-77 ISSN:0065-2164.Oxalic acid is the most ubiquitous and common low molecular weight organic acid produced by living organisms. Oxalic acid is produced by fungi, bacteria, plants, and animals. The aim of this review is to give an overview of current knowledge about the microbial cycling of oxalic acid through ecosystems. Here we review the production and degradation of oxalic acid, as well as its implications in the metabolism for fungi, bacteria, plants, and animals. Indeed, fungi are well known producers of oxalic acid, while bacteria are considered oxalic acid consumers. However, this framework may need to be modified, because the ability of fungi to degrade oxalic acid and the ability of bacteria to produce it, have been poorly investigated. Finally, we will highlight the role of fungi and bacteria in oxalic acid cycling in soil, plant and animal ecosystems.
- 26Rudnick, M. B.; van Veen, J. A.; de Boer, W. Oxalic acid: a signal molecule for fungus-feeding bacteria of the genus Collimonas ?. Environ. Microbiol. Rep. 2015, 7 (5), 709– 714, DOI: 10.1111/1758-2229.12290Google ScholarThere is no corresponding record for this reference.
- 27Hsieh, L.; Chen, C.; Wan, M.; Tsai, C.; Tsai, Y. Speciation and temporal characterization of dicarboxylic acids in PM2. 5 during a PM episode and a period of non-episodic pollution. Atmos. Environ. 2008, 42, 6836– 6850, DOI: 10.1016/j.atmosenv.2008.05.021Google ScholarThere is no corresponding record for this reference.
- 28Radek, M.; Savage, G. P. Oxalates in Some Indian Green Leafy Vegetables. Int. J. Food Sci. Nutr. 2008, 59, 246– 260, DOI: 10.1080/09637480701791176Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXjtVeqsbc%253D&md5=a5f887d5c343c62daa6a4166bc1ee17cOxalates in some Indian green leafy vegetablesRadek, M.; Savage, G. P.International Journal of Food Sciences and Nutrition (2008), 59 (3), 246-260CODEN: IJFNEH; ISSN:0963-7486. (Informa Healthcare)The sol. and total oxalate contents of 11 leafy vegetables grown in India were detd. Spinach, purple and green amaranth and colocasia contained high levels of total oxalates, which ranged from 5,138.0±37.6 mg/100 g dry matter up to 12,576.1±107.9 mg/100 g dry matter. Seven other leafy vegetables (curry, drumstick, shepu, fenugreek, coriander, radish and onion stalks) contained only insol. oxalate, which ranged from 209.0±5.0 mg/100 g dry matter to 2,774.9±18.4 mg/100 g dry matter. In vitro digestion of the samples showed that the gastric available oxalate was 10% lower than the values obtained from acid extn. and that intestinal available oxalate was 20% lower than the values obtained following hot water extn. The percentage calcium bound in the insol. oxalate fraction of the dried leafy vegetables ranged from 3.3% to 86.7% of the total calcium. Addn. of four different sources of calcium (low fat milk, whole milk, calcium carbonate and calcium sulfate) resulted in a range of 32-100% redns. of intestinal available oxalate in the mixt.
- 29Holmes, R. P.; Kennedy, M. Estimation of the Oxalate Content of Foods and Daily Oxalate Intake. Kidney Int. 2000, 57, 1662– 1667, DOI: 10.1046/j.1523-1755.2000.00010.xGoogle Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXjtVOrtbo%253D&md5=00c6a7ce30d73d679e1d11e1ebd985e3Estimation of the oxalate content of foods and daily oxalate intakeHolmes, Ross P.; Kennedy, MarthaKidney International (2000), 57 (4), 1662-1667CODEN: KDYIA5; ISSN:0085-2538. (Blackwell Science, Inc.)Background. The amt. of oxalate ingested may be an important risk factor in the development of idiopathic calcium oxalate nephrolithiasis. Reliable food tables listing the oxalate content of foods are currently not available. The aim of this research was to develop an accurate and reliable method to measure the food content of oxalate. Methods. Capillary electrophoresis (CE) and ion chromatog. (IC) were compared as direct techniques for the estn. of the oxalate content of foods. Foods were thoroughly homogenized in acid, heat extd., and clarified by centrifugation and filtration before diln. in water for anal. Five individuals consuming self-selected diets maintained food records for three days to det. their mean daily oxalate intakes. Results. Both techniques were capable of adequately measuring the oxalate in foods with a significant oxalate content. With foods of very low oxalate content (<1.8 mg/100 g), IC was more reliable than CE. The mean daily intake of oxalate by the five individuals tested was 152 ± 83 mg, ranging from 44 to 352 mg/day. Conclusions. CE appears to be the method of choice over IC for estg. the oxalate content of foods with a medium (>10 mg/100 g) to high oxalate content due to a faster anal. time and lower running costs, whereas IC may be better suited for the anal. of foods with a low oxalate content. Accurate ests. of the oxalate content of foods should permit the role of dietary oxalate in urinary oxalate excretion and stone formation to be clarified. Other factors, apart from the amt. of oxalate ingested, appear to exert a major influence over the amt. of oxalate excreted in the urine.
- 30Salgado, N.; Silva, M. A.; Figueira, M. E.; Costa, H. S.; Albuquerque, T. G. Oxalate in Foods: Extraction Conditions, Analytical Methods, Occurrence, and Health Implications. Foods 2023, 12, 3201, DOI: 10.3390/foods12173201Google ScholarThere is no corresponding record for this reference.
- 31Siener, R.; Seidler, A.; Hönow, R. Oxalate-rich foods. Food Sci. Technol. 2021, 41, 169– 173, DOI: 10.1590/fst.10620Google ScholarThere is no corresponding record for this reference.
- 32Mochizuki, T.; Kawamura, K.; Miyazaki, Y.; Wada, R.; Takahashi, Y.; Saigusa, N.; Tani, A. Secondary formation of oxalic acid and related organic species from biogenic sources in a larch forest at the northern slope of Mt. Fuji. Atmos. Environ. 2017, 166, 255– 262, DOI: 10.1016/j.atmosenv.2017.07.028Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXht1amsr3N&md5=e3c89cddfcc6e85b8622a0bcbc4f928bSecondary formation of oxalic acid and related organic species from biogenic sources in a larch forest at the northern slope of Mt. FujiMochizuki, Tomoki; Kawamura, Kimitaka; Miyazaki, Yuzo; Wada, Ryuichi; Takahashi, Yoshiyuki; Saigusa, Nobuko; Tani, AkiraAtmospheric Environment (2017), 166 (), 255-262CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)To better understand the formation of water-sol. org. aerosols in the forest atm., we measured low mol. wt. (LMW) dicarboxylic acids, oxocarboxylic acids, α-dicarbonyls, unsatd. fatty acids (UFAs), and water-sol. org. carbon (WSOC) in aerosols from a Larix kaempferi forest located at the northern slope of Mt. Fuji, Japan, in summer 2012. Concns. of dicarboxylic acids, oxocarboxylic acids, α-dicarbonyls, and WSOC showed maxima in daytime. Relative abundance of oxalic acid in LMW dicarboxylic acids was on av. 52% and its av. concn. was 214 ng m-3. We found that diurnal and temporal variations of oxalic acid are different from those of isoprene and α-pinene, whereas biogenic secondary org. aerosols (BSOAs) derived from isoprene and α-pinene showed similar variations with oxalic acid. The mass concn. ratios of oxalic acid/BSOAs were relatively const., although a large variation in the concns. of toluene that is an anthropogenic volatile org. compd. was obsd. These results suggest that formation of oxalic acid is assocd. with the oxidn. of isoprene and α-pinene with O3 and other oxidants in the forest atm. In addn., concns. of UFAs were obsd., for the first time, to decrease dramatically during daytime in the forest. This study demonstrates that forest ecosystem is an important source of oxalic acid and other dicarboxylic acids in the atm.
- 33Kawamura, K.; Bikkina, S. A review of dicarboxylic acids and related compounds in atmospheric aerosols: Molecular distributions, sources and transformation. Atmos. Res. 2016, 170, 140– 160, DOI: 10.1016/j.atmosres.2015.11.018Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXitVSqt7vF&md5=fcba54277b4abd2ddc9092cefe8141f6A review of dicarboxylic acids and related compounds in atmospheric aerosols: Molecular distributions, sources and transformationKawamura, Kimitaka; Bikkina, SrinivasAtmospheric Research (2016), 170 (), 140-160CODEN: ATREEW; ISSN:0169-8095. (Elsevier B.V.)This review aims to update our understanding on mol. distributions of water-sol. dicarboxylic acids and related compds. in atm. aerosols with a focus on their geog. variability, size distribution, sources and formation pathways. In general, mol. distributions of diacids in aerosols from the continental sites and over the open ocean waters are often characterized by the predominance of oxalic acid (C2) followed by malonic acid (C3) and/or succinic acid (C4), while those sampled over the polar regions often follow the order of C4 ≥ C2 and C3. The most abundant and ubiquitous diacid is oxalic acid, which is principally formed via atm. oxidn. of its higher homologues of long chain diacids and other pollution-derived org. precursors (e.g., olefins and arom. hydrocarbons). However, its occurrence in marine aerosols is mainly due to the transport from continental outflows (e.g., East Asian outflow during winter/spring to the North Pacific) and/or governed by photochem./aq. phase oxidn. of biogenic unsatd. fatty acids (e.g., oleic acid) and isoprene emitted from the productive open ocean waters. The long-range atm. transport of pollutants from mid latitudes to the Arctic in dark winter facilitates to accumulate the reactants prior to their intense photochem. oxidn. during springtime polar sunrise. Furthermore, the relative abundances of C2 in total diacid mass showed similar temporal trends with downward solar irradn. and ambient temps., suggesting the significance of atm. photochem. oxidn. processing. Compd.-specific isotopic analyses of oxalic acid showed the highest δ13C among diacids whereas azelaic acid showed the lowest value, corroborating the significance of atm. aging of oxalic acid. On the other hand, other diacids gave intermediate values between these two diacids, suggesting that aging of oxalic acid is assocd. with 13C enrichment.
- 34Claeys, M. Formation of secondary organic aerosols through photooxidation of isoprene. Science 2004, 303, 1173– 1176, DOI: 10.1126/science.1092805Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXhsVWgtb4%253D&md5=43cb23180fe04a94a189bf899aaa121fFormation of Secondary Organic Aerosols Through Photooxidation of IsopreneClaeys, Magda; Graham, Bim; Vas, Gyorgy; Wang, Wu; Vermeylen, Reinhilde; Pashynska, Vlada; Cafmeyer, Jan; Guyon, Pascal; Andreae, Meinrat O.; Artaxo, Paulo; Maenhaut, WillyScience (Washington, DC, United States) (2004), 303 (5661), 1173-1176CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Detailed org. anal. of natural aerosols from the Amazonian rain forest showed considerable quantities of previously unobserved polar org. compds. identified as a mixt. of 2 diastereoisomeric 2-methyltetrols: 2-methylthreitol and 2-methylerythritol. These polyols, which have an isoprene skeleton, can be explained by OH--initiated photooxidn. of isoprene. They have low vapor pressure, allowing them to condense onto pre-existing particles. It was estd. that photooxidn. of isoprene results in an annual global prodn. of ∼2 Tg of the polyols, a substantial fraction of the Intergovernmental Panel on Climate Change est. of 8-40 Tg/yr of secondary org. aerosol from biogenic sources.
- 35Kawamura, K.; Ikushima, K. Seasonal changes in the distribution of dicarboxylic acids in the urban atmosphere. Environ. Sci. Technol. 1993, 27, 2227– 2235, DOI: 10.1021/es00047a033Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXlsVGnt78%253D&md5=7c6f649ee0b03f2817ae1bf072414d93Seasonal changes in the distribution of dicarboxylic acids in the urban atmosphereKawamura, Kimitaka; Ikushima, KouichiEnvironmental Science and Technology (1993), 27 (10), 2227-35CODEN: ESTHAG; ISSN:0013-936X.Low-mol.-wt. dicarboxylic acids were studied in urban aerosol samples collected from Tokyo atm. in 1988-1989, by a capillary GC and GC-mass spectrometry method using a di-Bu ester derivatization technique. A homologous series of C2-C10 dicarboxylic acids were detected in the aerosols, including straight-chain satd., unsatd., branched-chain, and hydroxylated species. Mol. distributions of the diacids demonstrated that oxalic (C2) acid was the most abundant throughout the 4 seasons, followed by C3 and C4 diacids; the C2 diacid comprised 37-69% of the total diacid concns. Diacids with higher C nos. were less abundant, although C6 and C9 diacids were relatively abundant. Total concns. of diacids were 90-1370 ng/m3, which corresponded to 0.06-1.1% of the total aerosol mass. Their concns. relative to aerosol mass showed a seasonal trend with a max. in Aug. The C2-C4 diacid C normalized by aerosol total C also showed a strong seasonal change with a max. in the summer and indicated a pos. correlation with oxidant concns. Interestingly, lower diacids showed higher values for correlation coeff. (e.g., r = 0.80 for oxalic acid), suggesting a preferential prodn. of low mol. wt. dicarboxylic acids, esp., C2 diacid. Possible reaction mechanisms are discussed in relation to the preferential accumulation of lower diacids in the atm.
- 36Prenni, A. J.; DeMott, P. J.; Kreidenweis, S. M.; Sherman, D. E.; Russell, L. M.; Ming, Y. J. Phys. Chem. A 2001, 105, 11240– 11248, DOI: 10.1021/jp012427dGoogle Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXosVegtrc%253D&md5=924edef0b06242bbb43564ebe9065835The effects of low molecular weight dicarboxylic acids on cloud formationPrenni, Anthony J.; DeMott, Paul J.; Kreidenweis, Sonia M.; Sherman, D. Eli; Russell, Lynn M.; Ming, YiJournal of Physical Chemistry A (2001), 105 (50), 11240-11248CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The ubiquitous presence of org. compds. in tropospheric particles requires that their role in aerosol/cloud interactions be accounted for in climate models. In this paper, we present studies that investigate the hygroscopic behavior of org. compds. and their efficiency as ice nuclei. Specifically, results for sol. and partially sol. dicarboxylic acids that have been obsd. in atm. aerosol are discussed. At room temp., we use a humidified tandem differential mobility analyzer (HTDMA) and a condensation particle counter interfaced with a cloud condensation nuclei counter to characterize the water uptake behavior of these acids. The HTDMA data agree quite well with modeled hygroscopic behavior. However, we find that some of the compds. retain water to very low humidities, never exhibiting efflorescence. The studies are extended to lower temps. using a continuous flow ice thermal diffusion chamber to investigate the role of these species in ice nucleation at cirrus conditions. Results suggest that ice formation occurs via homogeneous nucleation for most of these acids, and that nucleation for these acids is not as efficient as that for sulfate aerosol.
- 37Makar, P. A. The estimation of organic gas vapour pressure. Atmos. Environ. 2001, 35, 961– 974, DOI: 10.1016/S1352-2310(00)00343-5Google ScholarThere is no corresponding record for this reference.
- 38Limbeck, A.; Puxbaum, H.; Otter, L.; Scholes, M. C. Atmos. Environ. 2001, 35, 1853– 1862, DOI: 10.1016/S1352-2310(00)00497-0Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXht1yqu7Y%253D&md5=0a1286ec3bec76bafa2486c4e090a43bSemivolatile behavior of dicarboxylic acids and other polar organic species at a rural background site (Nylsvley, RSA)Limbeck, A.; Puxbaum, H.; Otter, L.; Scholes, M. C.Atmospheric Environment (2001), 35 (10), 1853-1862CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Science Ltd.)In this study aerosol samples from the South African savanna were analyzed for their polar org. constituents. Samples were collected with a front/back-up filter tandem system of quartz fiber filters (dual filter strategy). In all samples (n=15) dicarboxylic acids and a variety of phthalates, aldehydes and monocarboxylic acids were obsd. Oxalic acid was the dominating compd. with an av. amt. of 79.2 ng m-3 on the front filter and 11.3 ng m-3 on the back-up filter. The presence of significant concns. of dicarboxylic acids on the back-up filter was rather unexpected. There are two possible sources to explain the presence of individual compds. on the back-up filter - particle penetration through the front filter or adsorption of compd. parts from the gas phase. Interpretation of the data indicates that the dicarboxylic acid concns. on the back-up filters appear to be caused by the adsorption of gaseous org. species. Dicarboxylic acids semivolatile behavior is evident with this results. This conclusion refutes the commonly held view that dicarboxylic acids in the atm. were assocd. with the aerosol phase only. Addnl., it was found that the distribution of dicarboxylic acids between the gas and particle phase in the atm. is not only dependent on their vapor pressures. The actual gas phase concn. appears to be more detd. by the chem. properties of the particles than by pure phys. influences. Surprisingly, malonic acid exhibits an anomaly, as it does not show a semivolatile tendency.
- 39Cheng, C.; Li, M.; Chan, C. K.; Tong, H.; Chen, C.; Chen, D.; Wu, D.; Li, L.; Wu, C.; Cheng, P. Mixing state of oxalic acid containing particles in the rural area of Pearl River Delta, China: implications for the formation mechanism of oxalic acid. Atmos. Chem. Phys. 2017, 17 (15), 9519– 9533, DOI: 10.5194/acp-17-9519-2017Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhs1CmtbjO&md5=b18ab18e9b09d1d72f2c7fa31c0ddfbbMixing state of oxalic acid containing particles in the rural area of pearl river delta, china: implications for the formation mechanism of oxalic acidCheng, Chunlei; Li, Mei; Chan, Chak K.; Tong, Haijie; Chen, Changhong; Chen, Duohong; Wu, Dui; Li, Lei; Wu, Cheng; Cheng, Peng; Gao, Wei; Huang, Zhengxu; Li, Xue; Zhang, Zhijuan; Fu, Zhong; Bi, Yanru; Zhou, ZhenAtmospheric Chemistry and Physics (2017), 17 (15), 9519-9533CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)The formation of oxalic acid and its mixing state in atm. particulate matter (PM) were studied using a single-particle aerosol mass spectrometer (SPAMS) in the summer and winter of 2014 in Heshan, a supersite in the rural area of the Pearl River Delta (PRD) region in China. Oxalic-acid-contg. particles accounted for 2.5 and 2.7 % in total detected ambient particles in summer and winter, resp. Oxalic acid was measured in particles classified as elemental carbon (EC), org. carbon (OC), elemental and org. carbon (ECOC), biomass burning (BB), heavy metal (HM), secondary (Sec), sodium-potassium (NaK), and dust. Oxalic acid was found predominantly mixing with sulfate and nitrate during the whole sampling period, likely due to aq.-phase reactions. In summer, oxalic-acid-contg. particle no. and ozone concn. followed a very similar trend, which may reflect the significant contribution of photochem. reactions to oxalic acid formation. The HM particles were the most abundant oxalic acid particles in summer and the diurnal variations in peak area of iron and oxalic acid show opposite trends, which suggests a possible loss of oxalic acid through the photolysis of iron oxalato-complexes during the strong photochem. activity period. In wintertime, carbonaceous particles contained a substantial amt. of oxalic acid as well as abundant carbon clusters and BB markers. The general existence of nitric acid in oxalic-acid-contg. particles indicates an acidic environment during the formation process of oxalic acid. The peak areas of nitrate, sulfate and oxalic had similar temporal change in the carbonaceous type oxalic acid particles, and the organosulfate-contg. oxalic acid particles correlated well with total oxalic acid particles during the haze episode, which suggests that the formation of oxalic acid is closely assocd. with the oxidn. of org. precursors in the aq. phase.
- 40Yu, S. Role of organic acids formic, acetic, pyruvic and/ oxalic in the formation of cloud condensation nuclei CCN: A Review. Atmos. Res. 2000, 53, 185– 217, DOI: 10.1016/S0169-8095(00)00037-5Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXitlKru7s%253D&md5=a057799ada7073f81e6d73aff1f9549dRole of organic acids (formic, acetic, pyruvic and oxalic) in the formation of cloud condensation nuclei (CCN): a reviewYu, S.Atmospheric Research (2000), 53 (4), 185-217CODEN: ATREEW; ISSN:0169-8095. (Elsevier Science B.V.)A review, with many refs., is given. Although it is believed that org. aerosols play a key role in cloud nucleation and make an important contribution to the cloud condensation nuclei (CCN) population, their specific species remain poorly characterized. This paper reviews the current knowledge of org. acids (mainly formic, acetic, pyruvic and oxalic acids). Without specification, org. acids in this paper refer to these four org. acids in the gas and aerosol phases. This paper analyzes the extent to which org. acids act as CCN and compares the phys. and chem. properties of org. acids with those of CCN. The results show that aerosol formate and acetate concns. range from 0.02 to 5.3 nmol m-3 and from 0.03 to 12.4 nmol m-3, resp., and that between 34 to 77% of formate and between 21 to 66% of acetate are present in the fine fraction of aerosols. It was found that although most (98-99%) of these volatile org. acids are present in the gas phase, their concns. in the aerosol particles are sufficient to make them a good candidate for CCN. The results also show that org. acids may make an important contribution to the formation of CCN in some special sources such as vegetation emissions and biomass-burning. Org. acids are expected to contribute significantly to the ests. of indirect (cloud-mediated) forcing due to aerosols.
- 41Peng, C.; Chan, M. N.; Chan, C. K. The Hygroscopic Properties of Dicarboxylic and Multifunctional Acids: Measurements and UNIFAC Predictions. Environ. Sci. Technol. 2001, 35, 4495– 4501, DOI: 10.1021/es0107531Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXnsFynsbY%253D&md5=fdf0eb4a3d286bd256b1e60985034db3The Hygroscopic Properties of Dicarboxylic and Multifunctional Acids: Measurements and UNIFAC PredictionsPeng, Changgeng; Chan, Man Nin; Chan, Chak K.Environmental Science and Technology (2001), 35 (22), 4495-4501CODEN: ESTHAG; ISSN:0013-936X. (American Chemical Society)The role of water-sol. org. compds. in regard to the hygroscopic properties of atm. aerosols has recently been the subject of many studies. In particular, low-mol.-wt. dicarboxylic acids and some multifunctional org. acids have been found or are expected to exist in atm. aerosols at urban, semiurban, rural, and remote sites. Unlike the case of their inorg. counterparts, the hygroscopic properties of org. acids have not been well characterized. In this study, the hygroscopic properties of selected water-sol. dicarboxylic acids (oxalic acid, malonic acid, succinic acid, and glutaric acid) and multifunctional acids (citric acid, DL-malic acid, and L-(+)-tartaric acid) were studied using single droplets levitated in an electrodynamic balance at 25°. The water activities of bulk samples of dil. solns. were also measured. Solute evapn. was obsd. for the dicarboxylic acids but not for the multifunctional acids. Oxalic acid, succinic acid, and glutaric acid droplets crystallize upon evapn. of water, but, except for glutaric acid droplets, do not deliquesce even at 90% relative humidity (RH). Mass transfer limitation of the deliquescence process was obsd. in glutaric acid. Neither crystn. nor deliquescence was obsd. in malonic acid, citric acid, DL-malic acid, or L-(+)-tartaric acid. Malonic acid and these three hydroxy carboxylic acids absorb water even at RHs much lower than their resp. deliquescence RH. The growth factor (Gf), defined as the ratio of the particle diam. at RH = 10% to that at RH = 90%, of oxalic acid and succinic acid was close to unity, indicating no hygroscopicity in this range. The remaining acids (malonic acid, glutaric acid, citric acid, malic acid, and tartaric acid) showed roughly similar hygroscopicity of a Gf of 1.30-1.53, which is similar to that of "more hygroscopic" aerosols in field measurements reported in the literature. A generalized equation for these four acids, Gf = (1-aw)-0.163, was developed to represent the hygroscopicity of these acids. Water activity predictions from calcns. using the UNIFAC model were found to agree with the measured water activity data to within 40% for most of the acids, but the deviations were as large as about 100% for malic acid and tartaric acid. We modified the functional group interaction parameters of the COOH-H2O, OH-H2O, and OH-COOH pairs by fitting the UNIFAC model with the measured data. The modified UNIFAC model improves the agreement of predictions and measurements to within 38% for all the acids studied.
