Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)
Article
Interaction Energies in Complexes of Zn and Amino-Acids: A Comparison of Ab-Initio and Force Field based Calculations Ran Friedman, Emma Ahlstrand, and Kersti Hermansson J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.6b12969 • Publication Date (Web): 08 Mar 2017 Downloaded from http://pubs.acs.org on March 14, 2017
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
The Journal of Physical Chemistry A 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 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
Interaction Energies in Complexes of Zn and Amino-Acids: a Comparison of Ab-Initio and Force Field based Calculations Emma Ahlstrand,†,‡ Kersti Hermansson,¶ and Ran Friedman∗,†,‡ 1
†Department of Chemistry and Biomedical Sciences, Linnæus University, 391 82 Kalmar, Sweden ‡Linnæus University Centre for Biomaterials Chemistry ¶Department of Chemistry, Ångström Laboratory, Uppsala University, BOX 538, 751 21 Uppsala, Sweden E-mail:
[email protected] Phone: +46 (0)480 446290
1 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
Abstract
2
3
Zinc plays important roles in structural stabilization of proteins, enzyme catalysis
4
and signal transduction. Many Zn binding sites are located at the interface between
5
the protein and the cellular fluid. In aqueous solutions, Zn ions adopt an octahedral
6
coordination, while in proteins zinc can have different coordinations with a tetrahedral
7
conformation found most frequently. The dynamics of Zn binding to proteins, and the
8
formation of complexes that involve Zn, are dictated by interactions between Zn and
9
its binding partners. We calculated the interaction energies between Zn and ligands in
10
complexes that mimic protein binding sites, and in Zn-complexes of water and one or
11
two amino acid moieties, using quantum-mechanics (QM) and molecular mechanics
12
(MM). It was found that MM calculations that neglect or only approximate polarizability
13
did not reproduce even the relative order of the interaction energies in these com-
14
plexes. Interaction energies calculated with the CHARMM-Drude polarizable force
15
field agreed better with the ab initio results, although the deviations between QM and
16
MM were still rather large (40–96 kcal/mol). In order to gain further insight into Zn-
17
ligand interactions, the free energies of interaction were estimated by QM calculations
18
with continuum solvent representation, and we performed energy decomposition anal-
19
ysis calculations to examine the characteristics of the different complexes. The ligands
20
were found to have high impact on the relative strength of polarization and electrostatic
21
interactions. Interestingly, ligand-ligand interactions did not play a significant role in
22
the binding of Zn. Finally, analysis of ligand exchange energies suggests that car-
23
boxylates could be exchanged with water molecules which explains the flexibility in Zn
24
binding dynamics. An exchange between carboxylate (Asp/Glu) and imidazole (His)
25
is less likely.
26
Introduction
27
Zn2+ is one of the most commonly encountered protein metal cofactors and the interplay
28
between Zn2+ and amino acid residues therefore plays an important role in biology. The 2 ACS Paragon Plus Environment
Page 2 of 39
Page 3 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
29
zinc ion participates in enzyme catalysis, structural stabilization of protein motifs, and bio-
30
logical regulation in enzymes. 1 Many of the catalytic Zn-binding sites are partially exposed
31
to solvent (intra- or extracellular fluid). The interactions between Zn and its ligands in the
32
protein ensure that the formation of Zn-protein complexes is thermodynamically and ki-
33
netically favored over its fully hydrated conformation. Moreover, the Zn-ligand interactions
34
tune the subtle dynamics and electrostatics of the protein. A sound description of the ion
35
in both these environments, i.e. protein and solution, is required for an understanding of
36
the ion’s binding and dynamics. In dilute aqueous solution, a Zn2+ ion is known to adopt
37
an octahedral coordination by binding to the oxygen atoms of six water molecules. 2,3 In
38
proteins, on the other hand, a tetrahedral conformation is most frequently found, 4 and the
39
most common amino acid ligands are cysteine, histidine, aspartate and glutamate. 5–8 The
40
ability of zinc to readily adopt different coordinations is an important feature of this ion.
