You are on page 1of 15

Subscriber access provided by Eastern Michigan University | Bruce T.

Halle Library

Surfaces, Interfaces, and Catalysis; Physical Properties of Nanomaterials and Materials


Increased Acid Dissociation at the Quartz/Water Interface
Shivam Parashar, Dominika Lesnicki, and Marialore Sulpizi
J. Phys. Chem. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.jpclett.8b00686 • Publication Date (Web): 10 Apr 2018
Downloaded from http://pubs.acs.org on April 10, 2018

Just Accepted

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

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 14 The Journal of Physical Chemistry Letters

1
2
3
4
5
6
7
8 Increased Acid Dissociation at the
9
10
11
12
Quartz/Water Interface
13
14
15 Shivam Parashar,† Dominika Lesnicki,‡ and Marialore Sulpizi∗,‡
16
17
18 †Department of Chemical Engineering, Indian Institute of Technology Roorkee, India.
19
20 ‡Institute of Physics, Johannes Gutenberg University Mainz, Staudingerweg 7, 55099
21
22 Mainz, Germany
23
24
25 E-mail: sulpizi@uni-mainz.de
26
27 Phone: +49 6131 39 23641. Fax: +49 6131 39 25441
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 1
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Letters Page 2 of 14

1
2
3
Abstract
4
5
6 As shown by a quite significant amount of literature, acids at the water surface tends
7
8 to be "less" acid, meaning that their associated form is favored over the conjugated base.
9
10 What happens at the solid/liquid interface? In the case of the silica/water interface
11
we show how the acidity of adsorbed molecules can instead increase. Using a free
12
13 energy perturbation approach in combination with electronic structure -based molecular
14
15 dynamics simulations, we show how the acidity of pyruvic acid at the quartz/water
16
17 interface is increased of almost two units. Such increased acidity is the result of the
18
19 specific microsolvation at the interface and in particular of the stabilization of the
20
21 deprotonated form by the silanols, on the quartz surface, and the special interfacial
22
23 water layer.
24
25
26
27 Graphical TOC Entry
28
29
30 Pyruvic acid
31
32
33 pKa = -1.85
34
35
36 Quartz 0001

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 2
60 ACS Paragon Plus Environment
Page 3 of 14 The Journal of Physical Chemistry Letters

1
2
3
4
Introduction
5
6 The acid dissociation constants of molecules in bulk aqueous solution are routinely deter-
7
8 mined using techniques like potentiometric titration, voltammetry, and electrophoresis. 1 A
9
10 more complicated task is to measure dissociation constants at interfaces since the signal orig-
11
12 inating from bulk molecules typically overwhelms that from surface molecules. The advent
13
14 of interface selective vibrational spectroscopy techniques, capable to probe the molecular
15
16 structure of the nanometric layer of a liquid interface, has recently permitted to selectively
17
18 address properties of molecules at interfaces, also providing information on the degree of
19
20 proton dissociation.
21
22 Pioneering interface selective spectroscopic studies of acid/base pairs have shown that
23
24 the surface favors the neutral form of the acid/base pair in comparison to the bulk. 2 For in-
25
26 stance, using second harmonic generation (SHG) and vibrational Sum Frequency Generation
27
28 (VSFG), it has been shown for phenol and carboxylic acids that the neutral acid species is
29
30 favored over the anionic conjugate base. 3–5 Similarly, it was found for molecules containing
31
32 an acid ammonium group that the neutral amine base species was favored over the cationic
33
34 quaternary ammonium acid. 6,7 Nitric acid, known as a strong acid in bulk aqueous solution
35
36 was also investigated at the water/air surface. Its dissociation does not occurs at the water
37
38 surface while generally occurs in the first and second surface layers depending on the acid’s
39
40 orientation and its solvation. 8–11 Amino acids, such as L-alanine and L-proline studied by
41
42 VSFG present a reduced acid dissociation at the water surface. Higher pH values are re-
43
44 quired to induce the dissociation of carboxylic acid groups of molecule at surface. 12 Tahara
45
46 and coworkers have evaluated the pH at model biological membranes (lipid/water interfaces),
47
48 which substantially deviates from the bulk pH. The pH at the lipid/water interface is higher
49
50 than that in the bulk when the head group of the lipid is positively charged, whereas the pH
51
52 at the lipid/water interface is lower when the lipid has a negatively charged head group. 13
53
54 Now the question is what happens instead at the solid/liquid interface? The solid surface
55
56 may contain hydrophilic groups (such as silanols in the case of quartz and silica), as well as
57
58
59 3
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Letters Page 4 of 14