- 42Kerminen, V.-M.; Ojanen, C.; Pakkanen, T.; Hillamo, R.; Aurela, M.; Meriläinen, J. Low-molecular-weight dicarboxylic acids in an urban and rural atmosphere. J. Aerosol Sci. 2000, 31, 349– 362, DOI: 10.1016/S0021-8502(99)00063-4Google Scholar42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXhslKgurY%253D&md5=2787e1a0647a9c1617370defb34771a7Low-molecular-weight dicarboxylic acids in an urban and rural atmosphereKerminen, Veli-Matti; Ojanen, Christina; Pakkanen, Tuomo; Hillamo, Risto; Aurela, Minna; Merilainen, JouniJournal of Aerosol Science (2000), 31 (3), 349-362CODEN: JALSB7; ISSN:0021-8502. (Elsevier Science Ltd.)Chem. of oxalic, malonic, and succinic acid was studied at the two sites representing the urban and rural conditions, and at a site intermediate between these two. The investigation was based on the particle collection with a virtual impactor and a Berner low-pressure impactor. Concns. of the three diacids displayed large seasonal amplitudes with low values in winter. Suggestive of common sources or atm. formation processes, the correlation between oxalic and malonic acid concns. was high. Both the local traffic and secondary prodn. in the long-range transported air masses seemed to be the important sources for these two acids. Contrary to oxalic and malonic acid, no enrichment at the urban site compared with the rural site was obsd. for succinic acid. The seasonal cycle of this acid resembled that of methanesulfonic acid. The most likely sources for succinic acid in our samples was the secondary prodn. in the long-range transported air, with potentially significant contribution coming from biogenic sources. The three diacids had quite different distributions over the particulate phase. Oxalic acid had a dominant accumulation mode, a clear Aitken mode at sizes below about 0.15 μm of particle diam., and modes corresponding to the sea-salt and crustal particle size ranges. Most of the malonic acid was assocd. with sea-salt particles, even though in a few samples an accumulation mode was also present. Succinic acid was distributed between the accumulation and the sea-salt particle modes, in addn. to which it frequently had quite a pronounced Aitken mode. Oxalic and succinic acids are among the orgs. that may contribute to the atm. cloud condensation nuclei prodn. Oxalic and malonic acid, and to a smaller extent succinic acid, participate in reactions occurring in sea-salt particles.
- 43Khwaja, H. A. Atmospheric concentrations of carboxylic acids and related compounds at a semiurban site. Atmos. Environ. 1995, 29, 127– 139, DOI: 10.1016/1352-2310(94)00211-3Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXjtVOltbk%253D&md5=4b4b585b07cfe89ea6931ba997442f9fAtmospheric concentrations of carboxylic acids and related compounds at a semiurban siteKhwaja, Haider A.Atmospheric Environment (1995), 29 (1), 127-39CODEN: AENVEQ; ISSN:1352-2310. (Elsevier)Atm.-gas- and particle-phase carboxylic acids and related compds. were measured during Oct. 1991 in a semiurban site in northeastern US. Formic and acetic acids were present in the atm. mostly in the gaseous form with <10% in the particle phase. Concns. of formic acid and acetic acid were in the 0.80-2.5 and 0.60-3.4 ppbv range, resp. Formic acid was correlated with acetic acid (r =0.94). Diurnal variations of formic acid were similar to those of O3. This pattern is consistent with photochem. prodn. as an important source. Formic-to-acetic acid ratios <1 were recorded, likely due to an increase in acetic acid contribution from direct emissions assocd. with vehicular traffic. HCHO levels began to decrease after the mid-afternoon maxima; concns. were 0.63-3.7 ppbv. Formic, acetic, pyruvic, glyoxalic, oxalic, succinic, and malonic acids in the particle phase were identified. Aerosol carboxylic acid concns. were 26-360 ng/m3. About 80% of the carboxylic acid aerosol mass was in the size fraction <1.0 μm diam. Of the dicarboxylic acids, oxalic acid was the most abundant species, followed by succinic acid and malonic acid. Particulate total org. C exhibited a concn. range 12.6 × 103-49.9 × 103 ng/cm3. The obsd. amts. of carboxylic acids in the particle phase accounted for a small fraction of the org. C. Results indicated that photochem. processes and anthropogenic emissions such as automobile exhaust are major sources of atm. carboxylic acids.
- 44Chebbi, A.; Carlier, P. Atmos. Environ. 1996, 30, 4233– 4249, DOI: 10.1016/1352-2310(96)00102-1Google Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XmtFaqsLo%253D&md5=c0a6e50637ff4490082745a52af84c1fCarboxylic acids in the troposphere, occurrence, sources, and sinks: a reviewChebbi, A.; Carlier, P.Atmospheric Environment (1996), 30 (24), 4233-4249CODEN: AENVEQ; ISSN:1352-2310. (Elsevier)A review, with many refs., is given. Carboxylic acids are ubiquitous and important components of the troposphere; they are currently measured in different environments. They are thought to have several sources comprising primary biogenic and anthropogenic emissions, hydrocarbons gas-phase oxidns., and some carbonyl compds. aq.-phase oxidns. We make a synthesis of the concns. of low mol. wt. carboxylic acids in tropospheric aq. and gaseous phases and in aerosol particles for different environments. We also successively exam. the major sources of carboxylic acids and discuss their relative contribution to tropospheric concns. for various environments as well as the principal sinks of these compds.
- 45Martinelango, P. K.; Dasgupta, P. K.; Al-Horr, R. S. Atmos. Environ. 2007, 41, 4258– 4269, DOI: 10.1016/j.atmosenv.2006.05.085Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXlsFaktrc%253D&md5=8cfe5cbfecf8eb9c71292811467424dbAtmospheric production of oxalic acid/oxalate and nitric acid/nitrate in the Tampa Bay airshed: Parallel pathwaysMartinelango, P. Kalyani; Dasgupta, Purnendu K.; Al-Horr, Rida S.Atmospheric Environment (2007), 41 (20), 4258-4269CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)Oxalic acid is the dominant dicarboxylic acid (DCA), and it constitutes up to 50% of total atm. DCAs, esp. in non-urban and marine atmospheres. A significant amt. of particulate H2Ox/oxalate (Ox) occurred in the coarse particle fraction of a dichotomous sampler, the ratio of oxalate concns. in the PM10 to PM2.5 fractions ranged from 1 to 2, with mean±sd being 1.4 ± 0.2. These results suggest that oxalate does not solely originate in the gas phase and condense into particles. Gaseous H2Ox concns. are much lower than particulate Ox concns. and are well correlated with HNO3, HCHO, and O3, supporting a photochem. origin. Of special relevance to the Bay Region Atm. Chem. Expt. (BRACE) is the extent of nitrogen deposition in the Tampa Bay estuary. Hydroxyl radical is primarily responsible for the conversion of NO2 to HNO3, the latter being much more easily deposited. Hydroxyl radical is also responsible for the aq. phase formation of oxalic acid from alkenes. Hence, we propose that an est. of ·OH can be obtained from H2Ox/Ox prodn. rate and we accordingly show that the product of total oxalate concn. and NO2 concn. approx. predicts the total nitrate concn. during the same period.
- 46Ma, X.; Sun, Y.; Huang, Z.; Zhang, Q.; Wang, W. A density functional theory study of the molecular interactions between a series of amides and sulfuric acid. Chemosphere 2019, 214, 781– 790, DOI: 10.1016/j.chemosphere.2018.08.152Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhvVOjsLzF&md5=5d4d3a26a63f7f4da197456377140e62A density functional theory study of the molecular interactions between a series of amides and sulfuric acidMa, Xiaohui; Sun, Yanhui; Huang, Zixiao; Zhang, Qingzhu; Wang, WenxingChemosphere (2019), 214 (), 781-790CODEN: CMSHAF; ISSN:0045-6535. (Elsevier Ltd.)Amides, a class of nitrogen-contg. org. pollutants in the atm., may affect the formation of atm. aerosols by the interactions with sulfuric acid. Here, the mol. interactions of sulfuric acid with formamide, methylformamide, DMF, acetamide, methylacetamide and dimethylacetamide was investigated by d. functional theory. Geometry optimization and Gibbs free energy calcn. were carried out at M06-2X/6-311++G(3df,3pd) level. The results indicate that the addn. of amides to H2SO4 might have a promoting effect on atm. new particle formation at 298.15 K and 1 atm. In the initial stage of new particle formation, the binding capacity of amides and sulfuric acid is stronger than ammonia, but weaker than methylamine. It is worth noting that the trans-methylacetamide could have similar capabilities of stabilizing sulfuric acid as dimethylamine. In the presence of water, amides are found to only have a weak enhancement capability on new particle formation. In addn., we can infer from evapn. rate that the small mol. clusters of formamide and sulfuric acid may be more energetically favorable than macromol. clusters.
- 47Hanson, D. R.; Eisele, F. L. Measurement of prenucleation molecular clusters in the NH3, H2SO4, H2O system. J. Geophys. Res.: Atmos. 2002, 107, AAC 10– 1-AAC 10-18, DOI: 10.1029/2001JD001100Google ScholarThere is no corresponding record for this reference.
- 48Sipilää, M.; Berndt, T.; Petäjä, T.; Brus, D.; Vanhanen, J.; Stratmann, F.; Patokoski, J.; Mauldin, R. L., 3; Hyvärinen, A.-P.; Lihavainen, H. The role of sulfuric acid in atmospheric nucleation. Science 2010, 327, 1243– 1246, DOI: 10.1126/science.1180315Google ScholarThere is no corresponding record for this reference.
- 49Kirkby, J.; Curtius, J.; Almeida, J.; Dunne, E.; Duplissy, J.; Ehrhart, S. Role of sulphuric acid, ammonia and galactic cosmic rays in atmospheric aerosol nucleation. Nature 2011, 476, 429– 433, DOI: 10.1038/nature10343Google Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtVOmsr7F&md5=ab78f81c052b4b7a47e9c02f6b670b94Role of sulphuric acid, ammonia and galactic cosmic rays in atmospheric aerosol nucleationKirkby, Jasper; Curtius, Joachim; Almeida, Joao; Dunne, Eimear; Duplissy, Jonathan; Ehrhart, Sebastian; Franchin, Alessandro; Gagne, Stephanie; Ickes, Luisa; Kuerten, Andreas; Kupc, Agnieszka; Metzger, Axel; Riccobono, Francesco; Rondo, Linda; Schobesberger, Siegfried; Tsagkogeorgas, Georgios; Wimmer, Daniela; Amorim, Antonio; Bianchi, Federico; Breitenlechner, Martin; David, Andre; Dommen, Josef; Downard, Andrew; Ehn, Mikael; Flagan, Richard C.; Haider, Stefan; Hansel, Armin; Hauser, Daniel; Jud, Werner; Junninen, Heikki; Kreissl, Fabian; Kvashin, Alexander; Laaksonen, Ari; Lehtipalo, Katrianne; Lima, Jorge; Lovejoy, Edward R.; Makhmutov, Vladimir; Mathot, Serge; Mikkilae, Jyri; Minginette, Pierre; Mogo, Sandra; Nieminen, Tuomo; Onnela, Antti; Pereira, Paulo; Petaejae, Tuukka; Schnitzhofer, Ralf; Seinfeld, John H.; Sipilae, Mikko; Stozhkov, Yuri; Stratmann, Frank; Tome, Antonio; Vanhanen, Joonas; Viisanen, Yrjo; Vrtala, Aron; Wagner, Paul E.; Walther, Hansueli; Weingartner, Ernest; Wex, Heike; Winkler, Paul M.; Carslaw, Kenneth S.; Worsnop, Douglas R.; Baltensperger, Urs; Kulmala, MarkkuNature (London, United Kingdom) (2011), 476 (7361), 429-433CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Model calcns. suggest that almost half of the global cloud condensation nuclei in the atm. boundary layer may originate from the nucleation of aerosols from trace condensable vapors, although the sensitivity of the no. of cloud condensation nuclei to changes of nucleation rate may be small. Despite extensive research, fundamental questions remain about the nucleation rate of sulfuric acid particles and the mechanisms responsible, including the roles of galactic cosmic rays and other chem. species such as ammonia. First results from the Cosmics Leaving OUtdoor Droplets (CLOUD) expt. at CERN are presented. It was found that atmospherically relevant ammonia mixing ratios of 100 parts per trillion by vol., or less, increase the nucleation rate of sulfuric acid particles more than 100-1,000-fold. Time-resolved mol. measurements reveal that nucleation proceeds by a base-stabilization mechanism involving the stepwise accretion of ammonia mols. Ions increase the nucleation rate by an addnl. factor of between two and more than ten at ground-level galactic-cosmic-ray intensities, provided that the nucleation rate lies below the limiting ion-pair prodn. rate. Ion-induced binary nucleation of H2SO4-H2O can occur in the mid-troposphere but is negligible in the boundary layer. However, even with the large enhancements in rate due to ammonia and ions, atm. concns. of ammonia and sulfuric acid are insufficient to account for obsd. boundary-layer nucleation.
- 50Elm, J.; Kubečka, J.; Besel, V.; Jääskeläinen, M. J.; Halonen, R.; Kurtén, T.; Vehkamäki, H. Modeling the formation and growth of atmospheric molecular clusters: A review. J. Aerosol Sci. 2020, 149, 105621, DOI: 10.1016/j.jaerosci.2020.105621Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhtlyisLfI&md5=5c62c26f1491d26b091d31fa8d341801Modeling the formation and growth of atmospheric molecular clusters: A reviewElm, Jonas; Kubecka, Jakub; Besel, Vitus; Jaaskelainen, Matias J.; Halonen, Roope; Kurten, Theo; Vehkamaki, HannaJournal of Aerosol Science (2020), 149 (), 105621CODEN: JALSB7; ISSN:0021-8502. (Elsevier Ltd.)A review,. Mol. clusters are ubiquitous constituents of the ambient atm., that can grow into larger sizes forming new aerosol particles. The formation and growth of small clusters into aerosol particles remain one of the largest uncertainties in global climate predictions. This has made the modeling of atm. mol. clustering into an active field of research, yielding direct mol. level information about the formation mechanism. We review the present state-of-the-art quantum chem. methods and cluster distribution dynamics models that are applied to study the formation and growth of atm. mol. clusters. We outline the current challenges in applying theor. methods and the future directions to move the field forward.
- 51Zhang, R.; Khalizov, A.; Wang, L.; Hu, M.; Xu, W. Nucleation and Growth of Nanoparticles in the Atmosphere. Chem. Rev. 2012, 112, 1957– 2011, DOI: 10.1021/cr2001756Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtl2itrnN&md5=a5b16c46ecf502c603291ce50de25604Nucleation and Growth of Nanoparticles in the AtmosphereZhang, Renyi; Khalizov, Alexei; Wang, Lin; Hu, Min; Xu, WenChemical Reviews (Washington, DC, United States) (2012), 112 (3), 1957-2011CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review which critically assesses recent findings related to nucleation and growth of atm. nanoparticles, emphasizing process understanding on a fundamental mol. level, is given. Topics discussed include: vapor nucleation overview (nucleation theories and computational approaches [classical nucleation, kinetic, and d. functional theories, mol. dynamics and Monte Carlo methods, nucleation theorem], nucleation expts. [adiabatic expansion approaches; diffusion, laminar flow, and turbulent mixing chambers; continuous generation of nucleating vapors from chem. reaction sources; exptl. results vs. nucleation theories]); atm. nanoparticle nucleation (measurements [concns. and size distribution, nanoparticle chem. compn., charged and neutral atm. clusters], lab. studies [binary nucleation of H2SO4/water; NH3 and amine-involved ternary nucleation of H2SO4/water; org. acid assisted nucleation of H2SO4/water; IOx and ion-induced nucleation; nucleating cluster chem. compn., reactivity, and thermodn.], theor./computational studies [quantum chem. calcns., mol. dynamics and Monte Carlo simulations], atm. nucleation parameterization); nanoparticle atm. growth (Kelvin [curvature] effect, condensation [H2SO4, low volatility orgs.], heterogeneous reactions [NH3, amines, aldehydes, α-dicarbonyls, alcs., other species]); numerical treatment of ambient nanoparticle nucleation and growth rates (measured, combined growth including condensation and intra-/extra-modal coagulation, nucleation rate derivation from atm. measurements); and summary and future research needs.
- 52Zheng, J.; Ma, Y.; Chen, M.; Zhang, Q.; Wang, L.; Khalizov, A. F. Measurement of atmospheric amines and ammonia using the high resolution time-of-flight chemical ionization mass spectrometry. Atmos. Environ. 2015, 102, 249– 259, DOI: 10.1016/j.atmosenv.2014.12.002Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXitVWgtLbE&md5=c4fae6404f558c380000c0c9b1939556Measurement of atmospheric amines and ammonia using the high resolution time-of-flight chemical ionization mass spectrometryZheng, Jun; Ma, Yan; Chen, Mindong; Zhang, Qi; Wang, Lin; Khalizov, Alexei F.; Yao, Lei; Wang, Zhen; Wang, Xing; Chen, LinxiAtmospheric Environment (2015), 102 (), 249-259CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)Ammonia (NH3) and amines play important roles in the nucleation and growth of atm. aerosols. To identify the sources of these chems. in the densely populated and industrialized Yangtze River Delta region of China, we conducted measurements of NH3 and several amines, including methylamine (CH3NH2), C2-amines (C2H7N), and C3-amines (C3H9N) at a suburban site of Nanjing, China, during summer 2012. Using a high-resoln. time-of-flight chem. ionization mass spectrometer (HRToF-CIMS, Aerodyne), 1-min-averaged concns. of NH3 and amines ranged from a few parts per trillions by vol. (pptv) to dozens of ppb by vol. (ppbv). The av. ± 1σ concns. of NH3 and total amines during the measurement period were 1.7 ± 2.3 ppbv and 7.2 ± 7.4 pptv, resp. Among the amines, C2-amines were the most abundant, accounting for 54% of the total amine loading. Significant correlations between NH3 and all three types of amines (0.65 < r2 < 0.80) indicate similar emission sources. Anal. of meteorol. conditions indicated that these NH3 and amine laden air masses mainly originated from nearby industrial areas where NH3 was used for selective catalytic redn. of nitrogen oxides (NOx). The results of this work indicate that industrial emissions in Nanjing, China may have a significant impact on local and regional aerosol chem. by supplying considerable amt. of amines.
- 53Nair, A. A.; Yu, F. Quantification of Atmospheric Ammonia Concentrations: A Review of Its Measurement and Modeling. Atmosphere 2020, 11, 1092, DOI: 10.3390/atmos11101092Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhslentL8%253D&md5=96c5c1f82ce9b9a9843ad6e175f3cf64Quantification of atmospheric ammonia concentrations: a review of its measurement and modelingNair, Arshad Arjunan; Yu, FangqunAtmosphere (2020), 11 (10), 1092CODEN: ATMOCZ; ISSN:2073-4433. (MDPI AG)Ammonia (NH3), the most prevalent alk. gas in the atm., plays a significant role in PM2.5 formation, atm. chem., and new particle formation. This paper reviews quantification of [NH3] through measurements, satellite-remote-sensing, and modeling reported in over 500 publications towards synthesizing the current knowledge of [NH3], focusing on spatiotemporal variations, controlling processes, and quantification issues. Most measurements are through regional passive sampler networks. [NH3] hotspots are typically over agricultural regions, such as the Midwest US and the North China Plain, with elevated concns. reaching monthly avs. of 20 and 74 ppbv, resp. Topog. effects dramatically increase [NH3] over the Indo-Gangetic Plains, North India and San Joaquin Valley, US. Measurements are sparse over oceans, where [NH3] ≈ a few tens of pptv, variations of which can affect aerosol formation. Satellite remote-sensing (AIRS, CrIS, IASI, TANSO-FTS, TES) provides global [NH3] quantification in the column and at the surface since 2002. Modeling is crucial for improving understanding of NH3 chem. and transport, its spatiotemporal variations, source apportionment, exploring physicochem. mechanisms, and predicting future scenarios. GEOS-Chem (global) and FRAME (UK) models are commonly applied for this. A synergistic approach of measurements↔satellite-inference↔modeling is needed towards improved understanding of atm. ammonia, which is of concern from the standpoint of human health and the ecosystem.
- 54Elm, J. An Atmospheric Cluster Database Consisting of Sulfuric Acid, Bases, Organics, and Water. ACS Omega 2019, 4 (6), 10965– 10974, DOI: 10.1021/acsomega.9b00860Google Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXht1aksr%252FP&md5=ebba43602afc4822561711a19d343118An Atmospheric Cluster Database Consisting of Sulfuric Acid, Bases, Organics, and WaterElm, JonasACS Omega (2019), 4 (6), 10965-10974CODEN: ACSODF; ISSN:2470-1343. (American Chemical Society)We have collected, re-computed and compiled a database consisting of 633 unique atmospherically relevant mol. clusters contg. sulfuric acid, bases, oxidized org. compds. and water. The database comprise of strongly hydrogen bonded mol. clusters, and span both neutral, neg. charged and pos. charged clusters of atm. relevance. All the cluster structures and vibrational frequencies were re-evaluated at the ωB97X-D/6-31++G(d,p) level of theory and the single point energies were refined using a high level DLPNO-CCSD(T)/aug-cc-pVTZ calcn. The database unifies published atm. mol. clusters under a single common methodol. and serves as an efficient look-up table for mol. cluster structures and thermochem. parameters. Utilizing the database the performance of four semi-empirical methodologies (PM6, PM7, B97-3c and PBEh-3c) in calcg. the binding energies of atm. mol. clusters is assessed. The B97-3c and PBEh-3c empirically cor. DFT methods yield low errors in the binding energies compared to DLPNO-CCSD(T)/aug-cc-pVTZ ref. results and that a simple linear model can be utilized for estg. accurate binding energies based on ωB97X-D/6-31++G(d,p) results.
- 55Lee, S.-H.; Gordon, H.; Yu, H.; Lehtipalo, K.; Haley, R.; Li, Y.; Zhang, R. New particle formation in the atmosphere: From molecular clusters to global climate. J. Geophys. Res.: Atmos. 2019, 124, 7098– 7146, DOI: 10.1029/2018JD029356Google Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhsVert7zF&md5=810a832a80a73b6ec7fada9f4299ee4eNew Particle Formation in the Atmosphere: From Molecular Clusters to Global ClimateLee, Shan-Hu; Gordon, Hamish; Yu, Huan; Lehtipalo, Katrianne; Haley, Ryan; Li, Yixin; Zhang, RenyiJournal of Geophysical Research: Atmospheres (2019), 124 (13), 7098-7146CODEN: JGRDE3; ISSN:2169-8996. (Wiley-Blackwell)New particle formation (NPF) represents the first step in the complex processes leading to formation of cloud condensation nuclei. Newly formed nanoparticles affect human health, air quality, weather, and climate. This review provides a brief history, synthesizes recent significant progresses, and outlines the challenges and future directions for research relevant to NPF. New developments include the emergence of state-of-the-art instruments that measure prenucleation clusters and newly nucleated nanoparticles down to about 1 nm; systematic lab. studies of multicomponent nucleation systems, including collaborative expts. conducted in the Cosmics Leaving Outdoor Droplets chamber at CERN; observations of NPF in different types of forests, extremely polluted urban locations, coastal sites, polar regions, and high-elevation sites; and improved nucleation theories and parameterizations to account for NPF in atm. models. The challenges include the lack of understanding of the fundamental chem. mechanisms responsible for aerosol nucleation and growth under diverse environments, the effects of SO2 and NOx on NPF, and the contribution of anthropogenic org. compds. to NPF. It is also crit. to develop instruments that can detect chem. compn. of particles from 3 to 20 nm and improve parameterizations to represent NPF over a wide range of atm. conditions of chem. precursor, temp., and humidity.