41
Clearly, acquiring adequate information about the energetics of protein-ion interac-
42
tions is required for an understanding of the biology of Zn metalloproteins. 9 A significant
43
component of the free energy of Zn complexation involves polarization of the ligands
44
and charge transfer. 10,11 Proper estimates of the interaction energies between the ion
45
and the relevant ligands can be derived from quantum mechanical calculations (QM), but
46
high-level QM calculations for a whole protein (or even a protein domain) are still com-
47
putationally prohibited. Approximating some of the system by employing e.g., a quantum
48
mechanics/molecular mechanics (QM/MM) approach (for example, see 12,13 ) can signif-
49
icantly reduce the computational requirements. However, QM/MM calculations present
50
their own challenges, such as the treatment of the boundary between the QM and MM
51
parts, and the computational speed for the QM part, particularly in dynamic systems that
52
contain multivalent metal ions. Moreover, in many such cases the QM region needs to be
53
large and the time-span to be covered need to be relatively long (hundreds of nanosec-
54
onds or longer).
55
A force field (i.e., MM-) based approach is routinely used in molecular dynamics (MD)
3 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
56
simulations of biological systems. 14,15 Available atomistic, non-polarizable force fields are
57
often fast and efficient but not always accurate enough to represent the intricate character
58
of protein-ion interactions. 16 Some force fields make use of a covalent representation for
59
the zinc-ligand interactions. However, although bonded models can be very useful in the
60
study of certain metalloproteins, 7,17 they are not suited for studies where the metal coor-
61
dination undergoes changes, necessitating the development of non-bonding methods.
62
Several attempts to develop non-bonded force fields for Zn were reported in the lit-
63
erature, 18–21 but modeling different binding conformations and ligand exchange correctly
64
remains a challenge. To exemplify this challenge for catalytic sites in contact with solu-
65
tion, we performed a 5 ns MD simulation of the S100A12 protein with the well established
66
CHARMM27 force field and the zinc ion parameters developed by Stote and Karplus. 18
67
Already after 2 ns of simulation, Zn becomes hydrated, and it adopts an octahedral coor-
68
dination that is not supported by experimental evidence (Figure 1).
69
Polarizable force fields are more complicated and (at least in principle) can be more
70
accurate than classical, non-polarizable ones. Drude-model based force fields that can
71
handle Zn ions have been developed. 22,23 A different approach to representing the polar-
72
izability of atoms is the use of a multipole expansion, as in the AMOEBA (Atomic Multipole
73
Optimized Energetics for Biomolecular Applications) force field. 24 It is important to note,
74
however, that polarizable force field may fail to represent systems in which a significant
75
charge transfer occurs. The so-called Sum of Interactions Between fragments Ab ini-
76
tio computed (SIBFA) 25 approach can deal with charge transfer but is not yet applicable
77
to dynamical metal coordination changes. Another charge transfer polarisable force-field
78
has been developed for simulations of Zn2+ and other ions in water, 26 but does not include
79
parameters for biomolecules, acetate or imidazole.
80
One reason why non-bonded models in force fields tend to fail to correctly represent
81
Zn2+ -protein systems is that the Lennard-Jones (LJ) parameters are usually calibrated
82
for ions in water and not for ions bound to protein-residues. Indeed, accounting for both
4 ACS Paragon Plus Environment
Page 4 of 39
Page 5 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
(a)
(b)
Figure 1: Simulation of S100A12 (PDBID:2WCB 27 ) with CHARMM27 and zinc ion parameters developed by Stote and Karplus; 18 (a) zinc (silver color) in the binding site with three His and one Asp and (b) after 2 ns simulation the zinc ion binding site is clearly more hydrated and the Zn complex adopts an octahedral conformation that is not supported by experiments. 83
ion-water and ion-residue interactions in the parametrisation of the force field was shown
84
to yield better accuracy in simulations of proteins with other ions (Ca2+ , K+ and Na+ ). 28,29
85
Likewise, Sakharov and Lim 11 developed an empirical scheme to handle polarizability and
86
charge transfer for Cys2 His2 Zn binding sites, but this model is not made to be used with
87
other binding conformations.