1
2
3
ions, which can influence the dissociation constant. Can the situation effectively be reversed?
4
5
Are hydrophilic surface enhancing the acid dissociation? Experimental investigation of acid/
6
7
base equilibrium at "buried" solid/liquid interface are more difficult. The group, e.g. of V.
8
9 Grassian has been extensively studying adsorption and dissociation of strong and weak acids
10
11 on silica particles, 14,15 however it is challenging for such experiments to quantify a change in
12
13 the acidity upon adsorption at the solid/liquid interface. Some recent computational studies
14
15 in the group of R.B. Gerber have evidenciated fast deprotonation of HCl and H2 SO4 on quartz
16
17 surface covered by a water monolayer, 16,17 although, again, a quantitative measument of the
18
19 acidity on such surface (and in comparison to liquid bulk conditions) was not provided.
20
21 The group of D. Marx has recently addressed with accurate free energy calculations, also
22
23 including nuclear quantum effects, the issue of acid dissociation in HCl-water clusters, and
24
25 has shown that, a minimum of four water molecules is needed to stabilize the fully dissociated
26
27 solvent-shared ion pair at low-temperature and thus to create the smallest droplet of acid. 18
28
29 In order to answer the question if adsorption at the solid/liquid interface can enhance acid
30
31 dissociation, we will discuss in details the dissociation of pyruvic acid, the simplest α-keto
32
33 acid, at silica/water interface. Pyruvic acid plays a key role in the intricate chemistry of the
34
35 atmosphere as intermediate in the oxidative channels of isopropene and of secondary organic
36
37 aerosols. 19 Silica is the main component of mineral dust particles in the atmosphere, and
38
39 reactions on its surface, in presence of variable levels of humidity, and also including proton
40
41 dissociation, have a direct impact on radical concentration and the ability of dust particles
42
43 to serve as cloud condensation nuclei or ice nuclei. 20 Noticeably, the dissociation constant
44
45 of pyruvic acid (bulk pKa 2.5) is close to the point of zero charge of α-quartz surface, 21,22
46
47 therefore pyruvic acid and silanols may compete for protons at the quartz surface. For
48
49 example, we have shown that on the quartz 0001 surface two species of silanols exists, with
50
51 the most acidic one, represented by the out-of-plane silanols presenting a pKa of 5.6. 21
52
53 Similarly at the amorphous silica/ water interface we also found that some convex geminals
54
55 groups, as well as some type of vicinals are very acid with pKa of 2.9 and 2.1, respectively, 23
56
57
58
59 4
60 ACS Paragon Plus Environment
Page 5 of 14 The Journal of Physical Chemistry Letters