- 56Nahlovska, Z.; Nahlovsky, B.; Strand, T. G. Molecular Structure of Gaseous Oxalic Acid from Electron Diffraction and IR Data. Acta Chem. Scand. 1970, 24 (24), 2617– 2628, DOI: 10.3891/acta.chem.scand.24-2617Google ScholarThere is no corresponding record for this reference.
- 57Redington, R. L.; Redington, T. E. Infrared Matrix-Isolation Spectra of Monomeric Oxalic Acid. J. Mol. Struct. 1978, 48, 165, DOI: 10.1016/0022-2860(78)80019-2Google ScholarThere is no corresponding record for this reference.
- 58Back, R. A. The ultraviolet absorption spectrum of oxalic acid vapor. Can. J. Chem. 1984, 62, 1414, DOI: 10.1139/v84-241Google ScholarThere is no corresponding record for this reference.
- 59Nieminen, J.; Rasanen, M.; Murto, J. Matrix Isolation and ab Initio Studies of Oxalic Acid. J. Phys. Chem. 1992, 96, 5303, DOI: 10.1021/j100192a024Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38XktFOjs7Y%253D&md5=c24dcd0c7962acaec37f24739762f0f8Matrix-isolation and ab initio studies of oxalic acidNieminen, J.; Rasanen, M.; Murto, J.Journal of Physical Chemistry (1992), 96 (13), 5303-8CODEN: JPCHAX; ISSN:0022-3654.Rotamerization and UV photochem. decompn. of oxalic acid were studied in solid Ne, Ar, and Xe matrixes. The energies and spectra of the conformers and photoproducts were also simulated by extensive ab initio calcns. at MP2/6-31G** (energies, geometries) and MP2/4-31G* (numerical frequencies) levels. Two conformers were present in matrixes, the lowest energy species being cTc (with two intramol. H bonds) and species cTt (with only one intramol. H bond) being at ∼7 kJ mol-1 higher in energy. The uphill process cTc → cTt could be initiated by UV excitation at wavelengths of ∼270 nm. IR irradn. at wavenumbers >2000 cm-1 promoted the reverse interconversion. Full Xe arc irradn. decomps. oxalic acid in matrixes, the product yield ratio being different in different hosts. On the basis of ab initio calcns., the possible transition states for proton shifts are discussed. The calcd. vibrational spectra are given for six conformers of oxalic acid, the spectra of the obsd. conformers are assigned, and their potential energy distributions are given.
- 60Maçôas, E. M. S.; Fausto, R.; Pettersson, M.; Khriachtchev, L.; Rasanen, M. Infrared-Induced Rotamerization of Oxalic Acid Monomer in Argon Matrix. J. Phys. Chem. A 2000, 104, 6956, DOI: 10.1021/jp000634sGoogle Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXks1WntrY%253D&md5=c4f721a7f9f3c9260f45d9922984fcceInfrared-Induced Rotamerization of Oxalic Acid Monomer in Argon MatrixMacoas, Ermelinda M. S.; Fausto, Rui; Pettersson, Mika; Khriachtchev, Leonid; Raesaenen, MarkkuJournal of Physical Chemistry A (2000), 104 (30), 6956-6961CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)IR-induced conformational isomerization of oxalic acid monomer isolated in an argon matrix at 7.5 K was studied by IR spectroscopy. For the first time, three conformational states of this mol. were identified exptl. and their vibrational spectra are assigned. In good agreement with d. functional theory predictions, all the obsd. conformers exhibit a trans O:C-C:O axis, differing in the relative conformation of their O-C-O-H axes. In the most stable conformer (belonging to the C2h symmetry point group), two intramol. OH···O: hydrogen bonds are present. The second (Cs) most stable conformer shows a single OH···O: bond, and the third one (C2h) does not exhibit any intramol. hydrogen bond. Using narrowband tunable irradn. in the near-IR region it was possible to promote very efficiently conformer interconversions, which was followed spectroscopically.
- 61Godfrey, P. D.; Mirabella, M. J.; Brown, R. D. Structural Studies of Higher Energy Conformers by Millimeter-Wave Spectroscopy: Oxalic Acid. J. Phys. Chem. A 2000, 104, 258, DOI: 10.1021/jp992499tGoogle Scholar61https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXnvFKgt7s%253D&md5=24b59c4bf9ff2379b6d57e6f4454dd97Structural Studies of Higher Energy Conformers by Millimeter-Wave Spectroscopy: Oxalic AcidGodfrey, Peter D.; Mirabella, Mathew J.; Brown, Ronald D.Journal of Physical Chemistry A (2000), 104 (2), 258-264CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The authors have detected a polar conformer of oxalic acid and studied its structure using Stark-modulated free jet microwave spectroscopy. By comparing obsd. and ab initio predicted spectroscopic rotational consts. and deuteration isotopic substitution coordinates, the species was identified with the cis, trans, trans conformer (cTt), in which one of the carboxyl H atoms is involved in an intramol. H bond with the carbonyl O of the other carboxyl group. Stark effect measurements yielded elec. dipole moment values of μa = 2.850(6) D; μb = 1.148(3) D, μc = 0.0 D (assumed), and hence μtotal = 3.073(6) D. Ab initio calcns. at the MP2/6-311++G(d,p) level predict and expt. confirms that no other polar conformers occur in detectable concns. in the vapor at the expt. spectral search scan preexpansion temp. of 458 K. However, 2 undetectable species, one nonpolar (tTt), the other of low polarity (cTc, μ = 0.6 D) are predicted to be present in greater concns. than the obsd. species. Predicted relative stabilities of the different conformers are appreciably affected by the inclusion of Gibbs free energy corrections via the vibrational partition function. Large amplitude anharmonic vibrations, for which numerical integration of the vibrational Schrodinger equation is required for obtaining the correction, have an important influence on predicting relative stabilities of the oxalic acid conformers.
- 62Buemi, G. DFT study of the hydrogen bond strength and IR spectra of formic, oxalic, glyoxylic and pyruvic acids in vacuum, acetone and water solution. J. Phys. Org. Chem. 2009, 22, 933– 947, DOI: 10.1002/poc.1543Google Scholar62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtFGlsbfO&md5=8fc7874503b3dde8f679d889f25a6f0dDFT study of the hydrogen bond strength and IR spectra of formic, oxalic, glyoxylic and pyruvic acids in vacuum, acetone and water solutionBuemi, GiuseppeJournal of Physical Organic Chemistry (2009), 22 (10), 933-947CODEN: JPOCEE; ISSN:0894-3230. (John Wiley & Sons Ltd.)The mol. geometries of the possible conformations of formic, oxalic, glyoxylic and pyruvic acids have been fully optimized at DFT B3LYP/6-311++G(d,p) levels of calcn. in vacuum as well as in water and acetone soln. Solns. were treated according to the SCRF PCM approach but some formic acid-water and formic acid-acetone clusters as well as adducts of oxalic acid with two or four water mols. were also taken into account for testing the importance of specific solute-solvent effects. All the most stable isomers of the title compds. are characterized by weak intramol. hydrogen bonds, whose strengths (EHB) cannot be correctly estd. as stability difference between the open and chelate forms since the energy of the former isomer is, in turn, stabilized by a weak hydrogen bridge due to the formic acid moiety. Following the Rotation Barrier Method (RBM), proposed some years ago, EHB in the examd. mols. (gas phase) falls in the range of 18-22 kJ/mol for oxalic acid (9.6 kJ/mol for the c-C-t isomer), 16.8 kJ/mol for glyoxylic acid and 19.8 kJ/mol for pyruvic acid. Most of them disappear at all, or nearly at all, both in acetone and aq. soln., in consequence of the solvent effect. The frequencies of the OH and C-O stretching modes, calcd. according to the anharmonic oscillator model, are in very good agreement with the exptl. literature data, where available. Copyright © 2009 John Wiley & Sons, Ltd.
- 63Feller, D.; Bross, D. H.; Ruscic, B. Enthalpy of Formation of C2H2O4 (Oxalic Acid) from High-Level Calculations and the Active Thermochemical Tables Approach. J. Phys. Chem. A 2019, 123, 3481– 3496, DOI: 10.1021/acs.jpca.8b12329Google Scholar63https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXlvVSmtr0%253D&md5=21c3ad5929bd881d97a29842d28c7598Enthalpy of Formation of C2H2O4 (Oxalic Acid) from High-Level Calculations and the Active Thermochemical Tables ApproachFeller, David; Bross, David H.; Ruscic, BrankoJournal of Physical Chemistry A (2019), 123 (16), 3481-3496CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)High-level coupled cluster calcns. obtained with the Feller-Peterson-Dixon (FPD) approach and new data from the most recent version of the Active Thermochem. Tables (ATcT) are used to reassess the enthalpy of formation of gas-phase C2H2O4 (oxalic acid). The theor. value was further calibrated by comparing FPD and ATcT gas-phase enthalpies of formation for H2CO (formaldehyde) and the two low-lying conformations of C2H4O2 (syn and anti acetic acid). The FPD approach produces a theor. enthalpy of formation of gas-phase oxalic acid of -732.2 ± 4.0 kJ/mol at 298.15 K (-721.8 ± 4.0 kJ/mol at 0 K). An independently obtained ATcT value, based on reassessing the existent exptl. detns. and expanding the resulting thermochem. network with select mid-level composite theor. results, disagrees with several earlier recommendations that were based solely on exptl. detns. but is in excellent accord with the current FPD value. The inclusion of the latter in the most recent ATcT thermochem. network produces a further refined value for the gas-phase enthalpy of formation, -731.6 ± 1.2 kJ/mol at 298.15 K (-721.0 ± 1.2 kJ/mol at 0 K). The condensed-phase ATcT enthalpy of formation of oxalic acid is -829.7 ± 0.5 kJ/mol, and the resulting sublimation enthalpy is 98.1 ± 1.3 kJ/mol, both at 298.15 K.
- 64Chen, C.; Shyu, S.-F. Conformers and intramolecular hydrogen bonding of the oxalic acid monomer and its anions. Int. J. Quantum Chem. 2000, 76, 541– 551, DOI: 10.1002/(SICI)1097-461X(2000)76:4<541::AID-QUA5>3.0.CO;2-UGoogle ScholarThere is no corresponding record for this reference.
- 65Van Alsenoy, C.; Klimkowski, V. J.; Schafer, L. Ab initio studies of structural features not easily amenable to experiment: Part 37. Structural and conformational investigations of the dicarbonyls glyoxal, biacetyl and oxalic acid. J. Mol. Struct.: THEOCHEM 1984, 109, 321– 330, DOI: 10.1016/0166-1280(84)80016-0Google ScholarThere is no corresponding record for this reference.
- 66Mohajeri, A.; Shakerin, N. The gas-phase acidity and intramolecular hydrogen bonding in oxalic acid. J. Mol. Struct.: THEOCHEM 2004, 711, 167, DOI: 10.1016/j.theochem.2004.10.002Google Scholar66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXhtVClsLnF&md5=366f66e9144c2ffccbd139868ad02bf9The gas-phase acidity and intramolecular hydrogen bonding in oxalic acidMohajeri, A.; Shakerin, N.Journal of Molecular Structure: THEOCHEM (2004), 711 (1-3), 167-172CODEN: THEODJ; ISSN:0166-1280. (Elsevier B.V.)Ab initio calcns. at MP2/6-311++G and G2 level were carried out to study the structures and the gas-phase acidities of different conformations in oxalic acid. The gas-phase acidities (free energies of first ionization reactions) at G2 level vary from 1297.27 kJ/mol for the most acidic conformer to 1359.53 kJ/mol for the least acidic one at 298.15 K. Analyzing intramol. hydrogen bonding in different conformers, shows that the gas-phase acidity has a direct relation with hydrogen bond stabilization of both acid and its conjugated base. The results indicate that the reaction in which the acidic hydrogen is less involved and the hydrogen of its conjugated base is more involved in the intramol. hydrogen bond is the easiest deprotonation reaction. The atoms in mol. (AIM) theory has been also used to obtain the electronic charge distribution and some other related properties.
- 67Hermida-Ramón, J. M.; Cabaleiro-Lago, E. M.; Rodríguez-Otero, J. Computational study of the dissociation of oxalic acid in water clusters. Chem. Phys. 2004, 302, 53– 60, DOI: 10.1016/j.chemphys.2004.02.021Google ScholarThere is no corresponding record for this reference.
- 68Chang, J. G.; Chen, H. T.; Xu, S. C.; Lin, M. C. Computational Study on the Kinetics and Mechanisms for the Unimolecular Decomposition of Formic and Oxalic Acids. J. Phys. Chem. A 2007, 111, 6789, DOI: 10.1021/jp069036pGoogle Scholar68https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXisVaisLo%253D&md5=bc140b40de3da68ce41189a5e3e6eb40Computational Study on the Kinetics and Mechanisms for the Unimolecular Decomposition of Formic and Oxalic AcidsChang, Jee-Gong; Chen, Hsin-Tsung; Xu, Shucheng; Lin, M. C.Journal of Physical Chemistry A (2007), 111 (29), 6789-6797CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The kinetics and mechanisms for the unimol. decompn. reactions of formic acid and oxalic acid have been studied computationally by the high-level G2M(CC1) method and microcanonical RRKM theory. There are two reaction pathways in the decompn. of formic acid: The dehydration process starting from the Z conformer is found to be the dominant, whereas the decarboxylation reaction starting from the E conformer is less competitive. The predicted rate consts. for the dehydration and decarboxylation reactions are in good agreement with the exptl. data. The calcd. CO/CO2 ratio, 13.6-13.9 between 1300 and 2000 K, is in close agreement with the ratio of 10 measured exptl. by Hsu et al. (In The 19th International Symposium on Combustion; The Combustion Institute: Pittsburgh, PA, 1983; p 89). For oxalic acid, its isomer with two intramol. hydrogen bonds is the most stable structure, similar to earlier reports. Two primary decompn. channels of oxalic acid producing CO2 + HOCOH with barriers of 33-36 kcal/mol and CO2 + CO + H2O with a barrier of 39 kcal/mol were found. At high temps., the latter process becomes more competitive. The rate const. predicted for the formation of CO2 and HOCOH (the precursor of HCOOH) agrees well with available exptl. data. The mechanism for the isomerization of HOCOH to HCOOH is also discussed.
- 69Higgins, J.; Zhou, X.; Liu, R.; Huang, T. T.-S. Theoretical Study of Thermal Decomposition Mechanism of Oxalic Acid. J. Phys. Chem. A 1997, 101, 2702– 2708, DOI: 10.1021/jp9638191Google Scholar69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXhslyqsro%253D&md5=2bb21e96f7d426594288668a58f46fcfTheoretical Study of Thermal Decomposition Mechanism of Oxalic AcidHiggins, James; Zhou, Xuefeng; Liu, Ruifeng; Huang, Thomas T.-S.Journal of Physical Chemistry A (1997), 101 (14), 2702-2708CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)D. functional theory B3LYP/6-31G** and ab initio MP2/6-31G** and MP4(SDQ)/6-311++G** calcns. were carried out to study the structures and isomerization and decompn. mechanisms of oxalic acid. The B3LYP structures and relative energies of the rotational isomers of oxalic acid are found very similar to MP2 results, confirming that the most stable isomer is the doubly intramol. hydrogen-bonded C2h structure E1, with four other planar isomers within 6 kcal/mol. It is predicted that unimol. formation of carbon dioxide and dihydroxycarbene (DHC) from oxalic acid has an activation barrier of 31 kcal/mol and that unimol. formation of HCOOH from DHC has an activation barrier about 31 kcal/mol higher. The unimol. formation of CO2, CO and H2O from oxalic acid via a concerted transition state has an activation barrier of only 42 kcal/mol, indicating it is a more favorable unimol. decompn. channel. On the other hand, hydrogen migration from oxygen to carbon of DHC to produce HCOOH can be accomplished through a hydrogen exchange with H2O (a model for oxalic acid) with an activation barrier of less than 37 kcal/mol. Transition state theory calcns. indicate that this bimol. channel might be responsible for the rapid formation of CO2 and HCOOH in gas phase oxalic acid thermal decompn., thus confirming the proposal of Bock and Redington. With increasing temp. the unimol. channel to produce CO2, CO, and H2O might become significant.
- 70Medeiros, F. S.; Oliveira, K. M. T.; Canuto, S.; Chaudhuri, P. A quantum chemical investigation of the interaction of perfluoropropionic acid with monoethanolamine and sulfuric acid in the atmosphere. Comput. Theor. Chem. 2024, 1233, 114485, DOI: 10.1016/j.comptc.2024.114485Google ScholarThere is no corresponding record for this reference.
- 71de Oliveira, T. S.; Ghosh, A.; Chaudhuri, P. Exploring the hydrogen-bonded interactions of vanillic acid with atmospheric bases: a DFT study. Struct. Chem. 2024, 35, 1601– 1611, DOI: 10.1007/s11224-024-02307-3Google ScholarThere is no corresponding record for this reference.
- 72Elm, J.; Mikkelsen, K. V. Computational approaches for efficiently modelling of small atmospheric clusters. Chem. Phys. Lett. 2014, 615, 26– 29, DOI: 10.1016/j.cplett.2014.09.060Google Scholar72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhs12ntL3F&md5=087d4f5b68da4e88ac9ae535a1bcb826Computational approaches for efficiently modelling of small atmospheric clustersElm, Jonas; Mikkelsen, Kurt V.Chemical Physics Letters (2014), 615 (), 26-29CODEN: CHPLBC; ISSN:0009-2614. (Elsevier B.V.)Utilizing a comprehensive test set of 205 clusters of atm. relevance, we investigate how different DFT functionals (M06-2X, PW91, ωB97X-D) and basis sets (6-311++G(3df,3pd), 6-31++G(d,p), 6-31+G(d)) affect the thermal contribution to the Gibbs free energy and single point energy. Reducing the basis set used in the geometry and frequency calcn. from 6-311++G(3df,3pd) → 6-31++G(d,p) implies a significant speed-up in computational time and only leads to small errors in the thermal contribution to the Gibbs free energy and subsequent coupled cluster single point energy calcn.
- 73Elm, J.; Bilde, M.; Mikkelsen, K. V. Assessment of binding energies of atmospherically relevant clusters. Phys. Chem. Chem. Phys. 2013, 15, 16442– 16445, DOI: 10.1039/c3cp52616jGoogle Scholar73https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsVCru7jP&md5=64f46480e82da3faf993770c931335feAssessment of binding energies of atmospherically relevant clustersElm, Jonas; Bilde, Merete; Mikkelsen, Kurt V.Physical Chemistry Chemical Physics (2013), 15 (39), 16442-16445CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)This work assesses the binding energies of atmospherically relevant clusters contg. H2SO4, H2O, NH3 and (CH3)2NH using d. functional theory. The performance of seven DFT functionals (B3LYP, CAM-B3LYP, M06-2X, PW91, LC-PW91, PBE0 and ωB97X-D) is evaluated against high level explicitly correlated coupled cluster methods using a test set of 107 atmospherically relevant clusters. Our studies show that all the tested functionals correlate well with the coupled cluster results, but with highly varying mean abs. errors. The PBE0, CAM-B3LYP, PW91 and M06-2X functionals are found to perform similarly with errors in the range of 2.53-3.46 kcal mol-1, while the B3LYP and LC-PW91 functionals yield higher errors of 6.95 kcal mol-1 and 10.66 kcal mol-1, resp. The ωB97X-D functional gives the best est. of the binding energies with a mean abs. error as low as 2.12 kcal mol-1 over the large test set of clusters.
- 74Elm, J.; Bilde, M.; Mikkelsen, K. V. Assessment of Density Functional Theory in Predicting Structures and Free Energies of Reaction of Atmospheric Prenucleation Clusters. Phys. Chem. Chem. Phys. 2012, 8, 2071– 2077, DOI: 10.1021/ct300192pGoogle ScholarThere is no corresponding record for this reference.
- 75Huff, A. K.; Mackenzie, R. B.; Smith, C. J.; Leopold, K. R. A Perfluorinated Carboxylic Sulfuric Anhydride: Microwave and Computational Studies of CF3COOSO2OH. J. Phys. Chem. A 2019, 123, 2237– 2243, DOI: 10.1021/acs.jpca.9b00300Google Scholar75https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXitFGht70%253D&md5=6f9d1a1c8867656a8507b8e9b9594c89A Perfluorinated Carboxylic Sulfuric Anhydride: Microwave and Computational Studies of CF3COOSO2OHHuff, Anna K.; MacKenzie, Rebecca B.; Smith, C. J.; Leopold, Kenneth R.Journal of Physical Chemistry A (2019), 123 (11), 2237-2243CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Trifluoroacetic sulfuric anhydride (CF3COOSO2OH, TFASA) and its deuterated isotopologue have been obsd. by pulsed-nozzle Fourier transform microwave spectroscopy. TFASA was generated in situ in a supersonic expansion from the reaction of CF3COOH or CF3COOD with SO3. The spectrum, which was notably weaker than those of previously studied carboxylic sulfuric anhydrides, is that of a simple asym. rotor with no evidence of internal rotation of the CF3 group. Calcns. at the M06-2X/6-311++G(3df,3pd) level indicate that the title compd. is produced via a mechanism involving a concerted cycloaddn., analogous to that found for other carboxylic sulfuric anhydrides. The calcns. further show that the equil. orientation of CF3 relative to the C=O bond changes upon formation of the anhydride, indicating that any path connecting the equil. structures of CF3COOH and CF3COOSO2OH necessarily includes both cycloaddn. and internal rotation. CCSD(T)/complete basis set with double and triple extrapolation [CBS(D-T)] single-point energy calcns. at key points on the potential surface indicate that the barrier to form TFASA from a putative CF3COOH···SO3 complex is about 1.2 kcal/mol after zero-point energy corrections. This value is significantly larger than the near-zero or slightly neg. barriers previously reported for the reactions of SO3 with nonfluorinated carboxylic acids and likely accounts, at least in part, for the reduced spectral intensity. Thus, TFASA is a somewhat unique addn. to the series of carboxylic sulfuric anhydrides studied to date. Theor. values of certain structural parameters, at. charges, and vibrational frequencies also support this point of view. Despite the differences, however, this work clearly demonstrates that the reaction RCOOH + SO3 → RCOOSO2OH readily occurs in the gas phase and is not restricted to acids with hydrocarbon R groups.
- 76Dennington, R. D.; Keith, T. A.; Millam, J. M. GaussView 5.0.9; Gaussian Inc.: Wallingford, CT, USA, 2008.Google ScholarThere is no corresponding record for this reference.