88
In our previous work, 30 we highlighted the limitations of some widely used QM-methods
89
(in particular DFT functionals) in describing simple, biologically-relevant Zn and Cd-complexes
90
and found that it is desirable to use the second-order Møller-Plesset theory (MP2) method
91
if more accurate alternative such as Coupled Cluster with Single, Double, and Triple ex-
92
citations (CCSD(T)) are not affordable. Here, we examined the variation of interaction
93
energies of structure-optimized Zn-ligand complexes, using MP2, in vacuum and in two
94
solvents described by continuum representations (water and a low dielectric solvent). We
95
compared the interaction energy differences for ten Zn-complexes calculated with QM
96
to those calculated with MM with the aim to elucidate the origin of the differences. The
5 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
97
amino acid combinations in the complexes represent zinc binding sites from the protein
98
data bank. In addition, we studied complexes where 1-2 ligands were replaced by wa-
99
ter molecules in order to represent partially solvated systems such as expected during
100
catalysis, when the protein binds or releases the Zn2+ ion, or during ligand exchange.
101
The models that were used in this study represent Zn-binding sites of different char-
102
acter with four to six ligands (Figure 2). Complex S1 represents His4 Zn, which has been
103
suggested to be involved in pathogenic Zn2+ catalyzed amyloid aggregation. 31 Model S2
104
represents His3 Glu/Asp as liganding residues. This combination is common among pro-
105
tein structures (e.g., MMP3 32 and S100A12 27 ). We were interested to see the effects of
106
replacing the acetate by a more basic group (methoxide) that represents serine. This is
107
modeled by complex S3. Serine is not a common ligand for Zn in proteins, 6,8 but in some
108
cases it can replace the native ligand without apparent loss in affinity. 33 Complex S4
109
represents His3 Wa (Wa for water). This coordination is found in several enzymes includ-
110
ing carbonic anhydrase 34 and MMP13. 35 Complexes S5–S9 represent intermediates that
111
could be found after ligand exchange, during protein folding and unfolding, or during pep-
112
tide aggregation. Finally, complex S10 is the fully hydrated ion. Cysteine ligands were
113
investigated previously 9,17,36–42 and were therefore not included in this study.
114
Given that many applications and models (e.g. force fields, solvation models) begin
115
with and rely on an accurate description of the gas-phase interaction energies, or at the
116
very least make use of them as a reference, we first performed calculations for nine Zn2+
117
complexes with model amino acid like ligands in vacuum. Thereafter we included also
118
solvent interactions to cover a more complete and realistic description of the interface
119
between solution and protein. 8 In this study we used tetrahydrofuran (THF) as a repre-
120
sentation of the protein environment outside the Zn2+ -ligand complexes, because of the
121
similarity between the dielectric properties of THF and those of proteins. 8,43
6 ACS Paragon Plus Environment
Page 6 of 39
Page 7 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
122
Methods
123
Model systems
124
The Zn-complexes included in this work are shown in Figure 2. The model systems were
125
numbered S1-10, with no particular order. We replaced the amino acid side chains of
126
histidine, serine and glutamic acid (or aspartic acid) by imidazole, methoxide and acetate,
127
respectively. 44,45 Of note, serine was modeled as methoxide and not as methanol because
128
QM calculations have shown that binding of methoxide was favored over water, but that
129
of methanol is not 46 and since metal coordination can reduce the pKa of serine enough
130
to deprotonate the oxygen 47 . Water was also a possible ligand. The carboxylate residue
131
of acetate can interact with the metal ion in a bi-dentate or mono-dentate manner. 48 S10
132
is a reference complex with only water ligands, [Zn(H2 O)6 ]2+
133
Quantum chemistry calculations
134
The GAMESS 49 (General Atomic and Molecular Electronic Structure System) software
135
was used for all QM calculations. Each of the model systems (S1-10) was geometry
136
optimized with the default Quadratic Approximation algorithm with the M06 50 functional
137
and the def2-TZVP 51 basis set. For these optimized structures, interaction energies were
138
then calculated at the MP2 52 level with the aug-cc-pVTZ 53 basis set. Solvent effects were
139
approximated by the iterative Conductor Polarizable Continuum Model (C-PCM). 54 The
140
solvents considered were water (dielectric coefficient r =78.39) and tetrahydrofuran (THF,
141
r =7.58). The atomic radius for Zn was then set to 1.39 Å. 8 For the PCM calculations the
142
aug-cc-pVDZ 53,55 basis set was used.