1
2
3
therefore very close to the pKa of bulk pyruvic acid of 2.5.
4
5
Anticipating our results, we find that pyruvic acid acidity is enhanced at the quartz/water
6
7
interface, of almost two units. This result is quite interesting, since it goes in the opposite
8
9 direction of what normally found for acids at the water/vapor interface. As we will explain
10
11 in details in this Letter, such increased acidity is the results of the specific micro-solvation
12
13 around the acid/abase groups and, in particular, of the stabilization of the conjugated base
14
15 by the silanols groups on the quartz surface.
16
17 In order to obtain a detailed microscopic understanding of the interface pKa variations,
18
19 molecular modeling at the atomistic level is required. In particular, using ab initio molecular
20
21 dynamics simulations, the acidity constants will be computed using the reversible proton
22
23 insertion/deletion method 24,25 that we already applied to a wide range of chemical groups,
24
25 also at solid/liquid interface. 21,23,26,27 Since our approach permits to explicitly take into
26
27 account the solvent and its specific interaction with the molecule as well as with the surface,
28
29 we will be able to pin down the origin of the pKa changes in terms of molecular orientation
30
31 and local hydrogen bond strength. The inclusion of the full electronic structure will permit
32
33 to address anharmonicity and polarization effects, which cannot be easily incorporated in a
34
35 simpler empirical force field approach.
36
37 We compute the pKa difference between the acid at the interface and the acid in the bulk,
38
39 using a thermodynamics integration which transforms the acid into its conjugated base. We
40
41 consider the α-quartz 0001 surface in contact with water in the neutral, fully hydroxilated
42
43 form. As schematically shown in Fig.1 we actually transform a model system which contains
44
45 the acid (AH) at the interface and the conjugated base (A-) in the bulk (left panel in Fig.1)
46
47 into a system which contains the base (A-) at the interface and the acid (AH) in the bulk
48
49 (right panel in Fig.1).
50
51 The difference in pKa ’s between interface and bulk (∆pKa = pKaI − pKaB ) is then given
52
53
54
55
56
57
58
59 5
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Letters Page 6 of 14

1
2
3 AH =
4
5
6
7
8 A- AH
9
10 A- =
AH A-
11
12
13
14 (a) E0 (b) E1
15
16 Figure 1: The model used for the calculation of ∆pKa value. On the left side the protonated
17
18 form of pyruvic acid (AH) and the deprotonated form (A− ) are shown. Schematic view of
19 the simulation setup (a) the initial state with protonated acid (AH) at the interface and
20 deprotonated acid (A-) in the bulk. (b) the final state with the acid in the bulk (AH) and
21 the deprotonated acid (A-) at the interface.
22
23
24
25 by the following equation:
26
27
28 (∆dp AAH )I − (∆dp AAH )B
29 ∆pKa = pKaI − pKaB = , (1)
30 ln10kB T
31
32
33 where (∆dp AAH )I,B is the deprotonation free energy for interface (I) and bulk (B), respec-
34
35 tively, and can be calculated with the thermodynamic integration approach. We remand the
36
37
reader to the Supplementary Information for all the details of the method implementation.
38
39
DFT based Born-Oppenheimer molecular dynamics (DFTMD) simulations were per-
40
41
formed using Becke exchange 28 and Lee, Yang and Parr 29 correlation functionals. All calcu-
42
43
lations have been carried out with the freely available DFT package CP2K/Quickstep which
44
45
is based on the hybrid Gaussian and plane wave method. 30 Analytic Goedecker-Tecter-Hutter
46
47
pseudopotentials, 31,32 a TZV2P level basis set for the orbitals and a density cutoff of 280
48
49
Ry were used. Dynamics were conducted in the canonical NVT ensemble with a timestep
50
51
of 0.5 fs using the CSVR thermostat with the target temperature of 330 K to avoid the
52
53
glassy behavior of BLYP water and a time constant of 1 ps. A periodic orthorhombic box
54
55
of dimensions 19.64×17.008×43.5 Å is considered (see Supplementary Material for details
56
57
58
59 6
60 ACS Paragon Plus Environment
Page 7 of 14 The Journal of Physical Chemistry Letters