- 77Kubečka, J.; Besel, V.; Kurtén, T.; Myllys, N.; Vehkamäki, H. Configurational sampling of noncovalent (atmospheric) molecular clusters: Sulfuric acid and guanidine. J. Phys. Chem. A 2019, 123, 6022– 6033, DOI: 10.1021/acs.jpca.9b03853Google Scholar77https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXht1eks7fL&md5=941e3ebf795469187f47126a4c65db3cConfigurational Sampling of Noncovalent (Atmospheric) Molecular Clusters: Sulfuric Acid and GuanidineKubecka, Jakub; Besel, Vitus; Kurten, Theo; Myllys, Nanna; Vehkamaki, HannaJournal of Physical Chemistry A (2019), 123 (28), 6022-6033CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)We studied the configurational sampling of noncovalently bonded mol. clusters relevant to the atm. In this article, we discuss possible approaches to searching for optimal configurations and present one alternative based on systematic configurational sampling, which seems able to overcome the typical problems assocd. with searching for global min. on multidimensional potential energy surfaces. Since atm. mol. clusters are usually held together by intermol. bonds, we also present a cost-effective strategy for treating hydrogen bonding and proton transferring by using rigid mols. and ions in different protonation states and illustrate its performance on clusters contg. guanidine and sulfuric acid.
- 78Partanen, L.; Vehkamäki, H.; Hansen, K.; Elm, J.; Henschel, H.; Kurtén, T.; Halonen, R.; Zapadinsky, E. Effect of Conformers on Free Energies of Atmospheric Complexes. J. Phys. Chem. A 2016, 120, 8613– 8624, DOI: 10.1021/acs.jpca.6b04452Google Scholar78https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xhs1SksbfM&md5=1390239ffe63c609adfbfd2561f5f63fEffect of Conformers on Free Energies of Atmospheric ComplexesPartanen, Lauri; Vehkamaki, Hanna; Hansen, Klavs; Elm, Jonas; Henschel, Henning; Kurten, Theo; Halonen, Roope; Zapadinsky, EvgeniJournal of Physical Chemistry A (2016), 120 (43), 8613-8624CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)In this article we show how to calc. free energies for atmospherically relevant complexes when multiple conformers and/or isomers are present. We explain why the thermal averaging methods used in several published works are not correct. Based on our two sample cases, the sulfuric acid - pinic acid complex and the (H2SO4)3(NH3)3(H2O)4 cluster, we provide numerical evidence that the use of these incorrect formulas can result in errors larger than one kcal/mol. We recommend that if vibrational frequencies and thus Gibbs free energies of the individual conformers are unavailable, one should not attempt to correct for the presence of multiple conformers and instead use only the global min. conformers for both reactants and products. On the other hand, if the free energies for the conformers are calcd. for both reactants and products, their effect can be accounted for by the statistical mech. methods presented in this article.
- 79Fileti, E. E.; Rivelino, R.; Canuto, S. Rayleigh light scattering of hydrogen bonded clusters investigated by means of ab initio calculations. J. Phys. B: At., Mol. Opt. Phys. 2003, 36, 399– 408, DOI: 10.1088/0953-4075/36/2/319Google Scholar79https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXntFGjtrs%253D&md5=5dbaf9c98e81fd7b512c06efdf3875b8Rayleigh light scattering of hydrogen bonded clusters investigated by means of ab initio calculationsFileti, Eudes E.; Rivelino, Roberto; Canuto, SylvioJournal of Physics B: Atomic, Molecular and Optical Physics (2003), 36 (2), 399-408CODEN: JPAPEH; ISSN:0953-4075. (Institute of Physics Publishing)Ab initio calcns. of depolarization ratios and intensities of classically scattered light, in terms of dipole polarizabilities and polarizability anisotropies, are reported for different H bonded mol. clusters. Five different groups of org. heterodimers formed with H2O are considered: HCHO···H2O, MeHO···H2O, HCOOH···H2O, MeCN···H2O, and Me2CO···H2O, together with the H2O dimer H2O···H2O. The geometries of all complexes were optimized by the 2nd-order Moller-Plesset many-body perturbation theory (MP2), using the augmented correlation-consistent basis set with polarized valence of double-zeta quality (aug-cc-pVDZ). The calcd. av. dipole polarizabilities of the isolated mols. are in good agreement with available exptl. results. The calcns. are then extended to the complexes and, from these, the Rayleigh scattering activities and depolarization ratio changes, upon H bond formation, are obtained and analyzed. The differences in activity and depolarization for Rayleigh scattered radiation between 2 groups of isomers, (i) HCN···H2O and H2O···HCN and (ii) MeHO···H2O and MeOH···OH2, also were studied.
- 80Chaudhuri, P.; Canuto, S. Rayleigh scattering properties of small polyglycine molecules. J. Mol. Struct. 2006, 760, 15– 20, DOI: 10.1016/j.theochem.2005.10.039Google Scholar80https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xit1WrsLo%253D&md5=8b44aec03394ddf333e2902da7d286ffRayleigh scattering properties of small polyglycine moleculesChaudhuri, Puspitapallab; Canuto, SylvioJournal of Molecular Structure: THEOCHEM (2006), 760 (1-3), 15-20CODEN: THEODJ; ISSN:0166-1280. (Elsevier B.V.)D.-functional theory calcns. of the Rayleigh optical activities of small isolated polyglycine mols. are reported. Fully extended β-sheet-like conformations of polypeptides of glycine, (Gly)n (with n = 1-5) are considered. After geometry optimizations, dipole moments and dipole polarizabilities (both the mean and the anisotropic components) are calcd. using the B3LYP and B3P86 functionals in 3 basis sets. The polarizabilities are used to analyze the Rayleigh scattering activities and depolarization ratios. The convergence of the av. dipole polarizability per monomer is analyzed. The differences in activity and depolarization for Rayleigh scattered radiation between the extended β-sheet-like and the folded α-helix-like forms of tetraglycine are analyzed and are relevant, suggesting its possible use in exptl. characterization.
- 81da Silva, A. M.; Chakraborty, S.; Chaudhuri, P. Rayleigh light scattering from hydrogen-bonded dimers of small astrophysical molecules. Int. J. Quantum Chem. 2012, 112, 2822– 2827, DOI: 10.1002/qua.23303Google ScholarThere is no corresponding record for this reference.
- 82Elm, J.; Norman, P.; Bilde, M.; Mikkelsen, K. V. Computational study of the Rayleigh light scattering properties of atmospheric prenucleation clusters. Phys. Chem. Chem. Phys. 2014, 16, 10883– 10890, DOI: 10.1039/C4CP01206BGoogle ScholarThere is no corresponding record for this reference.
- 83Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H., Gaussian 16, Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2016.Google ScholarThere is no corresponding record for this reference.
- 84Elm, J.; Ayoubi, D.; Engsvang, M.; Jensen, A. B.; Knattrup, Y.; Kubečka, J.; Bready, C. J.; Fowler, V. R.; Harold, S. E.; Longsworth, O. M. Quantum chemical modeling of organic enhanced atmospheric nucleation: A critical review. WIREs Comput. Mol. Sci. 2023, 13 (5), e1662 DOI: 10.1002/wcms.1662Google ScholarThere is no corresponding record for this reference.
Cited By
This article has not yet been cited by other publications.
Article Views
Altmetric
Citations
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
References
This article references 84 other publications.
- 1Huang, X.-F.; Hu, M.; He, L.-Y.; Tang, X.-Y. Chemical characterization of water-soluble organic acids in PM in Beijing, China. Atmos. Environ. 2005, 39, 2819– 2827, DOI: 10.1016/j.atmosenv.2004.08.038There is no corresponding record for this reference.
- 2Wagner, R.; Möhler, O.; Saathoff, H.; Schnaiter, M.; Leisner, T. High variability of the heterogeneous ice nucleation potential of oxalic acid dihydrate and sodium oxalate. Atmos. Chem. Phys. 2010, 10, 7617– 7641, DOI: 10.5194/acp-10-7617-2010There is no corresponding record for this reference.
- 3Xu, Y.; Nadykto, A. B.; Yu, F.; Jiang, L.; Wang, W. Formation and properties of hydrogen-bonded complexes of common organic oxalic acid with atmospheric nucleation precursors. J. Mol. Struct.: THEOCHEM 2010, 951, 28– 33, DOI: 10.1016/j.theochem.2010.04.0043https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXmtlKlt7o%253D&md5=76c544813ec5a3150586367e7398b9c1Formation and properties of hydrogen-bonded complexes of common organic oxalic acid with atmospheric nucleation precursorsXu, Yisheng; Nadykto, Alexey B.; Yu, Fangqun; Jiang, Lei; Wang, WeiJournal of Molecular Structure: THEOCHEM (2010), 951 (1-3), 28-33CODEN: THEODJ; ISSN:0166-1280. (Elsevier B.V.)Nucleation in the Earth's atm. plays a vital role in the aerosol radiative forcing assocd. directly with the prodn. of ultrafine particles responsible for adverse public health impacts and global climate changes. The interaction of the common org. oxalic acid with the atm. nucleation precursors and trace ionic species has been investigated using the D. Functional Theory (DFT). A comprehensive study of the hydrogen-bonded complexes of oxalic acid with neutral monomers and ionic clusters of the sulfuric acid, water and ammonia, the key atm. nucleation precursors, has been carried out. Bonding of the oxalic acid with the above-mentioned neutrals was found to be moderately weak, and thus, the oxalic acid cannot stabilize the binary sulfuric acid-water and ternary sulfuric acid-ammonia clusters. The interaction of the oxalic acid with pos. ionic species is, in contrast, strong and significantly enhances stability of the clusters formed over the ions. The hydration of pos. charged (C2H2O4)(H3O+), (C2H2O4)(NH4+) and (C2H2O4)(H3O+)(H2SO4) clusters is also strong and close to that of the sulfuric acid, the key atm. nucleation precursor. These considerations lead to the logical conclusion that oxalic acid can catalyze the prodn. of pos. charged pre-nucleation clusters and, thus, its role in nucleation of pos. ions in the Earth's troposphere should be studied in further details.
- 4Xu, W.; Zhang, R. Theoretical investigation of interaction of dicarboxylic acids with common aerosol nucleation precursors. J. Phys. Chem. A 2012, 116, 4539– 4550, DOI: 10.1021/jp301964u4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XkvFOnu7c%253D&md5=f83fc23ed930da24fdb0fd0719292623Theoretical Investigation of Interaction of Dicarboxylic Acids with Common Aerosol Nucleation PrecursorsXu, Wen; Zhang, RenyiJournal of Physical Chemistry A (2012), 116 (18), 4539-4550CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Dicarboxylic acids are important products from photooxidn. of volatile org. compds. and are believed to play an important role in the formation and growth of atm. secondary org. aerosols. In this paper, the interaction of five dicarboxylic acids, i.e., oxalic acid (C2H2O4), malonic acid (C3H4O4), maleic acid (C4H4O4), phthalic acid (C8H6O4), and succinic acid (C4H6O4), with sulfuric acid and ammonia has been studied, employing quantum chem. calcns., quantum theory of atoms in mols. (QTAIM), and the natural bond orbital (NBO) anal. methods. Several levels of quantum chem. calcns. are considered, including coupled-cluster theory with single and double excitations with perturbative corrections for the triple excitations (CCSD(T)) and two d. functionals, B3LYP and PW91PW91. The free energies of formation of the heterodimer and heterotrimer clusters suggest that dicarboxylic acids can contribute to the aerosol nucleation process by binding to sulfuric acid and ammonia. In particular, the formation energies and structures of the heterotrimer clusters show that dicarboxylic acids enhance nucleation in two directions, in contrast to monocarboxylic acids.
- 5Weber, K. H.; Morales, F. J.; Tao, F. Theoretical study on the structure and stabilities of molecular clusters of oxalic acid with water. J. Phys. Chem. A 2012, 116, 11601– 11617, DOI: 10.1021/jp308499f5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsFCrurrJ&md5=5f1a7fac54947abdbade8bb7b9ba4b5aTheoretical Study on the Structure and Stabilities of Molecular Clusters of Oxalic Acid with WaterWeber, Kevin H.; Morales, Francisco J.; Tao, Fu-MingJournal of Physical Chemistry A (2012), 116 (47), 11601-11617CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The importance of aerosols to humankind is well-known, playing an integral role in detg. Earth's climate and influencing human health. Despite this fact, much remains unknown about the initial events of nucleation. In this work, the mol. properties of common org. atm. pollutant oxalic acid and its gas phase interactions with water have been thoroughly examd. Local min. single-point energies for the monomer conformations were calcd. at the B3LYP and MP2 level of theory with both 6-311++G(d,p) and aug-cc-pVDZ basis sets and are compared with previous works. Optimized geometries, relative energies, and free energy changes for the stable clusters of oxalic acid conformers with up to six waters were then obtained from B3LYP calcns. with 6-31+G(d) and 6-311++G(d,p) basis sets. Initially, cooperative binding is predicted to be the most important factor in nucleation, but as the clusters grow, dipole cancellations are found to play a pivotal role. The clusters of oxalic acid hydrated purely with water tend to produce extremely stable and neutral core systems. Free energies of formation and atm. implications are discussed.
- 6Weber, K. H.; Liu, Q.; Tao, F.-M. Theoretical study on stable small clusters of oxalic acid with ammonia and water. J. Phys. Chem. A 2014, 118, 1451– 1468, DOI: 10.1021/jp41282266https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXht12lsrs%253D&md5=d06cb1431aa9546ee0ce4c56d193d850Theoretical Study on Stable Small Clusters of Oxalic Acid with Ammonia and WaterWeber, Kevin H.; Liu, Qian; Tao, Fu-MingJournal of Physical Chemistry A (2014), 118 (8), 1451-1468CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Thermodynamically stable small clusters of oxalic acid (CO2H)2, ammonia (NH3), and water (H2O) are studied through quantum chem. calcns. The (CO2H)2-NH3 core system with up to three waters of hydration was examd. by B3LYP d. functional theory and MP2 MO theory with the aug-cc-pVDZ basis set. The (CO2H)2-NH3 core complexes are obsd. to hydrogen bond strongly and should be found in appreciably significant concns. in the atm. Subsequent hydration of the (CO2H)2-NH3 core, however, is found to be somewhat prohibitive under ambient conditions. Relative populations of the examd. clusters are predicted and the binding patterns detailed. Atm. implications related to new particle formations are discussed.
- 7Peng, X.-Q.; Liu, Y.-R.; Huang, T.; Jiang, S.; Huang, W. Interaction of gas phase oxalic acid with ammonia and its atmospheric implications. Phys. Chem. Chem. Phys. 2015, 17, 9552– 9563, DOI: 10.1039/C5CP00027K7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXjtlWgtbc%253D&md5=eb6efa7d64ab2c0d30464a57cae4b0a6Interaction of gas phase oxalic acid with ammonia and its atmospheric implicationsPeng, Xiu-Qiu; Liu, Yi-Rong; Huang, Teng; Jiang, Shuai; Huang, WeiPhysical Chemistry Chemical Physics (2015), 17 (14), 9552-9563CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Oxalic acid is believed to play an important role in the formation and growth of atm. org. aerosols. However, as a common org. acid, the understanding of the larger clusters formed by gas phase oxalic acid with multiple ammonia mols. is incomplete. In this work, the structural characteristics and thermodn. of oxalic acid clusters with up to six ammonia mols. have been investigated at the PW91PW91/6-311++G(3df,3pd) level of theory. We found that oxalic acid forms relatively stable clusters with ammonia mols., and that ionization events play a key role. The analyses of the thermodn. and atm. relevance indicate that the heterodimer (H2C2O4)(NH3) shows an obvious relative concn. in the atm., and thus likely participates in new particle formation. However, with increasing no. of ammonia mols., the concn. of clusters decreases gradually. Addnl., clusters of oxalic acid with ammonia mols. are predicted to form favorably in low temp. conditions and show high Rayleigh scattering intensities.
- 8Miao, S.-K.; Jiang, S.; Chen, J.; Ma, Y.; Zhu, Y.-P.; Wen, Y. Hydration of a sulfuric acid-oxalic acid complex: acid dissociation and its atmospheric implication. RSC Adv. 2015, 5, 48638– 48646, DOI: 10.1039/C5RA06116D8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXptFClu7c%253D&md5=867367bf4afb469541b384519cc4ea02Hydration of a sulfuric acid-oxalic acid complex: acid dissociation and its atmospheric implicationMiao, Shou-Kui; Jiang, Shuai; Chen, Jiao; Ma, Yan; Zhu, Yu-Peng; Wen, Yang; Zhang, Miao-Miao; Huang, WeiRSC Advances (2015), 5 (60), 48638-48646CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Oxalic acid (OA), one of the most common org. acids in the Earth's atm., is expected to enhance the nucleation and growth of nanoparticles contg. sulfuric acid (SA) and water (W); however, the details about the hydration of OA-SA are poorly understood, esp. for the larger clusters with more water mols. We have investigated the structural characteristics and thermodn. of these clusters using d. functional theory at the PW91PW91/6-311++G(3df,3pd) level. The favorable free energies of formation and obvious concns. of the OA-SA-Wn (n = 0-6) clusters at 298.15 K predict that oxalic acid can contribute to the aerosol nucleation process by binding to sulfuric acid and water until n = 6. There is strong temp. dependence for the complexes formation, and the energy order of these complexes is altered from 100 to 400 K, regardless of different cluster sizes or different isomers within the same cluster size. The lower temp. and higher relative humidity promote the formation of hydrates. Addnl., the investigation of acid dissocn. predicts that several acid-dissocd. models could coexist in the atm., specifically when more water mols. are present. Fewer waters may be needed to cause the acid dissocn., as the relative acidity of the cluster increases, which plays a key role in forming relatively stable hydrated clusters of OA-SA. Finally, the Rayleigh scattering properties of OA-SA-Wn (n = 0-6) have been systematically investigated for the first time to further discuss its atm. implication.
- 9Pavuluri, C. M.; Kawamura, K.; Mihalopoulos, N.; Swaminathan, T. Laboratory photochemical processing of aqueous aerosols: Formation and degradation of dicarboxylic acids, oxocarboxylic acids and α- dicarbonyls. Atmos. Chem. Phys. 2015, 15, 7999– 8012, DOI: 10.5194/acp-15-7999-2015There is no corresponding record for this reference.
- 10Wang, G.; Cheng, C.; Meng, J.; Huang, Y.; Li, J.; Ren, Y. Field observation on secondary organic aerosols during Asian dust storm periods: Formation mechanism of oxalic acid and related compounds on dust surface. Atmos. Environ. 2015, 113, 169– 176, DOI: 10.1016/j.atmosenv.2015.05.01310https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXosVKqur0%253D&md5=41d96574d82a771f739b414cde6dcf2bField observation on secondary organic aerosols during Asian dust storm periods: Formation mechanism of oxalic acid and related compounds on dust surfaceWang, Gehui; Cheng, Chunlei; Meng, Jingjing; Huang, Yao; Li, Jianjun; Ren, YanqinAtmospheric Environment (2015), 113 (), 169-176CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)Chem. evolution of East Asian dust during transpacific transport has been given much attention for inorg. species such as sulfate, nitrate and ammonium. However, the role of org. species during the transport has almost entirely been ignored. To understand the formation mechanism of secondary org. aerosols (SOA) on dust surfaces, this study investigated the concns. and compns. of dicarboxylic acids, keto-carboxylic acids, α-dicarbonyls and inorg. ions in size-segregated aerosols (9-stages) collected in Xi'an, central China during the two dust storm episodes in the springs of 2009 and 2011 and compared with those in nondust storm periods. During the events the ambient particulate dicarboxylic acids were 932-2240 ng m-3, which are comparable and even higher than those in nondust periods. Mol. compns. of the above SOA are similar to those in nondust periods with oxalic acid being the leading species. In the presence of the dust storms, all the above mentioned SOA species in Xi'an were predominantly enriched on the coarse particles (>2.1 μm), and oxalic acid well correlated with NO3- (R2 = 0.72, p < 0.001) rather than SO42-. This phenomenon differs greatly from the SOA in any other nondust period that is usually characterized by an enrichment of oxalic acid in fine mode and a strong correlation of oxalic acid with SO42-. We propose a formation pathway to explain these observations, in which nitric acid and/or nitrogen oxides react with dust to produce Ca(NO3)2 and form a liq. phase on the surface of dust aerosols via water vapor-absorption of Ca(NO3)2, followed by a partitioning of the gas-phase water-sol. org. precursors (e.g.,glyoxal and methylglyoxal) into the aq.-phase and a subsequent oxidn. into oxalic acid. To the best of our knowledge, we found for the first time the enrichment of glyoxal and methylglyoxal on dust surface. Our data suggest an important role of nitrate in the heterogeneous formation process of SOA on the surface of dust.
- 11Kawamura, K.; Bikkina, S. A review of dicarboxylic acids and related compounds in atmospheric424 aerosols: Molecular distributions, sources and transformation. Atmos. Res. 2016, 170, 140– 160, DOI: 10.1016/j.atmosres.2015.11.01811https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXitVSqt7vF&md5=fcba54277b4abd2ddc9092cefe8141f6A review of dicarboxylic acids and related compounds in atmospheric aerosols: Molecular distributions, sources and transformationKawamura, Kimitaka; Bikkina, SrinivasAtmospheric Research (2016), 170 (), 140-160CODEN: ATREEW; ISSN:0169-8095. (Elsevier B.V.)This review aims to update our understanding on mol. distributions of water-sol. dicarboxylic acids and related compds. in atm. aerosols with a focus on their geog. variability, size distribution, sources and formation pathways. In general, mol. distributions of diacids in aerosols from the continental sites and over the open ocean waters are often characterized by the predominance of oxalic acid (C2) followed by malonic acid (C3) and/or succinic acid (C4), while those sampled over the polar regions often follow the order of C4 ≥ C2 and C3. The most abundant and ubiquitous diacid is oxalic acid, which is principally formed via atm. oxidn. of its higher homologues of long chain diacids and other pollution-derived org. precursors (e.g., olefins and arom. hydrocarbons). However, its occurrence in marine aerosols is mainly due to the transport from continental outflows (e.g., East Asian outflow during winter/spring to the North Pacific) and/or governed by photochem./aq. phase oxidn. of biogenic unsatd. fatty acids (e.g., oleic acid) and isoprene emitted from the productive open ocean waters. The long-range atm. transport of pollutants from mid latitudes to the Arctic in dark winter facilitates to accumulate the reactants prior to their intense photochem. oxidn. during springtime polar sunrise. Furthermore, the relative abundances of C2 in total diacid mass showed similar temporal trends with downward solar irradn. and ambient temps., suggesting the significance of atm. photochem. oxidn. processing. Compd.-specific isotopic analyses of oxalic acid showed the highest δ13C among diacids whereas azelaic acid showed the lowest value, corroborating the significance of atm. aging of oxalic acid. On the other hand, other diacids gave intermediate values between these two diacids, suggesting that aging of oxalic acid is assocd. with 13C enrichment.
- 12Peng, X.-Q.; Huang, T.; Miao, S.-K.; Chen, J.; Wen, H.; Feng, Y.-J. Hydration of oxalic acid–ammonia complex: atmospheric implication and Rayleigh-scattering properties. RSC Adv. 2016, 6, 46582– 47593, DOI: 10.1039/C6RA03164AThere is no corresponding record for this reference.
- 13Zhao, H.; Zhang, Q.; Du, L. Hydrogen bonding in cyclic complexes of carboxylic acid–sulfuric acid and their atmospheric implications. RSC Adv. 2016, 6, 71733– 71743, DOI: 10.1039/C6RA16782AThere is no corresponding record for this reference.