143
The Counterpoise (CP) correction method 56 was applied for Basis Set Superposi-
144
tion Error (BSSE) correction of interaction energies. The QM interaction energies were
145
further analyzed with the Localized Molecular Orbital Energy Decomposition Analysis
7 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
Page 8 of 39
S1 Imi4 (+2)
S2 Imi3 Ace (+1)
S3 Imi3 Meo (+1)
S4 Imi3 Wa (+2)
S5 Imi2 Ace (+1)
S6 Imi2 Wa2 (+2)
S7 AceWa2 (+1)
S8 AceWa4 (+1)
S9 AceWa5 (+1)
S10 Wa6 (+2)
Figure 2: Model systems, S1-10. Imi=imidazole (model complex for histidine), Ace=acetate (model complex for aspartic acid or glutamic acid) and Meo=methoxide ion (model complex for serine). Distances ≥ 2.1 Å between the interacting atoms are shown with dashed lines and distances < 2.1 Å are shown as bonds with solid lines. The net charge of the complex is shown in parenthesis.
8 ACS Paragon Plus Environment
Page 9 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
146
(LMOEDA) 57,58 scheme, which divides the interaction energy into contributions from elec-
147
trostatics, exchange, repulsion, polarization (which includes charge transfer) and disper-
148
sion. This Energy Decomposition Analysis (EDA) was carried out in order to shed light
149
on which interactions that play major (or minor) roles in the binding of the metal ion to
150
ligands. The EDAPCM 58 method was used to perform an EDA in solvated systems, using
151
a modified version of GAMESS courtesy of Peifeng Su.
152
Quantum mechanics interaction energies
153
The interaction energy in vacuo between the Zn and all the ligands (L) was calculated for
154
each system (S1-10) according to:
∆E int = Ecomplex − (EZn2+ +
X
EL )
(1)
allL
155
where Ecomplex is the total energy of the complex at its optimized structure. Individual lig-
156
and energies (BSSE-corrected), were calculated at the equilibrium geometry (i.e. lowest
157
energy structure).
158
To explore the influence of ligand-ligand interactions, calculations were also made for
159
the interaction energies of a "ligands-only cluster", i.e. a complex from which the ion has
160
been removed but the ligands were fixed at the positions of the optimized ion-ligands
161
complex, according to:
∆E int,shell = Ecomplex − (EZn2+ + ELC )
(2)
162
Here Ecomplex is the total energy of the complex at its optimized structure, and ELC is the
163
total energy of the "ligands-only cluster" kept at the same geometry as in the optimized Zn-
164
ligand complex. The difference between ∆E int,shell and ∆E int accounts for the influence
165
of ligand-ligand interactions on the formation of the complex.
166
To evaluate the energy cost or gain accompanying the exchange of one or more lig9 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
Page 10 of 39
167
ands, the difference between ∆E int,shell for selected complexes were calculated. Such
168
interaction energy differences are given the notation ∆∆E int,shell (SX → SY ) and were
169
calculated as the differences between two selected complexes, SX and SY.
∆∆E int,shell (SX → SY ) = ∆E int,shell (SY ) − ∆E int,shell (SX)
(3)
170
Solvent effects were approximated with the Zn-ligand systems embedded in a polar-
171
izable continuum model (PCM) 54 solvent. Using the PCM model, equation (2) is trans-
172
formed into a free-energy solvent-embedded variant, namely:
CM P CM ∆Gint,shell,P CM = GPcomplex − (GPZnCM ) 2+ + GLC
(4)
173
CM GPcomplex is the total free energy of the complex (using the structure optimized in vacuum)
174
and GPLCCM is the total free energy of the "ligands-only cluster".
175
Optimization of the structures was performed in GAMESS-US with the M06 DFT func-
176
tional. DFT calculations are significantly faster than those performed with MP2, whereas
177
our earlier study has shown that differences with respect to MP2-optimized structures
178
were very small. 30 In the same study, interaction energies calculated with various function-
179
als and with MP2 were compared to CCSD(T) calculations as reference, which revealed
180
good agreement between MP2 and CCSD(T) results (mean absolute error 0.9 kcal/mol),
181
but less convincing agreement between DFT and CCSD(T).