1
2
3
of the system preparation). For each simulation, an equilibrium trajectory of 5 ps and an
4
5
additional production run of 20 ps was used for the ∆pKa calculation.
6
7
For both bulk and interface the most stable Tce isomer was considered. At the interface
8
9 the acid is stabilized in a configuration with the heavy atoms plane parallel to the quartz
10
11 surface. The acid remains localized within the first adsorbed layer of water, as shown by
12
13 the position of the center of mass in Fig. 3 of the SI. In the protonated form the binding
14
15 between pyruvic acid and the surface silanols is not direct, but mediated by a water bridge
16
17 (see for example a snapshot from the simulation in Fig. 4 (b)). Additionally the methyl
18
19 group of pyruvic acid also accommodates on the hydrophobic patch on the quartz surface
20
21 (see Fig. 2 for a top view perspective). Such binding configuration is quite stable along the
22
23 simulation trajectory as shown by the time traces reported in Fig. 2. Essentially pyruvic
24
25 acid acid cannot freely diffuse at the interface, but remains quite strongly localized.
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48 Figure 2: Top view of the (0001) α-quartz interface where the pyruvic acid is located (position
49 in Å). Dots represent the average position of the Si (yellow), O (red) and H (black) atoms
50 from the quartz interface. The lines represent the time traces (position as function of time)
51
of the center of mass (cyan), the methyl group (magenta), the oxygen (red) and hydrogen
52
53 (black) atoms from the pyruvic acid and the oxygen atom (blue) from the water bridge.
54
55
56 From our calculation we find that the interfacial pyruvic acid is more acidic than the one
57
58
59 7
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Letters Page 8 of 14

1
2
3
in the bulk by 1.84 pKa units. The easier deprotonation at the quartz/water interface can
4
5
be understood through an analysis of the solvation structure around both protonated and
6
7
deprotonated species. In particular we will discuss the solvation structure in terms of radial
8
9 distribution functions and coordination numbers, specifically defined for the solid/liquid
10
11 interface as described in the Supplementary Material.
12
13 3 3
14
(a) HAH - Ow (b) OA_ - Hw
15
16 HAH - Ow OA_ - Hw 6
17 OA_ - HSi
18 2 2
19
20 4
rdf

rdf
21
22
1 1
23
2
24
25
26
27 0 0 0
28
0 2 4 6 8 0 2 4 6 8
29 r (Å) r (Å)
30
31
Figure 3: Radial distribution functions between (a) the hydrogen atom of the protonated
32 pyruvic acid (HAH ) and the oxygen atoms of the water solvent (Ow ) (acid in bulk in black,
33 acid at the interface in red), (b) the oxygen atom of the deprotonated pyruvic acid (OA− )
34 and the hydrogen atoms of the water solvent (Hw ) (conjugated base in bulk in black, at the
35 interface in red. The rdf between OA− and the silanol proton (HSi ) is also reported (red
36
37
dashed line, y axis values readable on the right).
38
39
40 In the protonated form the solvation structure around the OH group is similar when
41
42 comparing bulk and interface. In both cases the OH group donates a relatively strong
43
44 hydrogen bond, which however is slightly shorter at the interface (red line in Fig. 3, position
45
46 of the first peak at 1.61 Å) than in the bulk (black line in Fig. 3, position of the first peak
47
48 at 1.71 Å). This first observation would suggest that at the interface the proton is more
49
50 keen to leave its oxygen. Such a difference in the local solvation is the result of the different
51
52 water behavior in the bulk and at the interface. At the interface the interaction with the
53
54 hydrophilic surface renders the water molecules in the first adsorbed monolayer quite special.
55
56 Such waters have indeed much slower orientational and diffusion dynamics than those in bulk
57
58
59 8
60 ACS Paragon Plus Environment
Page 9 of 14 The Journal of Physical Chemistry Letters

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17 (a) (b)
18
19
20
21
22
23
24
25
26
27
28 (c) (d)
29
30
31
32
33
34
35
36
37
38
39 Figure 4: The H-bonds formed between the protonated/deprotonated pyruvic acid and the
40
surrounding atoms: (a) protonated acid in the bulk water, (b) protonated acid at the in-
41
42 terface, (c) deprotonated acid in the bulk water and (d) deprotonated acid at the interface.
43 Color code: Si, yellow; C, blue; O, red, H, white.
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 9
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Letters Page 10 of 14