- 14Elm, J.; Myllys, N.; Kurtén, T. What is required for highly oxidized molecules to form clusters with sulfuric acid?. J. Phys. Chem. A 2017, 121, 4578– 4587, DOI: 10.1021/acs.jpca.7b0375914https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXos1amtbY%253D&md5=bca0321a8b08702791e619e491c323daWhat Is Required for Highly Oxidized Molecules To Form Clusters with Sulfuric Acid?Elm, Jonas; Myllys, Nanna; Kurten, TheoJournal of Physical Chemistry A (2017), 121 (23), 4578-4587CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)We have studied the specific requirements of a given neutral org. mol. to act as a stabilizer in sulfuric acid induced new particle formation. Based on an anal. of the reaction Gibbs free energies between simple functional groups and sulfuric acid, carboxylic acid groups are identified to show the strongest hydrogen bonding interaction with sulfuric acid. The free energy assocd. with the hydrogen bonding between sulfuric acid and 14 different carboxylic acids of atm. relevance reveal that the binding strength is very dependent on the ability of sulfuric acid to form an addnl. hydrogen bond via its vacant S-OH group to a γ-carbonyl group in the org. mol. Extending the anal. to monoterpene oxidn. products and further to large dimer esters, we identify the following necessary criteria for a given org. oxidn. product to efficiently stabilize sulfuric acid clustering: (1) weak or no intramol. hydrogen bonds in the isolated monomer; (2) more than two carboxylic acid groups. As a proof of concept we show that these requirements correspond to the docking of a sulfuric acid mol. between two non-interacting carboxylic acid groups in the org. mol. These findings suggests that, for a given org. oxidn. product to participate in the initial steps in new particle formation involving sulfuric acid, very distinct mol. features are required.
- 15Hu, Y.-C.; Zhang, X.-H.; Li, Q.-S.; Zhang, Y.-H.; Li, Z.-S. Effect of water on the structure and stability of hydrogen-bonded oxalic acid dimer. ChemPhyschem 2017, 18, 3375– 3383, DOI: 10.1002/cphc.20170095015https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhsleiu7vK&md5=192a1c69aaf0b74c11242298629f69c3Effect of Water on the Structure and Stability of Hydrogen-Bonded Oxalic Acid DimerHu, Yuan-Chun; Zhang, Xiu-Hui; Li, Quan-Song; Zhang, Yun-Hong; Li, Ze-ShengChemPhysChem (2017), 18 (23), 3375-3383CODEN: CPCHFT; ISSN:1439-4235. (Wiley-VCH Verlag GmbH & Co. KGaA)As the simplest and most abundant dicarboxylic acid in the atm., oxalic acid (OA) not only plays a key role in aerosol nucleation, but also acts as a prototypical compd. for the investigation of intra- and intermol. hydrogen-bonding interactions. A systematic theor. study on the hydrated OA dimers performed by using DFT at the M06-2X/6-311++G(3df, 2p) level is discussed herein. The properties of hydrogen bonds in clusters are inspected through topol. anal. by using atoms in mols. (AIM) theory. The most stable OA dimer involves a cyclic structure with two intermol. hydrogen bonds. Calcns. show that one H2O has a slight effect on the hydrogen bonds, whereas two water mols. weaken and three water mols. break the two intermol. hydrogen bonds between OAs. Furthermore, there are no hydrogen-bond interactions between OAs in almost all stable clusters as the no. of H2O mols. increases to four and five. Addnl., ionization and isomerization of OA through water-assisted proton-transfer phenomena are obsd. in tetra- and pentahydrates. This work provides new insights into the conversion of anhyd. OA into hydrated clusters that are helpful for further understanding the atm. nucleation process and nature of hydrogen bond.
- 16Xu, J.; Finlayson-Pitts, B. J.; Gerber, R. B. Proton transfer in mixed clusters of methanesulfonic acid, methylamine, and oxalic acid: implications for atmospheric particle formation. J. Phys. Chem. A 2017, 121, 2377– 2385, DOI: 10.1021/acs.jpca.7b0122316https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXktV2kurc%253D&md5=63eae62077f46c6a94d7802ebab7b8f1Proton Transfer in Mixed Clusters of Methanesulfonic Acid, Methylamine, and Oxalic Acid: Implications for Atmospheric Particle FormationXu, Jing; Finlayson-Pitts, Barbara J.; Gerber, R. BennyJournal of Physical Chemistry A (2017), 121 (12), 2377-2385CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Understanding the properties of atm. particles made of several components is a very challenging problem. In this paper, we perform quantum chem. calcns. for small multicomponent clusters of atm. relevance that incorporate methanesulfonic acid (MSA), methylamine (MA), oxalic acid (OxA), and water (H2O). Potential correlations between theor. predictions of proton transfer in the small clusters and findings of recent expts. on formation of particles of detectable sizes (>2 nm) from the same components are studied. It is proposed that proton transfer from the acid to the amine in the 1:1 clusters correlates with expts. on particle formation in systems, such as MSA-MA and MSA-MA-OxA. In the case of OxA + MA, which has been obsd. to give few particles, proton transfer does not occur for the 1:1 cluster but does for the 2:2 cluster. Adding H2O to OxA-MA promotes the occurrence of proton transfer, and corresponding particles are slightly enhanced. The partial charge on the MA component increases by adding OxA or H2O to MSA-MA, which is correlated with enhanced particle formation compared to MSA-MA alone. Ab initio mol. dynamics simulations show that proton transfer at room temp. (T = 298 K) and high temp. (T = 500 K) is little affected compared with the equil. structure (T = 0 K). These results suggest that small cluster calcns. may be useful in predicting the formation of multicomponent particles in the atm.
- 17Chen, J.; Jiang, S.; Liu, Y.-R.; Huang, T.; Wang, C.-Y.; Miao, S.-K. Interaction of oxalic acid with dimethylamine and its atmospheric implications. RSC Adv. 2017, 7, 6374– 6388, DOI: 10.1039/C6RA27945G17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXht1amur4%253D&md5=9706a306eb06cfcc2afeadce68da99fcInteraction of oxalic acid with dimethylamine and its atmospheric implicationsChen, Jiao; Jiang, Shuai; Liu, Yi-Rong; Huang, Teng; Wang, Chun-Yu; Miao, Shou-Kui; Wang, Zhong-Quan; Zhang, Yang; Huang, WeiRSC Advances (2017), 7 (11), 6374-6388CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Oxalic acid and dimethylamine are the most common org. acid and base in the atm., and are recognized as significant precursor species in atm. new particle formation. However, the interaction between oxalic acid and dimethylamine in the presence of hydration is not yet understood. In this study, the most stable geometric structures and thermodn. of (C2H2O4)m(CH3NHCH3)(H2O)n (m = 1-2, n = 0-4) clusters are investigated using M06-2X coupled with the 6-311+G(2d,p) basis set. A high level explicitly cor. CCSD(T)-F12/VDZ-F12 method is utilized to benchmark the d. functional theory (DFT) methods. Hydration promotes proton transfer from oxalic acid to dimethylamine for (C2H2O4)(CH3NHCH3)(H2O)n (n = 0-4) clusters, while proton transfer from oxalic acid to dimethylamine occurs without hydration for (C2H2O4)2(CH3NHCH3)(H2O)n (n = 0-4) clusters. With regards to the isomer distribution at the potential energy surface, temp. seems not to be an important parameter, since almost all of the global min. for the investigated size range dominate within the investigated temp. range, except for in the (C2H2O4)m(CH3NHCH3)(H2O)2 clusters. Under atm. conditions, the peak hydration distribution shifts from unhydrated clusters to trihydrates for the (C2H2O4)(CH3NHCH3)(H2O)n (n = 0-4) clusters, while for the (C2H2O4)2(CH3NHCH3)(H2O)n (n = 0-4) clusters, unhydrated clusters clearly dominate the cluster distribution, irresp. of whether the humidity is low or high. Finally, the formation free energies obtained from quantum calcns. are used to calc. the evapn. rates. We find that evapn. of dimethylamine is preferred compared to oxalic acid for the (C2H2O4)(CH3NHCH3)(H2O)n clusters, while the results are reversed for the (C2H2O4)2(CH3NHCH3)(H2O)n clusters.
- 18Arquero, K. D.; Xu, J.; Gerber, B.; Finlayson-Pitts, B. J. Particle formation and growth from oxalic acid, methanesulfonic acid, trimethylamine and water: a combined experimental and theoretical study. Phys. Chem. Chem. Phys. 2017, 19, 28286– 28301, DOI: 10.1039/C7CP04468B18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhs1Crt7fI&md5=d4b35c04200187aab197360823bd56b0Particle formation and growth from oxalic acid, methanesulfonic acid, trimethylamine and water: a combined experimental and theoretical studyArquero, Kristine D.; Xu, Jing; Gerber, R. Benny; Finlayson-Pitts, Barbara J.Physical Chemistry Chemical Physics (2017), 19 (41), 28286-28301CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Atm. particles affect visibility, health, and climate but the mechanisms of their formation from initial clusters and their growth to detectable particles remain largely unknown. Previous studies showed methanesulfonic acid (MSA) reactions with NH3 and amines form particles, a process enhanced by water. Results of a combined exptl./theor. study of oxalic acid (OxA) effect on particle formation and growth from the MSA reaction with trimethylamine (TMA) in the presence/absence of water are reported. Gas phase reactants were mixed in an aerosol flow reactor at 1 atm pressure and 294° K. Particle no. concns. and size distributions were measured as a function of reaction time from 0.8-12 s. The interaction of OxA with TMA with/without water did not lead to significant particle formation. When OxA was present during the MSA reaction with TMA, there was little change (≤2 times more) in particle no. concn., but particles were larger compared vs. the base case (MSA with TMA alone). However, the presence of water with MSA and TMA overwhelmed the OxA effect so that OxA has no significant impact on particle no. concn. or size. Exptl. results suggested MSA hydrate is important for particle formation and growth in the four component OxA/MSA/TMA/water system. Results were compared to earlier studies of the OxA effect on the MSA/methylamine reaction and interpreted based on theor. calcd. properties of small clusters of the components.
- 19Hong, Y.; Liu, Y.-R.; Wen, H.; Miao, S.-K.; Huang, T.; Peng, X.-Q. Interaction of oxalic acid with methylamine and its atmospheric implications. RSC Adv. 2018, 8, 7225– 7234, DOI: 10.1039/C7RA13670F19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXislSmu7k%253D&md5=e17fd3798381e4fd8f7627d7e7be524dInteraction of oxalic acid with methylamine and its atmospheric implicationsHong, Yu; Liu, Yi-Rong; Wen, Hui; Miao, Shou-Kui; Huang, Teng; Peng, Xiu-Qiu; Jiang, Shuai; Feng, Ya-Juan; Huang, WeiRSC Advances (2018), 8 (13), 7225-7234CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Oxalic acid, which is one of the most common dicarboxylic acids, is expected to be an important component of atm. aerosols. However, the contribution of oxalic acid to the generation of new particles is still poorly understood. In this study, the structural characteristics and thermodn. of (C2H2O4)(CH3NH2)n (n = 1-4) were investigated at the PW91PW91/6-311++G(3df,3pd) level of theory. We found that clusters formed by oxalic acid and methylamine are relatively stable, and the more the atoms participating in the formation of a ring-like structure, the more stable is the cluster. In addn., via the anal. of atm. relevance, it can be revealed that clusters of (C2H2O4)(CH3NH2)n (n = 1-4) have a noteworthy concn. in the atm., which indicates that these clusters could be participating in new particle formation. Moreover, by comparison with (H2C2O4)(NH3)n (n = 1-6) species, it can be seen that oxalic acid is more readily bound to methylamine than to ammonia, which promotes nucleation or new particle formation. Finally, the Rayleigh scattering properties of clusters of (C2H2O4)(CH3NH2)n (n = 1-4) were investigated for the first time to det. their atm. implications.
- 20Boreddy, S. K. R.; Kawamura, K. Investigation on the hygroscopicity of oxalic acid and atmospherically relevant oxalate salts under sub- and supersaturated conditions. Environ. Sci.: Processes Impacts 2018, 20, 1069– 1080, DOI: 10.1039/C8EM00053KThere is no corresponding record for this reference.
- 21Lv, G.; Sun, X.; Zhang, C.; Li, M. Understanding the catalytic role of oxalic acid in SO3 hydration to form H2SO4 in the atmosphere. Atmos. Chem. Phys. 2019, 19, 2833– 2844, DOI: 10.5194/acp-19-2833-201921https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXosVahur4%253D&md5=44add63e2ab82d65cb6d2837a2b3d449Understanding the catalytic role of oxalic acid in SO3 hydration to form H2SO4 in the atmosphereLv, Guochun; Sun, Xiaomin; Zhang, Chenxi; Li, MeiAtmospheric Chemistry and Physics (2019), 19 (5), 2833-2844CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)The hydration of SO3 plays an important role in atm. sulfuric acid formation. Some atm. species can be involved in and facilitate the reaction. In this work, using quantum chem. calcns., we show that oxalic acid, the most common dicarboxylic acid in the atm., can effectively catalyze the hydration of SO3. The energy barrier of the SO3 hydration reaction catalyzed by oxalic acid (cTt, tTt, tCt and cCt conformers) is a little higher or less than 1 kcal mol-1, which is lower than the energy barrier of 5.17 kcal mol-1 for water-catalyzed SO3 hydration. Compared with the rates of the SO3 hydration reaction catalyzed by oxalic acid and water, it can be found that in the upper troposphere the OA-catalyzed SO3 hydration can play an important role in promoting SO3 hydration. It leads us to conclude that the involvement of oxalic acid in SO3 hydration to form H2SO4 is significant in the atm.
- 22Deshmukh, D. K.; Kawamura, K.; Gupta, T.; Haque, M. M.; Zhang, Y.-L.; Singh, D. K.; Tsai, Y. I. High Loadings of Water-soluble Oxalic Acid and Related Compounds in PM2.5 Aerosols in Eastern Central India: Influence of Biomass Burning and Photochemical Processing. Aerosol Air Qual. Res. 2019, 19, 2625– 2644, DOI: 10.4209/aaqr.2019.10.0543There is no corresponding record for this reference.
- 23Zhao, Y.; Liu, Y.-R.; Jiang, S.; Huang, T.; Wang, Z.-H.; Xu, C.-X. Volatile organic compounds enhancing sulfuric acid-based ternary homogeneous nucleation: the important role of synergistic effect. Atmos. Environ. 2020, 233, 117609, DOI: 10.1016/j.atmosenv.2020.11760923https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhtVWhs7bI&md5=3e257ecf2eb85da3ef364866615feff2Volatile organic compounds enhancing sulfuric acid-based ternary homogeneous nucleation: The important role of synergistic effectZhao, Yu; Liu, Yi-Rong; Jiang, Shuai; Huang, Teng; Wang, Zi-Hang; Xu, Cai-Xin; Feng, Ya-Juan; Huang, WeiAtmospheric Environment (2020), 233 (), 117609CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)New particle formation (NPF) is an important source of atm. aerosols. Sulfuric acid (SA) and water (W) are recognized as essential participating substances in the nucleation of the atm. In addn., as one of the most common org. acids, oxalic acid (OA) can improve NPF when amine such as methylamine (MA) is present. However, exploring the properties of atm. particles made up of different components is challenging and the role of volatile org. compds. in SA-based ternary homogeneous nucleation is still lacking in research. In this work, the structures and energies of (SA)x(OA)y(MA)z(W)m (0 ≤ x, y, z ≤ 3, 0 ≤ m ≤ 1) are investigated. The results indicate that it is accessible for SA to form clusters with OA and MA mols. through hydrogen bonds and proton transfer interactions. The anal. of non-covalent interactions and proton transfer reveals that ternary nucleation systems have stronger hydrogen bonds and more proton transfers than binary systems to stabilize the clusters. In terms of the thermodn. properties, the Gibbs free energies of the clusters will decrease as the addn. of SA, OA or MA mols. indicating that the synergistic effect of these three substances may be of potential in forming the initial cluster and subsequent growth processes. Moreover, the evapn. rates of clusters show that the synergistic effect of ternary clusters leads to a decrease in evapn. rate which may promote atm. nucleation.
- 24Chen, J. Theoretical analysis of sulfuric acid–dimethylamine–oxalic acid–water clusters and implications for atmospheric cluster formation. RSC Adv. 2022, 12, 22425– 22434, DOI: 10.1039/D2RA03492AThere is no corresponding record for this reference.
- 25Palmieri, F.; Estoppey, A.; House, G. L.; Lohberger, A.; Bindschedler, S.; Chain, P. S. G.; Junier, P. Oxalic acid, a molecule at the crossroads of bacterial-fungal interactions. Adv. Appl. Microbiol. 2019, 106, 49– 77, DOI: 10.1016/bs.aambs.2018.10.00125https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BB3cfmvFamtA%253D%253D&md5=7de8e8c2bf98155b660ecaa870f1ae77Oxalic acid, a molecule at the crossroads of bacterial-fungal interactionsPalmieri Fabio; Estoppey Aislinn; Lohberger Andrea; Bindschedler Saskia; House Geoffrey L; Chain Patrick S G; Junier PilarAdvances in applied microbiology (2019), 106 (), 49-77 ISSN:0065-2164.Oxalic acid is the most ubiquitous and common low molecular weight organic acid produced by living organisms. Oxalic acid is produced by fungi, bacteria, plants, and animals. The aim of this review is to give an overview of current knowledge about the microbial cycling of oxalic acid through ecosystems. Here we review the production and degradation of oxalic acid, as well as its implications in the metabolism for fungi, bacteria, plants, and animals. Indeed, fungi are well known producers of oxalic acid, while bacteria are considered oxalic acid consumers. However, this framework may need to be modified, because the ability of fungi to degrade oxalic acid and the ability of bacteria to produce it, have been poorly investigated. Finally, we will highlight the role of fungi and bacteria in oxalic acid cycling in soil, plant and animal ecosystems.
- 26Rudnick, M. B.; van Veen, J. A.; de Boer, W. Oxalic acid: a signal molecule for fungus-feeding bacteria of the genus Collimonas ?. Environ. Microbiol. Rep. 2015, 7 (5), 709– 714, DOI: 10.1111/1758-2229.12290There is no corresponding record for this reference.
- 27Hsieh, L.; Chen, C.; Wan, M.; Tsai, C.; Tsai, Y. Speciation and temporal characterization of dicarboxylic acids in PM2. 5 during a PM episode and a period of non-episodic pollution. Atmos. Environ. 2008, 42, 6836– 6850, DOI: 10.1016/j.atmosenv.2008.05.021There is no corresponding record for this reference.
- 28Radek, M.; Savage, G. P. Oxalates in Some Indian Green Leafy Vegetables. Int. J. Food Sci. Nutr. 2008, 59, 246– 260, DOI: 10.1080/0963748070179117628https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXjtVeqsbc%253D&md5=a5f887d5c343c62daa6a4166bc1ee17cOxalates in some Indian green leafy vegetablesRadek, M.; Savage, G. P.International Journal of Food Sciences and Nutrition (2008), 59 (3), 246-260CODEN: IJFNEH; ISSN:0963-7486. (Informa Healthcare)The sol. and total oxalate contents of 11 leafy vegetables grown in India were detd. Spinach, purple and green amaranth and colocasia contained high levels of total oxalates, which ranged from 5,138.0±37.6 mg/100 g dry matter up to 12,576.1±107.9 mg/100 g dry matter. Seven other leafy vegetables (curry, drumstick, shepu, fenugreek, coriander, radish and onion stalks) contained only insol. oxalate, which ranged from 209.0±5.0 mg/100 g dry matter to 2,774.9±18.4 mg/100 g dry matter. In vitro digestion of the samples showed that the gastric available oxalate was 10% lower than the values obtained from acid extn. and that intestinal available oxalate was 20% lower than the values obtained following hot water extn. The percentage calcium bound in the insol. oxalate fraction of the dried leafy vegetables ranged from 3.3% to 86.7% of the total calcium. Addn. of four different sources of calcium (low fat milk, whole milk, calcium carbonate and calcium sulfate) resulted in a range of 32-100% redns. of intestinal available oxalate in the mixt.
- 29Holmes, R. P.; Kennedy, M. Estimation of the Oxalate Content of Foods and Daily Oxalate Intake. Kidney Int. 2000, 57, 1662– 1667, DOI: 10.1046/j.1523-1755.2000.00010.x29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXjtVOrtbo%253D&md5=00c6a7ce30d73d679e1d11e1ebd985e3Estimation of the oxalate content of foods and daily oxalate intakeHolmes, Ross P.; Kennedy, MarthaKidney International (2000), 57 (4), 1662-1667CODEN: KDYIA5; ISSN:0085-2538. (Blackwell Science, Inc.)Background. The amt. of oxalate ingested may be an important risk factor in the development of idiopathic calcium oxalate nephrolithiasis. Reliable food tables listing the oxalate content of foods are currently not available. The aim of this research was to develop an accurate and reliable method to measure the food content of oxalate. Methods. Capillary electrophoresis (CE) and ion chromatog. (IC) were compared as direct techniques for the estn. of the oxalate content of foods. Foods were thoroughly homogenized in acid, heat extd., and clarified by centrifugation and filtration before diln. in water for anal. Five individuals consuming self-selected diets maintained food records for three days to det. their mean daily oxalate intakes. Results. Both techniques were capable of adequately measuring the oxalate in foods with a significant oxalate content. With foods of very low oxalate content (<1.8 mg/100 g), IC was more reliable than CE. The mean daily intake of oxalate by the five individuals tested was 152 ± 83 mg, ranging from 44 to 352 mg/day. Conclusions. CE appears to be the method of choice over IC for estg. the oxalate content of foods with a medium (>10 mg/100 g) to high oxalate content due to a faster anal. time and lower running costs, whereas IC may be better suited for the anal. of foods with a low oxalate content. Accurate ests. of the oxalate content of foods should permit the role of dietary oxalate in urinary oxalate excretion and stone formation to be clarified. Other factors, apart from the amt. of oxalate ingested, appear to exert a major influence over the amt. of oxalate excreted in the urine.
- 30Salgado, N.; Silva, M. A.; Figueira, M. E.; Costa, H. S.; Albuquerque, T. G. Oxalate in Foods: Extraction Conditions, Analytical Methods, Occurrence, and Health Implications. Foods 2023, 12, 3201, DOI: 10.3390/foods12173201There is no corresponding record for this reference.
- 31Siener, R.; Seidler, A.; Hönow, R. Oxalate-rich foods. Food Sci. Technol. 2021, 41, 169– 173, DOI: 10.1590/fst.10620There is no corresponding record for this reference.
- 32Mochizuki, T.; Kawamura, K.; Miyazaki, Y.; Wada, R.; Takahashi, Y.; Saigusa, N.; Tani, A. Secondary formation of oxalic acid and related organic species from biogenic sources in a larch forest at the northern slope of Mt. Fuji. Atmos. Environ. 2017, 166, 255– 262, DOI: 10.1016/j.atmosenv.2017.07.02832https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXht1amsr3N&md5=e3c89cddfcc6e85b8622a0bcbc4f928bSecondary formation of oxalic acid and related organic species from biogenic sources in a larch forest at the northern slope of Mt. FujiMochizuki, Tomoki; Kawamura, Kimitaka; Miyazaki, Yuzo; Wada, Ryuichi; Takahashi, Yoshiyuki; Saigusa, Nobuko; Tani, AkiraAtmospheric Environment (2017), 166 (), 255-262CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)To better understand the formation of water-sol. org. aerosols in the forest atm., we measured low mol. wt. (LMW) dicarboxylic acids, oxocarboxylic acids, α-dicarbonyls, unsatd. fatty acids (UFAs), and water-sol. org. carbon (WSOC) in aerosols from a Larix kaempferi forest located at the northern slope of Mt. Fuji, Japan, in summer 2012. Concns. of dicarboxylic acids, oxocarboxylic acids, α-dicarbonyls, and WSOC showed maxima in daytime. Relative abundance of oxalic acid in LMW dicarboxylic acids was on av. 52% and its av. concn. was 214 ng m-3. We found that diurnal and temporal variations of oxalic acid are different from those of isoprene and α-pinene, whereas biogenic secondary org. aerosols (BSOAs) derived from isoprene and α-pinene showed similar variations with oxalic acid. The mass concn. ratios of oxalic acid/BSOAs were relatively const., although a large variation in the concns. of toluene that is an anthropogenic volatile org. compd. was obsd. These results suggest that formation of oxalic acid is assocd. with the oxidn. of isoprene and α-pinene with O3 and other oxidants in the forest atm. In addn., concns. of UFAs were obsd., for the first time, to decrease dramatically during daytime in the forest. This study demonstrates that forest ecosystem is an important source of oxalic acid and other dicarboxylic acids in the atm.