182
Atomic charges
183
We calculated atomic charges by use of three different schemes: Mulliken population
184
analysis, 59 molecular electrostatic potential charge fitting, also known as the Kollman-
185
Singh or Merz-Kollman-Singh (MKS) scheme, 60 and the CHarges from ELectrostatic Po-
186
tentials using a Grid based method (CHELPG). 61 Mulliken’s method is widely used. How-
10 ACS Paragon Plus Environment
Page 11 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
187
ever, its drawbacks include strong dependence on the basis set and its reliance on equal
188
division of overlapping occupied orbitals even between atoms with differences in elec-
189
tronegativities. Electrostatic Potential (ESP) based methods such as MKS and CHELPG
190
do not suffer from these limitations. The main difference between the MKS method and
191
CHELPG is in the procedure used to fit the ESP.
192
Molecular mechanics interaction energy calculations
193
194
Non-polarizable force fields
195
We also performed force-field, i.e. molecular mechanics (MM), calculations of ∆E int,shell
196
for systems S1-S10. Here, the system preparation, energy minimization and interac-
197
tion energy calculations were performed with the GROMACS 62 (GROningen MAchine for
198
Chemical Simulations) software. For the ligands, the SPC/E 63 (extended simple point
199
charge model) water model was used within the OPLS-AA 64,65 (Optimized Potentials for
200
Liquid Simulations-All Atom) and Amber99SB 66 (Assisted Model Building with Energy
201
Refinement) force fields and the TIP3P (three-site Transferable Intermolecular Potential)
202
water model within the CHARMM27 67 force field (see Table 1). For the Zn2+ ion, three
203
different LJ-parameter sets were tested. 11,18,21 We used the combination of force fields
204
and parameters as originally developed, i.e., Li, Merz and co-workers 21 with Amber99SB,
205
Stote and Karplus 18 and Sakharov and Lim 11 with CHARMM27, and we also explored the
206
Zn-parameters from Stote and Karplus with all three force fields. All three LJ-parameter
207
sets were originally optimized for the divalent ion-water systems. Force field parameters
208
for acetate, imidazole and methoxide are presented in Tables S1–S3 in the Supporting
209
information.
210
Coordinates from the QM optimized structures (M06/def2-TZVP) were used as input
11 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
211
structures for the force field based calculations. The systems were then energy minimized
212
with the conjugate gradient algorithm in vacuum. A plain cutoff at 2.0 nm was used for the
213
Coulomb and van der Waals forces. Table 1: Molecular mechanics set ups used in this study, with non-polarizable force fields
A
LJ-parameter set Reference σZn (nm) Zn (kJ/mol) 18 Stote and Karplus 0.194216 1.0460
B C
Li et al. 21 Sakharov and Lim 11
0.227400 0.156800
0.0148 0.7657
Force field
Water model for ligands OPLS-AA SPC/E Amber99SB SPC/E CHARMM27 TIP3P Amber99SB SPC/E CHARMM27 TIP3P
12 ACS Paragon Plus Environment
Page 12 of 39
Page 13 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
214
215
Polarizable force-field
216
Interaction energies were calculated by applying the CHARMM Drude force field. 22,68 Wa-
217
ter molecules were represented by the polarizable SWM4-NDP model, 69 which is com-
218
patible with the same force field. System preparation was carried out in CHARMM, 70,71
219
based on scripts adapted from CHARMM-GUI. 72,73 Energy minimization and calculations
220
were performed in NAMD. 74,75
221
Results and discussion
222
Interaction energies of Zn complexes - comparison between ab initio
223
and molecular mechanics calculations
224
Complexes S1-S10 include four-, five- and six-coordinated systems (acetate can bind
225
Zn2+ in either a monodentate or bidentate fashion). Optimization of the structures with
226
ab-initio methods resulted in tetrahedral or octahedral coordinations with small distortions
227
(Figure 2). Interestingly, coordination numbers of 5 and 7 were observed for complexes
228
S2 and S9, respectively, with both the Amber99SB and CHARMM27 FFs as both oxygens
229
of the acetate coordinated the Zn2+ ion in the minimum. The conformations of all other
230
systems optimized using the MM methods were very similar to those obtained by DFT.