1
2
3
water. As a result the hydrogen bond established between the interfacial water molecule and
4
5
the OH group of pyruvic acid has a life time, larger than 20 ps, which even exceeds the
6
7
simulation length.
8
9 When we compare the behavior of the conjugated base (at the interface and in the
10
11 bulk) we also notice interesting differences. Indeed while in the bulk the conjugated base is
12
13 stabilized accepting hydrogen bonds from water (with an average number of Hbonds in the
14
15 first shell of 2.64), at the interface it is stabilized by hydrogen Hbonds from water (on average
16
17 1.87), but additionally also by accepting hydrogen bonds from a surface silanol (on average
18
19 0.58). The hydrogen bond established with the silanol (dashed red line in Fig.3) is shorter
20
21 (1.69 Å) than that accepted by water (continuous red line in Fig.3(b), position of the first
22
23 peak at 1.85 Å). The shorter hydrogen bond is due to specific properties of the OH group in
24
25 the silanols. As we have already shown in our previous work, 21,23 silanols may have acidity
26
27 constants as low as 5, substantially lower than water, in turn establishing stronger hydrogen
28
29 bonds than water. The silanols have therefore a special role in stabilizing the conjugated
30
31 base and therefore in favoring the deprotonated form of pyruvic acid at the interface.
32
33 The solvation behavior observed for pyruvic acid at the quartz/water interface is certainly
34
35 not unique, and indeed a similar mechanism for the conjugated base stabilization could be
36
37 in place also for other acids at hydroxilated interfaces. Of course the propensity to localize
38
39 at the interface is strongly depending on the degree of hydrophilicity and hydrophobiticity
40
41 of the acid as well as that of the solid surface.
42
43
44
45
46 Acknowledgement
47
48
49 This work was supported by the Deutsche Forschungsgemeinschaft (DFG) TRR146, project
50
51 A4. All the calculation were performed on the supercomputer of the High Performance
52
53 Computing Center (HLRS) of Stuttgart (grant 2DSFG).
54
55
56
57
58
59 10
60 ACS Paragon Plus Environment
Page 11 of 14 The Journal of Physical Chemistry Letters

1
2
3
4
Supporting Information Available
5
6 • Theory on the pKa calculation using thermodynamics integration
7
8
9 • Simulation details
10
11
12 • Vertical energy gaps
13
14
15 • Radial distribution function in presence of an interface
16
17
18
19 References
20
21
22 (1) Reijenga, J.; van Hoof, A.; van Loon, A. Development of Methods for the Determination
23
24 of pKa Values. Anal. Chem. Insights 2013, 8, 53.
25
26
27 (2) Eisenthal, K. Liquid Interfaces Probed by Second-Harmonic and Sum-Frequency Spec-
28
29 troscopy. Chem. Rev. 1996, 96, 1343–1360.
30
31
32
(3) Miranda, P.; Du, Q.; Shen, Y. Interaction of water with a fatty acid Langmuir film.
33
34
Chem. Phys. Lett. 1998, 286, 1–8.
35
36 (4) Rao, Y.; Subir, M.; McArthur, E. A.; Turro, N. J.; Eisenthal, K. B. Organic Ions at
37
38 the Air/Water Interface. Chem. Phys. Lett. 2009, 477, 241–244.
39
40
41 (5) Tang, C. Y.; Huang, Z.; Allen, H. C. Binding of Mg2+ and Ca2+ to Palmitic Acid
42
43 and Deprotonation of the COOH Headgroup Studied by Vibrational Sum Frequency
44
45 Generation Spectroscopy. J. Phys. Chem. B 2010, 114, 17068–17076.
46
47
48 (6) Zhao, X.; Ong, S.; Wang, H.; Eisenthal, K. B. New method for determination of surface
49
50 pKa using second harmonic generation. Chem. Phys. Lett. 1993, 214, 203–207.
51
52
(7) Wang, H.; Zhao, X.; Eisenthal, K. B. Effects of Monolayer Density and Bulk Ionic
53
54
Strength on Acid-Base Equilibria at the Air/Water Interface. J. Phys. Chem. B 2000,
55
56
104, 8855–8861.
57
58
59 11
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Letters Page 12 of 14