- 33Kawamura, K.; Bikkina, S. A review of dicarboxylic acids and related compounds in atmospheric aerosols: Molecular distributions, sources and transformation. Atmos. Res. 2016, 170, 140– 160, DOI: 10.1016/j.atmosres.2015.11.01833https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXitVSqt7vF&md5=fcba54277b4abd2ddc9092cefe8141f6A review of dicarboxylic acids and related compounds in atmospheric aerosols: Molecular distributions, sources and transformationKawamura, Kimitaka; Bikkina, SrinivasAtmospheric Research (2016), 170 (), 140-160CODEN: ATREEW; ISSN:0169-8095. (Elsevier B.V.)This review aims to update our understanding on mol. distributions of water-sol. dicarboxylic acids and related compds. in atm. aerosols with a focus on their geog. variability, size distribution, sources and formation pathways. In general, mol. distributions of diacids in aerosols from the continental sites and over the open ocean waters are often characterized by the predominance of oxalic acid (C2) followed by malonic acid (C3) and/or succinic acid (C4), while those sampled over the polar regions often follow the order of C4 ≥ C2 and C3. The most abundant and ubiquitous diacid is oxalic acid, which is principally formed via atm. oxidn. of its higher homologues of long chain diacids and other pollution-derived org. precursors (e.g., olefins and arom. hydrocarbons). However, its occurrence in marine aerosols is mainly due to the transport from continental outflows (e.g., East Asian outflow during winter/spring to the North Pacific) and/or governed by photochem./aq. phase oxidn. of biogenic unsatd. fatty acids (e.g., oleic acid) and isoprene emitted from the productive open ocean waters. The long-range atm. transport of pollutants from mid latitudes to the Arctic in dark winter facilitates to accumulate the reactants prior to their intense photochem. oxidn. during springtime polar sunrise. Furthermore, the relative abundances of C2 in total diacid mass showed similar temporal trends with downward solar irradn. and ambient temps., suggesting the significance of atm. photochem. oxidn. processing. Compd.-specific isotopic analyses of oxalic acid showed the highest δ13C among diacids whereas azelaic acid showed the lowest value, corroborating the significance of atm. aging of oxalic acid. On the other hand, other diacids gave intermediate values between these two diacids, suggesting that aging of oxalic acid is assocd. with 13C enrichment.
- 34Claeys, M. Formation of secondary organic aerosols through photooxidation of isoprene. Science 2004, 303, 1173– 1176, DOI: 10.1126/science.109280534https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXhsVWgtb4%253D&md5=43cb23180fe04a94a189bf899aaa121fFormation of Secondary Organic Aerosols Through Photooxidation of IsopreneClaeys, Magda; Graham, Bim; Vas, Gyorgy; Wang, Wu; Vermeylen, Reinhilde; Pashynska, Vlada; Cafmeyer, Jan; Guyon, Pascal; Andreae, Meinrat O.; Artaxo, Paulo; Maenhaut, WillyScience (Washington, DC, United States) (2004), 303 (5661), 1173-1176CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Detailed org. anal. of natural aerosols from the Amazonian rain forest showed considerable quantities of previously unobserved polar org. compds. identified as a mixt. of 2 diastereoisomeric 2-methyltetrols: 2-methylthreitol and 2-methylerythritol. These polyols, which have an isoprene skeleton, can be explained by OH--initiated photooxidn. of isoprene. They have low vapor pressure, allowing them to condense onto pre-existing particles. It was estd. that photooxidn. of isoprene results in an annual global prodn. of ∼2 Tg of the polyols, a substantial fraction of the Intergovernmental Panel on Climate Change est. of 8-40 Tg/yr of secondary org. aerosol from biogenic sources.
- 35Kawamura, K.; Ikushima, K. Seasonal changes in the distribution of dicarboxylic acids in the urban atmosphere. Environ. Sci. Technol. 1993, 27, 2227– 2235, DOI: 10.1021/es00047a03335https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXlsVGnt78%253D&md5=7c6f649ee0b03f2817ae1bf072414d93Seasonal changes in the distribution of dicarboxylic acids in the urban atmosphereKawamura, Kimitaka; Ikushima, KouichiEnvironmental Science and Technology (1993), 27 (10), 2227-35CODEN: ESTHAG; ISSN:0013-936X.Low-mol.-wt. dicarboxylic acids were studied in urban aerosol samples collected from Tokyo atm. in 1988-1989, by a capillary GC and GC-mass spectrometry method using a di-Bu ester derivatization technique. A homologous series of C2-C10 dicarboxylic acids were detected in the aerosols, including straight-chain satd., unsatd., branched-chain, and hydroxylated species. Mol. distributions of the diacids demonstrated that oxalic (C2) acid was the most abundant throughout the 4 seasons, followed by C3 and C4 diacids; the C2 diacid comprised 37-69% of the total diacid concns. Diacids with higher C nos. were less abundant, although C6 and C9 diacids were relatively abundant. Total concns. of diacids were 90-1370 ng/m3, which corresponded to 0.06-1.1% of the total aerosol mass. Their concns. relative to aerosol mass showed a seasonal trend with a max. in Aug. The C2-C4 diacid C normalized by aerosol total C also showed a strong seasonal change with a max. in the summer and indicated a pos. correlation with oxidant concns. Interestingly, lower diacids showed higher values for correlation coeff. (e.g., r = 0.80 for oxalic acid), suggesting a preferential prodn. of low mol. wt. dicarboxylic acids, esp., C2 diacid. Possible reaction mechanisms are discussed in relation to the preferential accumulation of lower diacids in the atm.
- 36Prenni, A. J.; DeMott, P. J.; Kreidenweis, S. M.; Sherman, D. E.; Russell, L. M.; Ming, Y. J. Phys. Chem. A 2001, 105, 11240– 11248, DOI: 10.1021/jp012427d36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXosVegtrc%253D&md5=924edef0b06242bbb43564ebe9065835The effects of low molecular weight dicarboxylic acids on cloud formationPrenni, Anthony J.; DeMott, Paul J.; Kreidenweis, Sonia M.; Sherman, D. Eli; Russell, Lynn M.; Ming, YiJournal of Physical Chemistry A (2001), 105 (50), 11240-11248CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The ubiquitous presence of org. compds. in tropospheric particles requires that their role in aerosol/cloud interactions be accounted for in climate models. In this paper, we present studies that investigate the hygroscopic behavior of org. compds. and their efficiency as ice nuclei. Specifically, results for sol. and partially sol. dicarboxylic acids that have been obsd. in atm. aerosol are discussed. At room temp., we use a humidified tandem differential mobility analyzer (HTDMA) and a condensation particle counter interfaced with a cloud condensation nuclei counter to characterize the water uptake behavior of these acids. The HTDMA data agree quite well with modeled hygroscopic behavior. However, we find that some of the compds. retain water to very low humidities, never exhibiting efflorescence. The studies are extended to lower temps. using a continuous flow ice thermal diffusion chamber to investigate the role of these species in ice nucleation at cirrus conditions. Results suggest that ice formation occurs via homogeneous nucleation for most of these acids, and that nucleation for these acids is not as efficient as that for sulfate aerosol.
- 37Makar, P. A. The estimation of organic gas vapour pressure. Atmos. Environ. 2001, 35, 961– 974, DOI: 10.1016/S1352-2310(00)00343-5There is no corresponding record for this reference.
- 38Limbeck, A.; Puxbaum, H.; Otter, L.; Scholes, M. C. Atmos. Environ. 2001, 35, 1853– 1862, DOI: 10.1016/S1352-2310(00)00497-038https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXht1yqu7Y%253D&md5=0a1286ec3bec76bafa2486c4e090a43bSemivolatile behavior of dicarboxylic acids and other polar organic species at a rural background site (Nylsvley, RSA)Limbeck, A.; Puxbaum, H.; Otter, L.; Scholes, M. C.Atmospheric Environment (2001), 35 (10), 1853-1862CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Science Ltd.)In this study aerosol samples from the South African savanna were analyzed for their polar org. constituents. Samples were collected with a front/back-up filter tandem system of quartz fiber filters (dual filter strategy). In all samples (n=15) dicarboxylic acids and a variety of phthalates, aldehydes and monocarboxylic acids were obsd. Oxalic acid was the dominating compd. with an av. amt. of 79.2 ng m-3 on the front filter and 11.3 ng m-3 on the back-up filter. The presence of significant concns. of dicarboxylic acids on the back-up filter was rather unexpected. There are two possible sources to explain the presence of individual compds. on the back-up filter - particle penetration through the front filter or adsorption of compd. parts from the gas phase. Interpretation of the data indicates that the dicarboxylic acid concns. on the back-up filters appear to be caused by the adsorption of gaseous org. species. Dicarboxylic acids semivolatile behavior is evident with this results. This conclusion refutes the commonly held view that dicarboxylic acids in the atm. were assocd. with the aerosol phase only. Addnl., it was found that the distribution of dicarboxylic acids between the gas and particle phase in the atm. is not only dependent on their vapor pressures. The actual gas phase concn. appears to be more detd. by the chem. properties of the particles than by pure phys. influences. Surprisingly, malonic acid exhibits an anomaly, as it does not show a semivolatile tendency.
- 39Cheng, C.; Li, M.; Chan, C. K.; Tong, H.; Chen, C.; Chen, D.; Wu, D.; Li, L.; Wu, C.; Cheng, P. Mixing state of oxalic acid containing particles in the rural area of Pearl River Delta, China: implications for the formation mechanism of oxalic acid. Atmos. Chem. Phys. 2017, 17 (15), 9519– 9533, DOI: 10.5194/acp-17-9519-201739https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhs1CmtbjO&md5=b18ab18e9b09d1d72f2c7fa31c0ddfbbMixing state of oxalic acid containing particles in the rural area of pearl river delta, china: implications for the formation mechanism of oxalic acidCheng, Chunlei; Li, Mei; Chan, Chak K.; Tong, Haijie; Chen, Changhong; Chen, Duohong; Wu, Dui; Li, Lei; Wu, Cheng; Cheng, Peng; Gao, Wei; Huang, Zhengxu; Li, Xue; Zhang, Zhijuan; Fu, Zhong; Bi, Yanru; Zhou, ZhenAtmospheric Chemistry and Physics (2017), 17 (15), 9519-9533CODEN: ACPTCE; ISSN:1680-7324. (Copernicus Publications)The formation of oxalic acid and its mixing state in atm. particulate matter (PM) were studied using a single-particle aerosol mass spectrometer (SPAMS) in the summer and winter of 2014 in Heshan, a supersite in the rural area of the Pearl River Delta (PRD) region in China. Oxalic-acid-contg. particles accounted for 2.5 and 2.7 % in total detected ambient particles in summer and winter, resp. Oxalic acid was measured in particles classified as elemental carbon (EC), org. carbon (OC), elemental and org. carbon (ECOC), biomass burning (BB), heavy metal (HM), secondary (Sec), sodium-potassium (NaK), and dust. Oxalic acid was found predominantly mixing with sulfate and nitrate during the whole sampling period, likely due to aq.-phase reactions. In summer, oxalic-acid-contg. particle no. and ozone concn. followed a very similar trend, which may reflect the significant contribution of photochem. reactions to oxalic acid formation. The HM particles were the most abundant oxalic acid particles in summer and the diurnal variations in peak area of iron and oxalic acid show opposite trends, which suggests a possible loss of oxalic acid through the photolysis of iron oxalato-complexes during the strong photochem. activity period. In wintertime, carbonaceous particles contained a substantial amt. of oxalic acid as well as abundant carbon clusters and BB markers. The general existence of nitric acid in oxalic-acid-contg. particles indicates an acidic environment during the formation process of oxalic acid. The peak areas of nitrate, sulfate and oxalic had similar temporal change in the carbonaceous type oxalic acid particles, and the organosulfate-contg. oxalic acid particles correlated well with total oxalic acid particles during the haze episode, which suggests that the formation of oxalic acid is closely assocd. with the oxidn. of org. precursors in the aq. phase.
- 40Yu, S. Role of organic acids formic, acetic, pyruvic and/ oxalic in the formation of cloud condensation nuclei CCN: A Review. Atmos. Res. 2000, 53, 185– 217, DOI: 10.1016/S0169-8095(00)00037-540https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXitlKru7s%253D&md5=a057799ada7073f81e6d73aff1f9549dRole of organic acids (formic, acetic, pyruvic and oxalic) in the formation of cloud condensation nuclei (CCN): a reviewYu, S.Atmospheric Research (2000), 53 (4), 185-217CODEN: ATREEW; ISSN:0169-8095. (Elsevier Science B.V.)A review, with many refs., is given. Although it is believed that org. aerosols play a key role in cloud nucleation and make an important contribution to the cloud condensation nuclei (CCN) population, their specific species remain poorly characterized. This paper reviews the current knowledge of org. acids (mainly formic, acetic, pyruvic and oxalic acids). Without specification, org. acids in this paper refer to these four org. acids in the gas and aerosol phases. This paper analyzes the extent to which org. acids act as CCN and compares the phys. and chem. properties of org. acids with those of CCN. The results show that aerosol formate and acetate concns. range from 0.02 to 5.3 nmol m-3 and from 0.03 to 12.4 nmol m-3, resp., and that between 34 to 77% of formate and between 21 to 66% of acetate are present in the fine fraction of aerosols. It was found that although most (98-99%) of these volatile org. acids are present in the gas phase, their concns. in the aerosol particles are sufficient to make them a good candidate for CCN. The results also show that org. acids may make an important contribution to the formation of CCN in some special sources such as vegetation emissions and biomass-burning. Org. acids are expected to contribute significantly to the ests. of indirect (cloud-mediated) forcing due to aerosols.
- 41Peng, C.; Chan, M. N.; Chan, C. K. The Hygroscopic Properties of Dicarboxylic and Multifunctional Acids: Measurements and UNIFAC Predictions. Environ. Sci. Technol. 2001, 35, 4495– 4501, DOI: 10.1021/es010753141https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXnsFynsbY%253D&md5=fdf0eb4a3d286bd256b1e60985034db3The Hygroscopic Properties of Dicarboxylic and Multifunctional Acids: Measurements and UNIFAC PredictionsPeng, Changgeng; Chan, Man Nin; Chan, Chak K.Environmental Science and Technology (2001), 35 (22), 4495-4501CODEN: ESTHAG; ISSN:0013-936X. (American Chemical Society)The role of water-sol. org. compds. in regard to the hygroscopic properties of atm. aerosols has recently been the subject of many studies. In particular, low-mol.-wt. dicarboxylic acids and some multifunctional org. acids have been found or are expected to exist in atm. aerosols at urban, semiurban, rural, and remote sites. Unlike the case of their inorg. counterparts, the hygroscopic properties of org. acids have not been well characterized. In this study, the hygroscopic properties of selected water-sol. dicarboxylic acids (oxalic acid, malonic acid, succinic acid, and glutaric acid) and multifunctional acids (citric acid, DL-malic acid, and L-(+)-tartaric acid) were studied using single droplets levitated in an electrodynamic balance at 25°. The water activities of bulk samples of dil. solns. were also measured. Solute evapn. was obsd. for the dicarboxylic acids but not for the multifunctional acids. Oxalic acid, succinic acid, and glutaric acid droplets crystallize upon evapn. of water, but, except for glutaric acid droplets, do not deliquesce even at 90% relative humidity (RH). Mass transfer limitation of the deliquescence process was obsd. in glutaric acid. Neither crystn. nor deliquescence was obsd. in malonic acid, citric acid, DL-malic acid, or L-(+)-tartaric acid. Malonic acid and these three hydroxy carboxylic acids absorb water even at RHs much lower than their resp. deliquescence RH. The growth factor (Gf), defined as the ratio of the particle diam. at RH = 10% to that at RH = 90%, of oxalic acid and succinic acid was close to unity, indicating no hygroscopicity in this range. The remaining acids (malonic acid, glutaric acid, citric acid, malic acid, and tartaric acid) showed roughly similar hygroscopicity of a Gf of 1.30-1.53, which is similar to that of "more hygroscopic" aerosols in field measurements reported in the literature. A generalized equation for these four acids, Gf = (1-aw)-0.163, was developed to represent the hygroscopicity of these acids. Water activity predictions from calcns. using the UNIFAC model were found to agree with the measured water activity data to within 40% for most of the acids, but the deviations were as large as about 100% for malic acid and tartaric acid. We modified the functional group interaction parameters of the COOH-H2O, OH-H2O, and OH-COOH pairs by fitting the UNIFAC model with the measured data. The modified UNIFAC model improves the agreement of predictions and measurements to within 38% for all the acids studied.
- 42Kerminen, V.-M.; Ojanen, C.; Pakkanen, T.; Hillamo, R.; Aurela, M.; Meriläinen, J. Low-molecular-weight dicarboxylic acids in an urban and rural atmosphere. J. Aerosol Sci. 2000, 31, 349– 362, DOI: 10.1016/S0021-8502(99)00063-442https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXhslKgurY%253D&md5=2787e1a0647a9c1617370defb34771a7Low-molecular-weight dicarboxylic acids in an urban and rural atmosphereKerminen, Veli-Matti; Ojanen, Christina; Pakkanen, Tuomo; Hillamo, Risto; Aurela, Minna; Merilainen, JouniJournal of Aerosol Science (2000), 31 (3), 349-362CODEN: JALSB7; ISSN:0021-8502. (Elsevier Science Ltd.)Chem. of oxalic, malonic, and succinic acid was studied at the two sites representing the urban and rural conditions, and at a site intermediate between these two. The investigation was based on the particle collection with a virtual impactor and a Berner low-pressure impactor. Concns. of the three diacids displayed large seasonal amplitudes with low values in winter. Suggestive of common sources or atm. formation processes, the correlation between oxalic and malonic acid concns. was high. Both the local traffic and secondary prodn. in the long-range transported air masses seemed to be the important sources for these two acids. Contrary to oxalic and malonic acid, no enrichment at the urban site compared with the rural site was obsd. for succinic acid. The seasonal cycle of this acid resembled that of methanesulfonic acid. The most likely sources for succinic acid in our samples was the secondary prodn. in the long-range transported air, with potentially significant contribution coming from biogenic sources. The three diacids had quite different distributions over the particulate phase. Oxalic acid had a dominant accumulation mode, a clear Aitken mode at sizes below about 0.15 μm of particle diam., and modes corresponding to the sea-salt and crustal particle size ranges. Most of the malonic acid was assocd. with sea-salt particles, even though in a few samples an accumulation mode was also present. Succinic acid was distributed between the accumulation and the sea-salt particle modes, in addn. to which it frequently had quite a pronounced Aitken mode. Oxalic and succinic acids are among the orgs. that may contribute to the atm. cloud condensation nuclei prodn. Oxalic and malonic acid, and to a smaller extent succinic acid, participate in reactions occurring in sea-salt particles.
- 43Khwaja, H. A. Atmospheric concentrations of carboxylic acids and related compounds at a semiurban site. Atmos. Environ. 1995, 29, 127– 139, DOI: 10.1016/1352-2310(94)00211-343https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXjtVOltbk%253D&md5=4b4b585b07cfe89ea6931ba997442f9fAtmospheric concentrations of carboxylic acids and related compounds at a semiurban siteKhwaja, Haider A.Atmospheric Environment (1995), 29 (1), 127-39CODEN: AENVEQ; ISSN:1352-2310. (Elsevier)Atm.-gas- and particle-phase carboxylic acids and related compds. were measured during Oct. 1991 in a semiurban site in northeastern US. Formic and acetic acids were present in the atm. mostly in the gaseous form with <10% in the particle phase. Concns. of formic acid and acetic acid were in the 0.80-2.5 and 0.60-3.4 ppbv range, resp. Formic acid was correlated with acetic acid (r =0.94). Diurnal variations of formic acid were similar to those of O3. This pattern is consistent with photochem. prodn. as an important source. Formic-to-acetic acid ratios <1 were recorded, likely due to an increase in acetic acid contribution from direct emissions assocd. with vehicular traffic. HCHO levels began to decrease after the mid-afternoon maxima; concns. were 0.63-3.7 ppbv. Formic, acetic, pyruvic, glyoxalic, oxalic, succinic, and malonic acids in the particle phase were identified. Aerosol carboxylic acid concns. were 26-360 ng/m3. About 80% of the carboxylic acid aerosol mass was in the size fraction <1.0 μm diam. Of the dicarboxylic acids, oxalic acid was the most abundant species, followed by succinic acid and malonic acid. Particulate total org. C exhibited a concn. range 12.6 × 103-49.9 × 103 ng/cm3. The obsd. amts. of carboxylic acids in the particle phase accounted for a small fraction of the org. C. Results indicated that photochem. processes and anthropogenic emissions such as automobile exhaust are major sources of atm. carboxylic acids.
- 44Chebbi, A.; Carlier, P. Atmos. Environ. 1996, 30, 4233– 4249, DOI: 10.1016/1352-2310(96)00102-144https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XmtFaqsLo%253D&md5=c0a6e50637ff4490082745a52af84c1fCarboxylic acids in the troposphere, occurrence, sources, and sinks: a reviewChebbi, A.; Carlier, P.Atmospheric Environment (1996), 30 (24), 4233-4249CODEN: AENVEQ; ISSN:1352-2310. (Elsevier)A review, with many refs., is given. Carboxylic acids are ubiquitous and important components of the troposphere; they are currently measured in different environments. They are thought to have several sources comprising primary biogenic and anthropogenic emissions, hydrocarbons gas-phase oxidns., and some carbonyl compds. aq.-phase oxidns. We make a synthesis of the concns. of low mol. wt. carboxylic acids in tropospheric aq. and gaseous phases and in aerosol particles for different environments. We also successively exam. the major sources of carboxylic acids and discuss their relative contribution to tropospheric concns. for various environments as well as the principal sinks of these compds.
- 45Martinelango, P. K.; Dasgupta, P. K.; Al-Horr, R. S. Atmos. Environ. 2007, 41, 4258– 4269, DOI: 10.1016/j.atmosenv.2006.05.08545https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXlsFaktrc%253D&md5=8cfe5cbfecf8eb9c71292811467424dbAtmospheric production of oxalic acid/oxalate and nitric acid/nitrate in the Tampa Bay airshed: Parallel pathwaysMartinelango, P. Kalyani; Dasgupta, Purnendu K.; Al-Horr, Rida S.Atmospheric Environment (2007), 41 (20), 4258-4269CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)Oxalic acid is the dominant dicarboxylic acid (DCA), and it constitutes up to 50% of total atm. DCAs, esp. in non-urban and marine atmospheres. A significant amt. of particulate H2Ox/oxalate (Ox) occurred in the coarse particle fraction of a dichotomous sampler, the ratio of oxalate concns. in the PM10 to PM2.5 fractions ranged from 1 to 2, with mean±sd being 1.4 ± 0.2. These results suggest that oxalate does not solely originate in the gas phase and condense into particles. Gaseous H2Ox concns. are much lower than particulate Ox concns. and are well correlated with HNO3, HCHO, and O3, supporting a photochem. origin. Of special relevance to the Bay Region Atm. Chem. Expt. (BRACE) is the extent of nitrogen deposition in the Tampa Bay estuary. Hydroxyl radical is primarily responsible for the conversion of NO2 to HNO3, the latter being much more easily deposited. Hydroxyl radical is also responsible for the aq. phase formation of oxalic acid from alkenes. Hence, we propose that an est. of ·OH can be obtained from H2Ox/Ox prodn. rate and we accordingly show that the product of total oxalate concn. and NO2 concn. approx. predicts the total nitrate concn. during the same period.