231
Bond lengths are presented in the Supporting Information Table S4.
232
The interaction energies between the Zn and its ligands (systems S1-S10) as cal-
233
culated with MP2/aug-cc-pVTZ for the DFT-optimized structures are presented in Ta-
234
ble 2. ∆Eint,shell values are presented with CP correction. However, the corrections
235
were 5 kcal/mol or smaller in all systems (Table S5), indicating that the basis set is large
236
enough for these calculations in view of the fact that the interaction energies themselves 13 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
237
are large. The least favorable interaction energy (-360 kcal/mol), was calculated for the
238
fully hydrated Zn ion, whereas the most favorable one (-620 kcal/mol) was calculated for
239
complex S3 (Imi3 Meo). Furthermore, a comparison of ∆Eint,shell and ∆E int in Table 2,
240
revealed that the ligand-ligand interactions in the complexes contributed 20 kcal/mol or
241
less to the interaction energies, and were thus fairly insignificant in comparison with the
242
overall magnitude of the interactions. To estimate the influence of the geometric distortion
243
of the ligands in the complex, i.e., deviations from the optimized isolated ligand structures,
244
we recalculated ∆Eint for systems S2, S6, and S10 in the same geometry as in the com-
245
plex rather than at the optimal isolated ligand geometries. The interaction energies were
246
rather similar irrespective of whether optimization of the isolated ligands was performed
247
in all three cases (Table S5).
248
Page 14 of 39
The interaction energies for the ten Zn-ligand systems, calculated with the force fields
249
OPLS-AA, Amber99SB and CHARMM27 in combination with the three different LJ-parameters
250
for Zn, are presented in Table 2 and Figure 3a. The differences between the MM and MP2
251
interaction energies were overall large (65 kcal/mol in average and up to 170 kcal/mol),
252
see Figure 3b. Examination of the three force fields used here reveals that although the
253
actual ∆E int,shell values differed substantially from those calculated with MP2, the three
254
force fields displayed quite similar trends. All overestimated the interaction energies of
255
complexes S8 to S10, and underestimated the rest, where the interaction energies of S1
256
and S4 were the most heavily underestimated. The similarity between the force fields can
257
be attributed to the similarity in partial charges (Table S1-S3 in Supporting Information).
258
Calculations with the Stote and Karplus parameters for Zn in combination with the
259
CHARMM27 force field performed significantly better than the Amber99SB and OPLS-AA
260
force fields for all model systems except S7 (Figure 1). Stote and Karplus developed their
261
parameters to be compatible with the CHARMM series of force fields. As can be seen
262
from Table 2 and Figure 3a, the Zn parameters from Sakharov and Lim, 11 also calculated
263
with the CHARMM27 force field, resulted in better agreement with MP2 than the param-
14 ACS Paragon Plus Environment
Page 15 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
264
eters from Stote and Karplus for complexes S1 and S3-S7, but not for the others. Bond
265
length values are given in the Supporting Information, Table S6. We also note that the
266
optimized Zn parameters from Li et al. (in combination with the Amber99SB force field)
267
led to a better agreement with the MP2 calculation for complexes S1-S7 compared with
268
the LJ-parameters developed by Stote and Karplus, whereas the opposite was true for
269
complexes S8-S10.
270
The calculated interaction energies for complexes S1 to S9, with respect to the hy-
271
drated ion (S10) as a reference, are presented in Table 3. At the MP2 level in vacuum,
272
complexes S1-S9 yielded lower, i.e. more favorable, interaction energies than the hy-
273
drated ion. This was not the case for the interaction energies calculated with the different
274
force fields, where model systems S1, S4 and S6 were less stable than the all-water
275
complex (S10). This may be due to larger contribution from polarization (see section 3.4
276
below).