1
2
3
(8) Shamay, E. S.; Buch, V.; Parrinello, M.; Richmond, G. L. At the Water’s Edge: Nitric
4
5
Acid as a Weak Acid. Journal of the American Chemical Society 2007, 129, 12910–
6
7
12911, PMID: 17915872.
8
9
10 (9) Kido Soule, M. C.; Blower, P. G.; Richmond, G. L. Nonlinear Vibrational Spectroscopic
11
12 Studies of the Adsorption and Speciation of Nitric Acid at the Vapor/Acid Solution In-
13
14 terface. The Journal of Physical Chemistry A 2007, 111, 3349–3357, PMID: 17419597.
15
16
17 (10) Wang, S.; Bianco, R.; Hynes, J. T. Depth-Dependent Dissociation of Nitric Acid at an
18
19 Aqueous Surface: Car-Parrinello Molecular Dynamics. The Journal of Physical Chem-
20
21 istry A 2009, 113, 1295–1307, PMID: 19173580.
22
23
24 (11) Wang, S.; Bianco, R.; Hynes, J. T. Dissociation of nitric acid at an aqueous surface:
25
26
Large amplitude motions in the contact ion pair to solvent-separated ion pair conver-
27
28
sion. Phys. Chem. Chem. Phys. 2010, 12, 8241–8249.
29
30 (12) Strazdaite, S.; Meister, K.; Bakker, H. J. Reduced Acid Dissociation of Amino-Acids at
31
32 the Surface of Water. Journal of the American Chemical Society 2017, 139, 3716–3720,
33
34 PMID: 28177623.
35
36
37 (13) Kundu, A.; Yamaguchi, S.; Tahara, T. Evaluation of pH at Charged Lipid/Water Inter-
38
39 faces by Heterodyne-Detected Electronic Sum Frequency Generation. J. Phys. Chem.
40
41 Lett. 2014, 5, 762–766.
42
43
44 (14) Tang, M.; Larish, W. A.; Fang, Y.; Gankanda, A. n.; Grassian, V. H. Heterogeneous
45
46 Reactions of Acetic Acid with Oxide Surfaces: Effects of Mineralogy and Relative Hu-
47
48 midity. The Journal of Physical Chemistry A 2016, 120, 5609–5616, PMID: 27322707.
49
50
(15) Fang, Y.; Tang, M.; Grassian, V. H. Competition between Displacement and Dissocia-
51
52
tion of a Strong Acid Com pared to a Weak Acid Adsorbed on Silica Particle Surfaces:
53
54
The Role of Adsorbed Water. The Journal of Physical Chemistry A 2016, 120, 4016–
55
56
4024, PMID: 27220375.
57
58
59 12
60 ACS Paragon Plus Environment
Page 13 of 14 The Journal of Physical Chemistry Letters