- 46Ma, X.; Sun, Y.; Huang, Z.; Zhang, Q.; Wang, W. A density functional theory study of the molecular interactions between a series of amides and sulfuric acid. Chemosphere 2019, 214, 781– 790, DOI: 10.1016/j.chemosphere.2018.08.15246https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhvVOjsLzF&md5=5d4d3a26a63f7f4da197456377140e62A density functional theory study of the molecular interactions between a series of amides and sulfuric acidMa, Xiaohui; Sun, Yanhui; Huang, Zixiao; Zhang, Qingzhu; Wang, WenxingChemosphere (2019), 214 (), 781-790CODEN: CMSHAF; ISSN:0045-6535. (Elsevier Ltd.)Amides, a class of nitrogen-contg. org. pollutants in the atm., may affect the formation of atm. aerosols by the interactions with sulfuric acid. Here, the mol. interactions of sulfuric acid with formamide, methylformamide, DMF, acetamide, methylacetamide and dimethylacetamide was investigated by d. functional theory. Geometry optimization and Gibbs free energy calcn. were carried out at M06-2X/6-311++G(3df,3pd) level. The results indicate that the addn. of amides to H2SO4 might have a promoting effect on atm. new particle formation at 298.15 K and 1 atm. In the initial stage of new particle formation, the binding capacity of amides and sulfuric acid is stronger than ammonia, but weaker than methylamine. It is worth noting that the trans-methylacetamide could have similar capabilities of stabilizing sulfuric acid as dimethylamine. In the presence of water, amides are found to only have a weak enhancement capability on new particle formation. In addn., we can infer from evapn. rate that the small mol. clusters of formamide and sulfuric acid may be more energetically favorable than macromol. clusters.
- 47Hanson, D. R.; Eisele, F. L. Measurement of prenucleation molecular clusters in the NH3, H2SO4, H2O system. J. Geophys. Res.: Atmos. 2002, 107, AAC 10– 1-AAC 10-18, DOI: 10.1029/2001JD001100There is no corresponding record for this reference.
- 48Sipilää, M.; Berndt, T.; Petäjä, T.; Brus, D.; Vanhanen, J.; Stratmann, F.; Patokoski, J.; Mauldin, R. L., 3; Hyvärinen, A.-P.; Lihavainen, H. The role of sulfuric acid in atmospheric nucleation. Science 2010, 327, 1243– 1246, DOI: 10.1126/science.1180315There is no corresponding record for this reference.
- 49Kirkby, J.; Curtius, J.; Almeida, J.; Dunne, E.; Duplissy, J.; Ehrhart, S. Role of sulphuric acid, ammonia and galactic cosmic rays in atmospheric aerosol nucleation. Nature 2011, 476, 429– 433, DOI: 10.1038/nature1034349https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtVOmsr7F&md5=ab78f81c052b4b7a47e9c02f6b670b94Role of sulphuric acid, ammonia and galactic cosmic rays in atmospheric aerosol nucleationKirkby, Jasper; Curtius, Joachim; Almeida, Joao; Dunne, Eimear; Duplissy, Jonathan; Ehrhart, Sebastian; Franchin, Alessandro; Gagne, Stephanie; Ickes, Luisa; Kuerten, Andreas; Kupc, Agnieszka; Metzger, Axel; Riccobono, Francesco; Rondo, Linda; Schobesberger, Siegfried; Tsagkogeorgas, Georgios; Wimmer, Daniela; Amorim, Antonio; Bianchi, Federico; Breitenlechner, Martin; David, Andre; Dommen, Josef; Downard, Andrew; Ehn, Mikael; Flagan, Richard C.; Haider, Stefan; Hansel, Armin; Hauser, Daniel; Jud, Werner; Junninen, Heikki; Kreissl, Fabian; Kvashin, Alexander; Laaksonen, Ari; Lehtipalo, Katrianne; Lima, Jorge; Lovejoy, Edward R.; Makhmutov, Vladimir; Mathot, Serge; Mikkilae, Jyri; Minginette, Pierre; Mogo, Sandra; Nieminen, Tuomo; Onnela, Antti; Pereira, Paulo; Petaejae, Tuukka; Schnitzhofer, Ralf; Seinfeld, John H.; Sipilae, Mikko; Stozhkov, Yuri; Stratmann, Frank; Tome, Antonio; Vanhanen, Joonas; Viisanen, Yrjo; Vrtala, Aron; Wagner, Paul E.; Walther, Hansueli; Weingartner, Ernest; Wex, Heike; Winkler, Paul M.; Carslaw, Kenneth S.; Worsnop, Douglas R.; Baltensperger, Urs; Kulmala, MarkkuNature (London, United Kingdom) (2011), 476 (7361), 429-433CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Model calcns. suggest that almost half of the global cloud condensation nuclei in the atm. boundary layer may originate from the nucleation of aerosols from trace condensable vapors, although the sensitivity of the no. of cloud condensation nuclei to changes of nucleation rate may be small. Despite extensive research, fundamental questions remain about the nucleation rate of sulfuric acid particles and the mechanisms responsible, including the roles of galactic cosmic rays and other chem. species such as ammonia. First results from the Cosmics Leaving OUtdoor Droplets (CLOUD) expt. at CERN are presented. It was found that atmospherically relevant ammonia mixing ratios of 100 parts per trillion by vol., or less, increase the nucleation rate of sulfuric acid particles more than 100-1,000-fold. Time-resolved mol. measurements reveal that nucleation proceeds by a base-stabilization mechanism involving the stepwise accretion of ammonia mols. Ions increase the nucleation rate by an addnl. factor of between two and more than ten at ground-level galactic-cosmic-ray intensities, provided that the nucleation rate lies below the limiting ion-pair prodn. rate. Ion-induced binary nucleation of H2SO4-H2O can occur in the mid-troposphere but is negligible in the boundary layer. However, even with the large enhancements in rate due to ammonia and ions, atm. concns. of ammonia and sulfuric acid are insufficient to account for obsd. boundary-layer nucleation.
- 50Elm, J.; Kubečka, J.; Besel, V.; Jääskeläinen, M. J.; Halonen, R.; Kurtén, T.; Vehkamäki, H. Modeling the formation and growth of atmospheric molecular clusters: A review. J. Aerosol Sci. 2020, 149, 105621, DOI: 10.1016/j.jaerosci.2020.10562150https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhtlyisLfI&md5=5c62c26f1491d26b091d31fa8d341801Modeling the formation and growth of atmospheric molecular clusters: A reviewElm, Jonas; Kubecka, Jakub; Besel, Vitus; Jaaskelainen, Matias J.; Halonen, Roope; Kurten, Theo; Vehkamaki, HannaJournal of Aerosol Science (2020), 149 (), 105621CODEN: JALSB7; ISSN:0021-8502. (Elsevier Ltd.)A review,. Mol. clusters are ubiquitous constituents of the ambient atm., that can grow into larger sizes forming new aerosol particles. The formation and growth of small clusters into aerosol particles remain one of the largest uncertainties in global climate predictions. This has made the modeling of atm. mol. clustering into an active field of research, yielding direct mol. level information about the formation mechanism. We review the present state-of-the-art quantum chem. methods and cluster distribution dynamics models that are applied to study the formation and growth of atm. mol. clusters. We outline the current challenges in applying theor. methods and the future directions to move the field forward.
- 51Zhang, R.; Khalizov, A.; Wang, L.; Hu, M.; Xu, W. Nucleation and Growth of Nanoparticles in the Atmosphere. Chem. Rev. 2012, 112, 1957– 2011, DOI: 10.1021/cr200175651https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtl2itrnN&md5=a5b16c46ecf502c603291ce50de25604Nucleation and Growth of Nanoparticles in the AtmosphereZhang, Renyi; Khalizov, Alexei; Wang, Lin; Hu, Min; Xu, WenChemical Reviews (Washington, DC, United States) (2012), 112 (3), 1957-2011CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review which critically assesses recent findings related to nucleation and growth of atm. nanoparticles, emphasizing process understanding on a fundamental mol. level, is given. Topics discussed include: vapor nucleation overview (nucleation theories and computational approaches [classical nucleation, kinetic, and d. functional theories, mol. dynamics and Monte Carlo methods, nucleation theorem], nucleation expts. [adiabatic expansion approaches; diffusion, laminar flow, and turbulent mixing chambers; continuous generation of nucleating vapors from chem. reaction sources; exptl. results vs. nucleation theories]); atm. nanoparticle nucleation (measurements [concns. and size distribution, nanoparticle chem. compn., charged and neutral atm. clusters], lab. studies [binary nucleation of H2SO4/water; NH3 and amine-involved ternary nucleation of H2SO4/water; org. acid assisted nucleation of H2SO4/water; IOx and ion-induced nucleation; nucleating cluster chem. compn., reactivity, and thermodn.], theor./computational studies [quantum chem. calcns., mol. dynamics and Monte Carlo simulations], atm. nucleation parameterization); nanoparticle atm. growth (Kelvin [curvature] effect, condensation [H2SO4, low volatility orgs.], heterogeneous reactions [NH3, amines, aldehydes, α-dicarbonyls, alcs., other species]); numerical treatment of ambient nanoparticle nucleation and growth rates (measured, combined growth including condensation and intra-/extra-modal coagulation, nucleation rate derivation from atm. measurements); and summary and future research needs.
- 52Zheng, J.; Ma, Y.; Chen, M.; Zhang, Q.; Wang, L.; Khalizov, A. F. Measurement of atmospheric amines and ammonia using the high resolution time-of-flight chemical ionization mass spectrometry. Atmos. Environ. 2015, 102, 249– 259, DOI: 10.1016/j.atmosenv.2014.12.00252https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXitVWgtLbE&md5=c4fae6404f558c380000c0c9b1939556Measurement of atmospheric amines and ammonia using the high resolution time-of-flight chemical ionization mass spectrometryZheng, Jun; Ma, Yan; Chen, Mindong; Zhang, Qi; Wang, Lin; Khalizov, Alexei F.; Yao, Lei; Wang, Zhen; Wang, Xing; Chen, LinxiAtmospheric Environment (2015), 102 (), 249-259CODEN: AENVEQ; ISSN:1352-2310. (Elsevier Ltd.)Ammonia (NH3) and amines play important roles in the nucleation and growth of atm. aerosols. To identify the sources of these chems. in the densely populated and industrialized Yangtze River Delta region of China, we conducted measurements of NH3 and several amines, including methylamine (CH3NH2), C2-amines (C2H7N), and C3-amines (C3H9N) at a suburban site of Nanjing, China, during summer 2012. Using a high-resoln. time-of-flight chem. ionization mass spectrometer (HRToF-CIMS, Aerodyne), 1-min-averaged concns. of NH3 and amines ranged from a few parts per trillions by vol. (pptv) to dozens of ppb by vol. (ppbv). The av. ± 1σ concns. of NH3 and total amines during the measurement period were 1.7 ± 2.3 ppbv and 7.2 ± 7.4 pptv, resp. Among the amines, C2-amines were the most abundant, accounting for 54% of the total amine loading. Significant correlations between NH3 and all three types of amines (0.65 < r2 < 0.80) indicate similar emission sources. Anal. of meteorol. conditions indicated that these NH3 and amine laden air masses mainly originated from nearby industrial areas where NH3 was used for selective catalytic redn. of nitrogen oxides (NOx). The results of this work indicate that industrial emissions in Nanjing, China may have a significant impact on local and regional aerosol chem. by supplying considerable amt. of amines.
- 53Nair, A. A.; Yu, F. Quantification of Atmospheric Ammonia Concentrations: A Review of Its Measurement and Modeling. Atmosphere 2020, 11, 1092, DOI: 10.3390/atmos1110109253https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhslentL8%253D&md5=96c5c1f82ce9b9a9843ad6e175f3cf64Quantification of atmospheric ammonia concentrations: a review of its measurement and modelingNair, Arshad Arjunan; Yu, FangqunAtmosphere (2020), 11 (10), 1092CODEN: ATMOCZ; ISSN:2073-4433. (MDPI AG)Ammonia (NH3), the most prevalent alk. gas in the atm., plays a significant role in PM2.5 formation, atm. chem., and new particle formation. This paper reviews quantification of [NH3] through measurements, satellite-remote-sensing, and modeling reported in over 500 publications towards synthesizing the current knowledge of [NH3], focusing on spatiotemporal variations, controlling processes, and quantification issues. Most measurements are through regional passive sampler networks. [NH3] hotspots are typically over agricultural regions, such as the Midwest US and the North China Plain, with elevated concns. reaching monthly avs. of 20 and 74 ppbv, resp. Topog. effects dramatically increase [NH3] over the Indo-Gangetic Plains, North India and San Joaquin Valley, US. Measurements are sparse over oceans, where [NH3] ≈ a few tens of pptv, variations of which can affect aerosol formation. Satellite remote-sensing (AIRS, CrIS, IASI, TANSO-FTS, TES) provides global [NH3] quantification in the column and at the surface since 2002. Modeling is crucial for improving understanding of NH3 chem. and transport, its spatiotemporal variations, source apportionment, exploring physicochem. mechanisms, and predicting future scenarios. GEOS-Chem (global) and FRAME (UK) models are commonly applied for this. A synergistic approach of measurements↔satellite-inference↔modeling is needed towards improved understanding of atm. ammonia, which is of concern from the standpoint of human health and the ecosystem.
- 54Elm, J. An Atmospheric Cluster Database Consisting of Sulfuric Acid, Bases, Organics, and Water. ACS Omega 2019, 4 (6), 10965– 10974, DOI: 10.1021/acsomega.9b0086054https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXht1aksr%252FP&md5=ebba43602afc4822561711a19d343118An Atmospheric Cluster Database Consisting of Sulfuric Acid, Bases, Organics, and WaterElm, JonasACS Omega (2019), 4 (6), 10965-10974CODEN: ACSODF; ISSN:2470-1343. (American Chemical Society)We have collected, re-computed and compiled a database consisting of 633 unique atmospherically relevant mol. clusters contg. sulfuric acid, bases, oxidized org. compds. and water. The database comprise of strongly hydrogen bonded mol. clusters, and span both neutral, neg. charged and pos. charged clusters of atm. relevance. All the cluster structures and vibrational frequencies were re-evaluated at the ωB97X-D/6-31++G(d,p) level of theory and the single point energies were refined using a high level DLPNO-CCSD(T)/aug-cc-pVTZ calcn. The database unifies published atm. mol. clusters under a single common methodol. and serves as an efficient look-up table for mol. cluster structures and thermochem. parameters. Utilizing the database the performance of four semi-empirical methodologies (PM6, PM7, B97-3c and PBEh-3c) in calcg. the binding energies of atm. mol. clusters is assessed. The B97-3c and PBEh-3c empirically cor. DFT methods yield low errors in the binding energies compared to DLPNO-CCSD(T)/aug-cc-pVTZ ref. results and that a simple linear model can be utilized for estg. accurate binding energies based on ωB97X-D/6-31++G(d,p) results.
- 55Lee, S.-H.; Gordon, H.; Yu, H.; Lehtipalo, K.; Haley, R.; Li, Y.; Zhang, R. New particle formation in the atmosphere: From molecular clusters to global climate. J. Geophys. Res.: Atmos. 2019, 124, 7098– 7146, DOI: 10.1029/2018JD02935655https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhsVert7zF&md5=810a832a80a73b6ec7fada9f4299ee4eNew Particle Formation in the Atmosphere: From Molecular Clusters to Global ClimateLee, Shan-Hu; Gordon, Hamish; Yu, Huan; Lehtipalo, Katrianne; Haley, Ryan; Li, Yixin; Zhang, RenyiJournal of Geophysical Research: Atmospheres (2019), 124 (13), 7098-7146CODEN: JGRDE3; ISSN:2169-8996. (Wiley-Blackwell)New particle formation (NPF) represents the first step in the complex processes leading to formation of cloud condensation nuclei. Newly formed nanoparticles affect human health, air quality, weather, and climate. This review provides a brief history, synthesizes recent significant progresses, and outlines the challenges and future directions for research relevant to NPF. New developments include the emergence of state-of-the-art instruments that measure prenucleation clusters and newly nucleated nanoparticles down to about 1 nm; systematic lab. studies of multicomponent nucleation systems, including collaborative expts. conducted in the Cosmics Leaving Outdoor Droplets chamber at CERN; observations of NPF in different types of forests, extremely polluted urban locations, coastal sites, polar regions, and high-elevation sites; and improved nucleation theories and parameterizations to account for NPF in atm. models. The challenges include the lack of understanding of the fundamental chem. mechanisms responsible for aerosol nucleation and growth under diverse environments, the effects of SO2 and NOx on NPF, and the contribution of anthropogenic org. compds. to NPF. It is also crit. to develop instruments that can detect chem. compn. of particles from 3 to 20 nm and improve parameterizations to represent NPF over a wide range of atm. conditions of chem. precursor, temp., and humidity.
- 56Nahlovska, Z.; Nahlovsky, B.; Strand, T. G. Molecular Structure of Gaseous Oxalic Acid from Electron Diffraction and IR Data. Acta Chem. Scand. 1970, 24 (24), 2617– 2628, DOI: 10.3891/acta.chem.scand.24-2617There is no corresponding record for this reference.
- 57Redington, R. L.; Redington, T. E. Infrared Matrix-Isolation Spectra of Monomeric Oxalic Acid. J. Mol. Struct. 1978, 48, 165, DOI: 10.1016/0022-2860(78)80019-2There is no corresponding record for this reference.
- 58Back, R. A. The ultraviolet absorption spectrum of oxalic acid vapor. Can. J. Chem. 1984, 62, 1414, DOI: 10.1139/v84-241There is no corresponding record for this reference.
- 59Nieminen, J.; Rasanen, M.; Murto, J. Matrix Isolation and ab Initio Studies of Oxalic Acid. J. Phys. Chem. 1992, 96, 5303, DOI: 10.1021/j100192a02459https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38XktFOjs7Y%253D&md5=c24dcd0c7962acaec37f24739762f0f8Matrix-isolation and ab initio studies of oxalic acidNieminen, J.; Rasanen, M.; Murto, J.Journal of Physical Chemistry (1992), 96 (13), 5303-8CODEN: JPCHAX; ISSN:0022-3654.Rotamerization and UV photochem. decompn. of oxalic acid were studied in solid Ne, Ar, and Xe matrixes. The energies and spectra of the conformers and photoproducts were also simulated by extensive ab initio calcns. at MP2/6-31G** (energies, geometries) and MP2/4-31G* (numerical frequencies) levels. Two conformers were present in matrixes, the lowest energy species being cTc (with two intramol. H bonds) and species cTt (with only one intramol. H bond) being at ∼7 kJ mol-1 higher in energy. The uphill process cTc → cTt could be initiated by UV excitation at wavelengths of ∼270 nm. IR irradn. at wavenumbers >2000 cm-1 promoted the reverse interconversion. Full Xe arc irradn. decomps. oxalic acid in matrixes, the product yield ratio being different in different hosts. On the basis of ab initio calcns., the possible transition states for proton shifts are discussed. The calcd. vibrational spectra are given for six conformers of oxalic acid, the spectra of the obsd. conformers are assigned, and their potential energy distributions are given.
- 60Maçôas, E. M. S.; Fausto, R.; Pettersson, M.; Khriachtchev, L.; Rasanen, M. Infrared-Induced Rotamerization of Oxalic Acid Monomer in Argon Matrix. J. Phys. Chem. A 2000, 104, 6956, DOI: 10.1021/jp000634s60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXks1WntrY%253D&md5=c4f721a7f9f3c9260f45d9922984fcceInfrared-Induced Rotamerization of Oxalic Acid Monomer in Argon MatrixMacoas, Ermelinda M. S.; Fausto, Rui; Pettersson, Mika; Khriachtchev, Leonid; Raesaenen, MarkkuJournal of Physical Chemistry A (2000), 104 (30), 6956-6961CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)IR-induced conformational isomerization of oxalic acid monomer isolated in an argon matrix at 7.5 K was studied by IR spectroscopy. For the first time, three conformational states of this mol. were identified exptl. and their vibrational spectra are assigned. In good agreement with d. functional theory predictions, all the obsd. conformers exhibit a trans O:C-C:O axis, differing in the relative conformation of their O-C-O-H axes. In the most stable conformer (belonging to the C2h symmetry point group), two intramol. OH···O: hydrogen bonds are present. The second (Cs) most stable conformer shows a single OH···O: bond, and the third one (C2h) does not exhibit any intramol. hydrogen bond. Using narrowband tunable irradn. in the near-IR region it was possible to promote very efficiently conformer interconversions, which was followed spectroscopically.
- 61Godfrey, P. D.; Mirabella, M. J.; Brown, R. D. Structural Studies of Higher Energy Conformers by Millimeter-Wave Spectroscopy: Oxalic Acid. J. Phys. Chem. A 2000, 104, 258, DOI: 10.1021/jp992499t61https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXnvFKgt7s%253D&md5=24b59c4bf9ff2379b6d57e6f4454dd97Structural Studies of Higher Energy Conformers by Millimeter-Wave Spectroscopy: Oxalic AcidGodfrey, Peter D.; Mirabella, Mathew J.; Brown, Ronald D.Journal of Physical Chemistry A (2000), 104 (2), 258-264CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The authors have detected a polar conformer of oxalic acid and studied its structure using Stark-modulated free jet microwave spectroscopy. By comparing obsd. and ab initio predicted spectroscopic rotational consts. and deuteration isotopic substitution coordinates, the species was identified with the cis, trans, trans conformer (cTt), in which one of the carboxyl H atoms is involved in an intramol. H bond with the carbonyl O of the other carboxyl group. Stark effect measurements yielded elec. dipole moment values of μa = 2.850(6) D; μb = 1.148(3) D, μc = 0.0 D (assumed), and hence μtotal = 3.073(6) D. Ab initio calcns. at the MP2/6-311++G(d,p) level predict and expt. confirms that no other polar conformers occur in detectable concns. in the vapor at the expt. spectral search scan preexpansion temp. of 458 K. However, 2 undetectable species, one nonpolar (tTt), the other of low polarity (cTc, μ = 0.6 D) are predicted to be present in greater concns. than the obsd. species. Predicted relative stabilities of the different conformers are appreciably affected by the inclusion of Gibbs free energy corrections via the vibrational partition function. Large amplitude anharmonic vibrations, for which numerical integration of the vibrational Schrodinger equation is required for obtaining the correction, have an important influence on predicting relative stabilities of the oxalic acid conformers.