1
2
3
(16) Murdachaew, G.; Gaigeot, M.-P.; Halonen, L. r.; Gerber, R. B. First and second de-
4
5
protonation of H2SO4 on wet hydroxylated (0 001) alpha-quartz. Physical Chemistry
6
7
Chemical Physics 2014, 16, 22287–22298.
8
9
10 (17) Murdachaew, G.; Gaigeot, M.-P.; Halonen, L.; Gẽrber, R. B. Dissociation of HCl into
11
12 Ions on Wet Hydroxylated (0001) Îś-Quartz. The Journal of Physical Chemistry Letters
13
14 2013, 4, 3500–3507.
15
16
17 (18) Pérez de Tudela, R.; Marx, D. Acid Dissociation in HCl-Water Clusters is Temperature
18
19 Dependent and Cannot be Detected Based on Dipole Moments. Phys. Rev. Lett. 2017,
20
21 119, 223001.
22
23
24 (19) Reed Harris, A. E.; Pajunoja, A.; Cazaunau, M.; Gr̃atien, A.; Pangui, E.; Monod, A.;
25
26
Griffith, A., Elizabeth C. an d Virtanen; Doussin, J.-F.; Vaida, V. Multiphase Pho-
27
28
tochemistry of Pyruvic Acid under Atmospheric Conditions˙The Journal of Physical
29
30
Chemistry A 2017, 121, 3327–3339, PMID: 28388049.
31
32 (20) Cwiertny, D. M.; Young, M. A.; Grassian, V. H. Chemistry and Photochemistry of
33
34 Mineral Dust Aerosol. Annual Review of Physical Chemistry 2008, 59, 27–51, PMID:
35
36 18393675.
37
38
39 (21) Sulpizi, M.; Gaigeot, M.-P.; Sprik, M. The silica–water interface: how the silanols
40
41 determine the surface acidity and modulate the water properties. Journal of chemical
42
43 theory and computation 2012, 8, 1037–1047.
44
45
46 (22) Leung, K.; Nielsen, I. M. B.; Criscenti, L. J. Elucidating the Bimodal Acid-Base Be-
47
48 havior of the Water-Silica Interface from First Principles. Journal of the American
49
50 Chemical Society 2009, 131, 18358–18365, PMID: 19947602.
51
52
(23) Pfeiffer-Laplaud, M.; Costa, D.; Tielens, F.; Gaigeot, M.-P.; Sulpizi, M. Bimodal Acidity
53
54
at the Amorphous Silica/Water Interface. The Journal of Physical Chemistry C 2015,
55
56
119, 27354–27362.
57
58
59 13
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Letters Page 14 of 14

1
2
3
(24) Sulpizi, M.; Sprik, M. Acidity constants from vertical energy gaps: density functional
4
5
theory based molecular dynamics implementation. Phys. Chem. Chem. Phys. 2008, 10,
6
7
5238–5249.
8
9
10 (25) Costanzo, F.; Sulpizi, M.; R.G., V.; Sprik, M. The oxidation of tyrosine and tryptophan
11
12 studied by a molecular dynamics normal hydrogen electrode. J. Chem. Phys. 2011, 134,
13
14 244508.
15
16
17 (26) Tazi, S.; Rotenberg, B.; Salanne, M.; Sprik, M.; Sulpizi, M. Absolute acidity of clay
18
19 edge sites from ab-initio simulations. Geochimica et Cosmochimica Acta 2012, 94, 1 –
20
21 11.
22
23
24 (27) Pfeiffer-Laplaud, M.; Gaigeot, M.-P.; Sulpizi, M. pKa at Quartz/Electrolyte Interfaces.
25
26 The Journal of Physical Chemistry Letters 2016, 7, 3229–3234, PMID: 27483195.
27
28
29 (28) Becke, A. D. Density-functional exchange-energy approximation with correct asymp-
30
31 totic behavior. Phys. Rev. A 1988, 38, 3098–3100.
32
33
(29) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy
34
35
formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789.
36
37
38 (30) VandeVondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.; Hutter, J.
39
40 Quickstep: Fast and accurate density functional calculations using a mixed Gaussian
41
42 and plane waves approach. Computer Physics Communications 2005, 167, 103 – 128.
43
44
45 (31) Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials.
46
47 Phys. Rev. B 1996, 54, 1703–1710.
48
49
50 (32) Hartwigsen, C.; Goedecker, S.; Hutter, J. Relativistic separable dual-space Gaussian
51
52 pseudopotentials from H to Rn. Phys. Rev. B 1998, 58, 3641–3662.
53
54
55
56
57
58
59 14
60 ACS Paragon Plus Environment

You might also like