- 62Buemi, G. DFT study of the hydrogen bond strength and IR spectra of formic, oxalic, glyoxylic and pyruvic acids in vacuum, acetone and water solution. J. Phys. Org. Chem. 2009, 22, 933– 947, DOI: 10.1002/poc.154362https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtFGlsbfO&md5=8fc7874503b3dde8f679d889f25a6f0dDFT study of the hydrogen bond strength and IR spectra of formic, oxalic, glyoxylic and pyruvic acids in vacuum, acetone and water solutionBuemi, GiuseppeJournal of Physical Organic Chemistry (2009), 22 (10), 933-947CODEN: JPOCEE; ISSN:0894-3230. (John Wiley & Sons Ltd.)The mol. geometries of the possible conformations of formic, oxalic, glyoxylic and pyruvic acids have been fully optimized at DFT B3LYP/6-311++G(d,p) levels of calcn. in vacuum as well as in water and acetone soln. Solns. were treated according to the SCRF PCM approach but some formic acid-water and formic acid-acetone clusters as well as adducts of oxalic acid with two or four water mols. were also taken into account for testing the importance of specific solute-solvent effects. All the most stable isomers of the title compds. are characterized by weak intramol. hydrogen bonds, whose strengths (EHB) cannot be correctly estd. as stability difference between the open and chelate forms since the energy of the former isomer is, in turn, stabilized by a weak hydrogen bridge due to the formic acid moiety. Following the Rotation Barrier Method (RBM), proposed some years ago, EHB in the examd. mols. (gas phase) falls in the range of 18-22 kJ/mol for oxalic acid (9.6 kJ/mol for the c-C-t isomer), 16.8 kJ/mol for glyoxylic acid and 19.8 kJ/mol for pyruvic acid. Most of them disappear at all, or nearly at all, both in acetone and aq. soln., in consequence of the solvent effect. The frequencies of the OH and C-O stretching modes, calcd. according to the anharmonic oscillator model, are in very good agreement with the exptl. literature data, where available. Copyright © 2009 John Wiley & Sons, Ltd.
- 63Feller, D.; Bross, D. H.; Ruscic, B. Enthalpy of Formation of C2H2O4 (Oxalic Acid) from High-Level Calculations and the Active Thermochemical Tables Approach. J. Phys. Chem. A 2019, 123, 3481– 3496, DOI: 10.1021/acs.jpca.8b1232963https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXlvVSmtr0%253D&md5=21c3ad5929bd881d97a29842d28c7598Enthalpy of Formation of C2H2O4 (Oxalic Acid) from High-Level Calculations and the Active Thermochemical Tables ApproachFeller, David; Bross, David H.; Ruscic, BrankoJournal of Physical Chemistry A (2019), 123 (16), 3481-3496CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)High-level coupled cluster calcns. obtained with the Feller-Peterson-Dixon (FPD) approach and new data from the most recent version of the Active Thermochem. Tables (ATcT) are used to reassess the enthalpy of formation of gas-phase C2H2O4 (oxalic acid). The theor. value was further calibrated by comparing FPD and ATcT gas-phase enthalpies of formation for H2CO (formaldehyde) and the two low-lying conformations of C2H4O2 (syn and anti acetic acid). The FPD approach produces a theor. enthalpy of formation of gas-phase oxalic acid of -732.2 ± 4.0 kJ/mol at 298.15 K (-721.8 ± 4.0 kJ/mol at 0 K). An independently obtained ATcT value, based on reassessing the existent exptl. detns. and expanding the resulting thermochem. network with select mid-level composite theor. results, disagrees with several earlier recommendations that were based solely on exptl. detns. but is in excellent accord with the current FPD value. The inclusion of the latter in the most recent ATcT thermochem. network produces a further refined value for the gas-phase enthalpy of formation, -731.6 ± 1.2 kJ/mol at 298.15 K (-721.0 ± 1.2 kJ/mol at 0 K). The condensed-phase ATcT enthalpy of formation of oxalic acid is -829.7 ± 0.5 kJ/mol, and the resulting sublimation enthalpy is 98.1 ± 1.3 kJ/mol, both at 298.15 K.
- 64Chen, C.; Shyu, S.-F. Conformers and intramolecular hydrogen bonding of the oxalic acid monomer and its anions. Int. J. Quantum Chem. 2000, 76, 541– 551, DOI: 10.1002/(SICI)1097-461X(2000)76:4<541::AID-QUA5>3.0.CO;2-UThere is no corresponding record for this reference.
- 65Van Alsenoy, C.; Klimkowski, V. J.; Schafer, L. Ab initio studies of structural features not easily amenable to experiment: Part 37. Structural and conformational investigations of the dicarbonyls glyoxal, biacetyl and oxalic acid. J. Mol. Struct.: THEOCHEM 1984, 109, 321– 330, DOI: 10.1016/0166-1280(84)80016-0There is no corresponding record for this reference.
- 66Mohajeri, A.; Shakerin, N. The gas-phase acidity and intramolecular hydrogen bonding in oxalic acid. J. Mol. Struct.: THEOCHEM 2004, 711, 167, DOI: 10.1016/j.theochem.2004.10.00266https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXhtVClsLnF&md5=366f66e9144c2ffccbd139868ad02bf9The gas-phase acidity and intramolecular hydrogen bonding in oxalic acidMohajeri, A.; Shakerin, N.Journal of Molecular Structure: THEOCHEM (2004), 711 (1-3), 167-172CODEN: THEODJ; ISSN:0166-1280. (Elsevier B.V.)Ab initio calcns. at MP2/6-311++G and G2 level were carried out to study the structures and the gas-phase acidities of different conformations in oxalic acid. The gas-phase acidities (free energies of first ionization reactions) at G2 level vary from 1297.27 kJ/mol for the most acidic conformer to 1359.53 kJ/mol for the least acidic one at 298.15 K. Analyzing intramol. hydrogen bonding in different conformers, shows that the gas-phase acidity has a direct relation with hydrogen bond stabilization of both acid and its conjugated base. The results indicate that the reaction in which the acidic hydrogen is less involved and the hydrogen of its conjugated base is more involved in the intramol. hydrogen bond is the easiest deprotonation reaction. The atoms in mol. (AIM) theory has been also used to obtain the electronic charge distribution and some other related properties.
- 67Hermida-Ramón, J. M.; Cabaleiro-Lago, E. M.; Rodríguez-Otero, J. Computational study of the dissociation of oxalic acid in water clusters. Chem. Phys. 2004, 302, 53– 60, DOI: 10.1016/j.chemphys.2004.02.021There is no corresponding record for this reference.
- 68Chang, J. G.; Chen, H. T.; Xu, S. C.; Lin, M. C. Computational Study on the Kinetics and Mechanisms for the Unimolecular Decomposition of Formic and Oxalic Acids. J. Phys. Chem. A 2007, 111, 6789, DOI: 10.1021/jp069036p68https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXisVaisLo%253D&md5=bc140b40de3da68ce41189a5e3e6eb40Computational Study on the Kinetics and Mechanisms for the Unimolecular Decomposition of Formic and Oxalic AcidsChang, Jee-Gong; Chen, Hsin-Tsung; Xu, Shucheng; Lin, M. C.Journal of Physical Chemistry A (2007), 111 (29), 6789-6797CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The kinetics and mechanisms for the unimol. decompn. reactions of formic acid and oxalic acid have been studied computationally by the high-level G2M(CC1) method and microcanonical RRKM theory. There are two reaction pathways in the decompn. of formic acid: The dehydration process starting from the Z conformer is found to be the dominant, whereas the decarboxylation reaction starting from the E conformer is less competitive. The predicted rate consts. for the dehydration and decarboxylation reactions are in good agreement with the exptl. data. The calcd. CO/CO2 ratio, 13.6-13.9 between 1300 and 2000 K, is in close agreement with the ratio of 10 measured exptl. by Hsu et al. (In The 19th International Symposium on Combustion; The Combustion Institute: Pittsburgh, PA, 1983; p 89). For oxalic acid, its isomer with two intramol. hydrogen bonds is the most stable structure, similar to earlier reports. Two primary decompn. channels of oxalic acid producing CO2 + HOCOH with barriers of 33-36 kcal/mol and CO2 + CO + H2O with a barrier of 39 kcal/mol were found. At high temps., the latter process becomes more competitive. The rate const. predicted for the formation of CO2 and HOCOH (the precursor of HCOOH) agrees well with available exptl. data. The mechanism for the isomerization of HOCOH to HCOOH is also discussed.
- 69Higgins, J.; Zhou, X.; Liu, R.; Huang, T. T.-S. Theoretical Study of Thermal Decomposition Mechanism of Oxalic Acid. J. Phys. Chem. A 1997, 101, 2702– 2708, DOI: 10.1021/jp963819169https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXhslyqsro%253D&md5=2bb21e96f7d426594288668a58f46fcfTheoretical Study of Thermal Decomposition Mechanism of Oxalic AcidHiggins, James; Zhou, Xuefeng; Liu, Ruifeng; Huang, Thomas T.-S.Journal of Physical Chemistry A (1997), 101 (14), 2702-2708CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)D. functional theory B3LYP/6-31G** and ab initio MP2/6-31G** and MP4(SDQ)/6-311++G** calcns. were carried out to study the structures and isomerization and decompn. mechanisms of oxalic acid. The B3LYP structures and relative energies of the rotational isomers of oxalic acid are found very similar to MP2 results, confirming that the most stable isomer is the doubly intramol. hydrogen-bonded C2h structure E1, with four other planar isomers within 6 kcal/mol. It is predicted that unimol. formation of carbon dioxide and dihydroxycarbene (DHC) from oxalic acid has an activation barrier of 31 kcal/mol and that unimol. formation of HCOOH from DHC has an activation barrier about 31 kcal/mol higher. The unimol. formation of CO2, CO and H2O from oxalic acid via a concerted transition state has an activation barrier of only 42 kcal/mol, indicating it is a more favorable unimol. decompn. channel. On the other hand, hydrogen migration from oxygen to carbon of DHC to produce HCOOH can be accomplished through a hydrogen exchange with H2O (a model for oxalic acid) with an activation barrier of less than 37 kcal/mol. Transition state theory calcns. indicate that this bimol. channel might be responsible for the rapid formation of CO2 and HCOOH in gas phase oxalic acid thermal decompn., thus confirming the proposal of Bock and Redington. With increasing temp. the unimol. channel to produce CO2, CO, and H2O might become significant.
- 70Medeiros, F. S.; Oliveira, K. M. T.; Canuto, S.; Chaudhuri, P. A quantum chemical investigation of the interaction of perfluoropropionic acid with monoethanolamine and sulfuric acid in the atmosphere. Comput. Theor. Chem. 2024, 1233, 114485, DOI: 10.1016/j.comptc.2024.114485There is no corresponding record for this reference.
- 71de Oliveira, T. S.; Ghosh, A.; Chaudhuri, P. Exploring the hydrogen-bonded interactions of vanillic acid with atmospheric bases: a DFT study. Struct. Chem. 2024, 35, 1601– 1611, DOI: 10.1007/s11224-024-02307-3There is no corresponding record for this reference.
- 72Elm, J.; Mikkelsen, K. V. Computational approaches for efficiently modelling of small atmospheric clusters. Chem. Phys. Lett. 2014, 615, 26– 29, DOI: 10.1016/j.cplett.2014.09.06072https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhs12ntL3F&md5=087d4f5b68da4e88ac9ae535a1bcb826Computational approaches for efficiently modelling of small atmospheric clustersElm, Jonas; Mikkelsen, Kurt V.Chemical Physics Letters (2014), 615 (), 26-29CODEN: CHPLBC; ISSN:0009-2614. (Elsevier B.V.)Utilizing a comprehensive test set of 205 clusters of atm. relevance, we investigate how different DFT functionals (M06-2X, PW91, ωB97X-D) and basis sets (6-311++G(3df,3pd), 6-31++G(d,p), 6-31+G(d)) affect the thermal contribution to the Gibbs free energy and single point energy. Reducing the basis set used in the geometry and frequency calcn. from 6-311++G(3df,3pd) → 6-31++G(d,p) implies a significant speed-up in computational time and only leads to small errors in the thermal contribution to the Gibbs free energy and subsequent coupled cluster single point energy calcn.
- 73Elm, J.; Bilde, M.; Mikkelsen, K. V. Assessment of binding energies of atmospherically relevant clusters. Phys. Chem. Chem. Phys. 2013, 15, 16442– 16445, DOI: 10.1039/c3cp52616j73https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsVCru7jP&md5=64f46480e82da3faf993770c931335feAssessment of binding energies of atmospherically relevant clustersElm, Jonas; Bilde, Merete; Mikkelsen, Kurt V.Physical Chemistry Chemical Physics (2013), 15 (39), 16442-16445CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)This work assesses the binding energies of atmospherically relevant clusters contg. H2SO4, H2O, NH3 and (CH3)2NH using d. functional theory. The performance of seven DFT functionals (B3LYP, CAM-B3LYP, M06-2X, PW91, LC-PW91, PBE0 and ωB97X-D) is evaluated against high level explicitly correlated coupled cluster methods using a test set of 107 atmospherically relevant clusters. Our studies show that all the tested functionals correlate well with the coupled cluster results, but with highly varying mean abs. errors. The PBE0, CAM-B3LYP, PW91 and M06-2X functionals are found to perform similarly with errors in the range of 2.53-3.46 kcal mol-1, while the B3LYP and LC-PW91 functionals yield higher errors of 6.95 kcal mol-1 and 10.66 kcal mol-1, resp. The ωB97X-D functional gives the best est. of the binding energies with a mean abs. error as low as 2.12 kcal mol-1 over the large test set of clusters.
- 74Elm, J.; Bilde, M.; Mikkelsen, K. V. Assessment of Density Functional Theory in Predicting Structures and Free Energies of Reaction of Atmospheric Prenucleation Clusters. Phys. Chem. Chem. Phys. 2012, 8, 2071– 2077, DOI: 10.1021/ct300192pThere is no corresponding record for this reference.
- 75Huff, A. K.; Mackenzie, R. B.; Smith, C. J.; Leopold, K. R. A Perfluorinated Carboxylic Sulfuric Anhydride: Microwave and Computational Studies of CF3COOSO2OH. J. Phys. Chem. A 2019, 123, 2237– 2243, DOI: 10.1021/acs.jpca.9b0030075https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXitFGht70%253D&md5=6f9d1a1c8867656a8507b8e9b9594c89A Perfluorinated Carboxylic Sulfuric Anhydride: Microwave and Computational Studies of CF3COOSO2OHHuff, Anna K.; MacKenzie, Rebecca B.; Smith, C. J.; Leopold, Kenneth R.Journal of Physical Chemistry A (2019), 123 (11), 2237-2243CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Trifluoroacetic sulfuric anhydride (CF3COOSO2OH, TFASA) and its deuterated isotopologue have been obsd. by pulsed-nozzle Fourier transform microwave spectroscopy. TFASA was generated in situ in a supersonic expansion from the reaction of CF3COOH or CF3COOD with SO3. The spectrum, which was notably weaker than those of previously studied carboxylic sulfuric anhydrides, is that of a simple asym. rotor with no evidence of internal rotation of the CF3 group. Calcns. at the M06-2X/6-311++G(3df,3pd) level indicate that the title compd. is produced via a mechanism involving a concerted cycloaddn., analogous to that found for other carboxylic sulfuric anhydrides. The calcns. further show that the equil. orientation of CF3 relative to the C=O bond changes upon formation of the anhydride, indicating that any path connecting the equil. structures of CF3COOH and CF3COOSO2OH necessarily includes both cycloaddn. and internal rotation. CCSD(T)/complete basis set with double and triple extrapolation [CBS(D-T)] single-point energy calcns. at key points on the potential surface indicate that the barrier to form TFASA from a putative CF3COOH···SO3 complex is about 1.2 kcal/mol after zero-point energy corrections. This value is significantly larger than the near-zero or slightly neg. barriers previously reported for the reactions of SO3 with nonfluorinated carboxylic acids and likely accounts, at least in part, for the reduced spectral intensity. Thus, TFASA is a somewhat unique addn. to the series of carboxylic sulfuric anhydrides studied to date. Theor. values of certain structural parameters, at. charges, and vibrational frequencies also support this point of view. Despite the differences, however, this work clearly demonstrates that the reaction RCOOH + SO3 → RCOOSO2OH readily occurs in the gas phase and is not restricted to acids with hydrocarbon R groups.
- 76Dennington, R. D.; Keith, T. A.; Millam, J. M. GaussView 5.0.9; Gaussian Inc.: Wallingford, CT, USA, 2008.There is no corresponding record for this reference.
- 77Kubečka, J.; Besel, V.; Kurtén, T.; Myllys, N.; Vehkamäki, H. Configurational sampling of noncovalent (atmospheric) molecular clusters: Sulfuric acid and guanidine. J. Phys. Chem. A 2019, 123, 6022– 6033, DOI: 10.1021/acs.jpca.9b0385377https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXht1eks7fL&md5=941e3ebf795469187f47126a4c65db3cConfigurational Sampling of Noncovalent (Atmospheric) Molecular Clusters: Sulfuric Acid and GuanidineKubecka, Jakub; Besel, Vitus; Kurten, Theo; Myllys, Nanna; Vehkamaki, HannaJournal of Physical Chemistry A (2019), 123 (28), 6022-6033CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)We studied the configurational sampling of noncovalently bonded mol. clusters relevant to the atm. In this article, we discuss possible approaches to searching for optimal configurations and present one alternative based on systematic configurational sampling, which seems able to overcome the typical problems assocd. with searching for global min. on multidimensional potential energy surfaces. Since atm. mol. clusters are usually held together by intermol. bonds, we also present a cost-effective strategy for treating hydrogen bonding and proton transferring by using rigid mols. and ions in different protonation states and illustrate its performance on clusters contg. guanidine and sulfuric acid.
- 78Partanen, L.; Vehkamäki, H.; Hansen, K.; Elm, J.; Henschel, H.; Kurtén, T.; Halonen, R.; Zapadinsky, E. Effect of Conformers on Free Energies of Atmospheric Complexes. J. Phys. Chem. A 2016, 120, 8613– 8624, DOI: 10.1021/acs.jpca.6b0445278https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xhs1SksbfM&md5=1390239ffe63c609adfbfd2561f5f63fEffect of Conformers on Free Energies of Atmospheric ComplexesPartanen, Lauri; Vehkamaki, Hanna; Hansen, Klavs; Elm, Jonas; Henschel, Henning; Kurten, Theo; Halonen, Roope; Zapadinsky, EvgeniJournal of Physical Chemistry A (2016), 120 (43), 8613-8624CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)In this article we show how to calc. free energies for atmospherically relevant complexes when multiple conformers and/or isomers are present. We explain why the thermal averaging methods used in several published works are not correct. Based on our two sample cases, the sulfuric acid - pinic acid complex and the (H2SO4)3(NH3)3(H2O)4 cluster, we provide numerical evidence that the use of these incorrect formulas can result in errors larger than one kcal/mol. We recommend that if vibrational frequencies and thus Gibbs free energies of the individual conformers are unavailable, one should not attempt to correct for the presence of multiple conformers and instead use only the global min. conformers for both reactants and products. On the other hand, if the free energies for the conformers are calcd. for both reactants and products, their effect can be accounted for by the statistical mech. methods presented in this article.
- 79Fileti, E. E.; Rivelino, R.; Canuto, S. Rayleigh light scattering of hydrogen bonded clusters investigated by means of ab initio calculations. J. Phys. B: At., Mol. Opt. Phys. 2003, 36, 399– 408, DOI: 10.1088/0953-4075/36/2/31979https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXntFGjtrs%253D&md5=5dbaf9c98e81fd7b512c06efdf3875b8Rayleigh light scattering of hydrogen bonded clusters investigated by means of ab initio calculationsFileti, Eudes E.; Rivelino, Roberto; Canuto, SylvioJournal of Physics B: Atomic, Molecular and Optical Physics (2003), 36 (2), 399-408CODEN: JPAPEH; ISSN:0953-4075. (Institute of Physics Publishing)Ab initio calcns. of depolarization ratios and intensities of classically scattered light, in terms of dipole polarizabilities and polarizability anisotropies, are reported for different H bonded mol. clusters. Five different groups of org. heterodimers formed with H2O are considered: HCHO···H2O, MeHO···H2O, HCOOH···H2O, MeCN···H2O, and Me2CO···H2O, together with the H2O dimer H2O···H2O. The geometries of all complexes were optimized by the 2nd-order Moller-Plesset many-body perturbation theory (MP2), using the augmented correlation-consistent basis set with polarized valence of double-zeta quality (aug-cc-pVDZ). The calcd. av. dipole polarizabilities of the isolated mols. are in good agreement with available exptl. results. The calcns. are then extended to the complexes and, from these, the Rayleigh scattering activities and depolarization ratio changes, upon H bond formation, are obtained and analyzed. The differences in activity and depolarization for Rayleigh scattered radiation between 2 groups of isomers, (i) HCN···H2O and H2O···HCN and (ii) MeHO···H2O and MeOH···OH2, also were studied.
- 80Chaudhuri, P.; Canuto, S. Rayleigh scattering properties of small polyglycine molecules. J. Mol. Struct. 2006, 760, 15– 20, DOI: 10.1016/j.theochem.2005.10.03980https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xit1WrsLo%253D&md5=8b44aec03394ddf333e2902da7d286ffRayleigh scattering properties of small polyglycine moleculesChaudhuri, Puspitapallab; Canuto, SylvioJournal of Molecular Structure: THEOCHEM (2006), 760 (1-3), 15-20CODEN: THEODJ; ISSN:0166-1280. (Elsevier B.V.)D.-functional theory calcns. of the Rayleigh optical activities of small isolated polyglycine mols. are reported. Fully extended β-sheet-like conformations of polypeptides of glycine, (Gly)n (with n = 1-5) are considered. After geometry optimizations, dipole moments and dipole polarizabilities (both the mean and the anisotropic components) are calcd. using the B3LYP and B3P86 functionals in 3 basis sets. The polarizabilities are used to analyze the Rayleigh scattering activities and depolarization ratios. The convergence of the av. dipole polarizability per monomer is analyzed. The differences in activity and depolarization for Rayleigh scattered radiation between the extended β-sheet-like and the folded α-helix-like forms of tetraglycine are analyzed and are relevant, suggesting its possible use in exptl. characterization.
- 81da Silva, A. M.; Chakraborty, S.; Chaudhuri, P. Rayleigh light scattering from hydrogen-bonded dimers of small astrophysical molecules. Int. J. Quantum Chem. 2012, 112, 2822– 2827, DOI: 10.1002/qua.23303There is no corresponding record for this reference.
- 82Elm, J.; Norman, P.; Bilde, M.; Mikkelsen, K. V. Computational study of the Rayleigh light scattering properties of atmospheric prenucleation clusters. Phys. Chem. Chem. Phys. 2014, 16, 10883– 10890, DOI: 10.1039/C4CP01206BThere is no corresponding record for this reference.
- 83Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H., Gaussian 16, Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2016.There is no corresponding record for this reference.
- 84Elm, J.; Ayoubi, D.; Engsvang, M.; Jensen, A. B.; Knattrup, Y.; Kubečka, J.; Bready, C. J.; Fowler, V. R.; Harold, S. E.; Longsworth, O. M. Quantum chemical modeling of organic enhanced atmospheric nucleation: A critical review. WIREs Comput. Mol. Sci. 2023, 13 (5), e1662 DOI: 10.1002/wcms.1662There is no corresponding record for this reference.
Supporting Information
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.4c06290.
Containing the optimized bond lengths and bond angles of different oxalic acid conformers; the Cartesian coordinates of the optimized geometries of binary and ternary clusters of OA with AM and SA; relevant structural parameters related to hydrogen bond formation in (OA)(SA) and (OA)(SA)2 clusters, supported by the ternary cluster images (Figure S1); successive binding free energies (ΔGS) for the formation of various (OA)(SA)2 ternary clusters and Boltzman-averaged values of Rayleigh scattering intensities of the OA conformers and their binary and ternatry clusters with AM and SA (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.