277
Interaction energies in continuum representations of solvents
278
The QM interaction energies were estimated also in a continuum representation of water
279
and THF (the latter may roughly represent interactions in proteins 8 ). Energy values with a
280
PCM model, formally correspond to free energies of interactions and should therefore not
281
be directly compared with ∆E int,shell . As expected, differences between ∆Gint,shell values
282
in THF and water were seen to be most apparent for systems that include charged ligands
283
(acetate and methoxide), due to the difference in dielectric constant (Figure S2, Support-
284
ing information). For those complexes, the ∆Gint,shell values were about 30 kcal/mol more
285
negative (i.e. more stable) in THF than in water. The differences for the neutral systems,
286
containing only imidazole or water as ligands were much smaller. For the fully hydrated
287
ion, embedding in THF increased the free energy of interaction by as little as 1 kcal/mol,
288
indicating that in PCM is the ion in this sense regarded as fully solvated by a single shell
289
of water molecules. Interestingly, whereas ∆E int,shell values were the least favorable for 15 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
290
the hydrated ion (S10, Table 3), this was not the case for ∆Gint,shell . In PCM water, the
291
interaction free energies calculated for complexes S5-S7 were less favorable than S10,
292
which suggests that such conformations would not be stable in solution. One may argue
293
that the low dielectric constant of THF does not well represent the systems that are par-
294
tially hydrated (S6 and S7), and that a higher dielectric constant ( ≥20 or more) would
295
be better. 43 However, we find that these systems yield interaction free energies that are
296
even less favorable in PCM water relative to system S10.
297
Ligand exchange energies
298
To get a measure of the interaction energies for the individual ligands in the complex, the
299
difference between two selected systems, ∆∆E int (SX → SY ) or ∆∆E int,shell (SX → SY )
300
was calculated from the MP2 values. Such ligand exchange interaction energies for (i)
301
exchange from water to imidazole, (ii) from water to acetate, and (iii) from imidazole to
302
acetate are presented in Table 4. An exchange of a water molecule by imidazole, i.e.
303
S6 → S4, is accompanied by an energy gain of about 30–40 kcal/mol, and an exchange
304
of two water molecules by two imidazoles (S7 → S5) by almost twice as much. Ligand
305
exchange of two water molecules by one acetate (S10 → S8) is very favorable in vacuum
306
(-200 kcal/mol), while exchange from two imidazoles to one acetate (S6 → S7) yields
307
-140 kcal/mol.
308
The difference in the interaction energies with regard to ligand exchange in PCM water
309
and THF, ∆∆Gint,P CM (SX → SY ), were similar in magnitude for water to imidazole.
310
However, the exchange from water to acetate yielded only 7–30 kcal/mol in PCM, which
311
was much less than the corresponding value in vacuum. It is apparent that the effects of
312
ligand exchange are less pronounced when a water molecule is replaced by a charged
313
acetate in solvent, especially in water, where the favorable interaction between Zn and the
314
charged ligand is offset by the loss of solvation free energy for the complex who carries out
315
a smaller charge (+1 versus +2 when a water molecule is replaced by an acetate). The 16 ACS Paragon Plus Environment
Page 16 of 39
Page 17 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
316
∆∆Gint,shellP CM (SX → SY ) values for a replacement of two imidazoles to one acetate
317
were +40–+70 kcal/mol in PCM, suggesting that such an exchange is not likely to occur.
318
To assess the performance of the three force fields the same ligand exchange ener-
319
gies were calculated by MM. OPLS-AA and Amber99SB yielded no difference in inter-
320
action energies upon exchange from water to imidazole. In contrast, calculations with
321
the CHARMM27 force field resulted in interaction energies which were more favorable
322
by about 20 kcal/mol per molecule. These values are in better agreement with the MP2
323
results (30–40 kcal/mol). It is apparent that the Zn+2 -imidazole (His) interaction is not
324
favorable enough in MM. The distances between the Zn atom and its nitrogen ligands on
325
imidazole are accurate with the OPLS-AA MM, whilst ∼0.10 Å too long with Amber99SB
326
and CHARMM27 (Table S4). Ligand exchange from two water to one acetate is very
327
favorable in vacuum (-200 kcal/mol) also when calculated by MM.
328
For a ligand exchange from two imidazole to one acetate the two force fields (OPLS-
329
AA and Amber99SB) predicted too favorable ∆∆E int,shell (SX → SY ) compared to MP2,
330
whereas the interaction energies calculated with CHARMM27 were closer to MP2.
331
Energy decomposition and atomic charge analyses
332
Energy decomposition analysis 57,58 is used here in order to estimate the relative con-
333
tributions of interactions that govern the binding of Zn in the different complexes. The
334
decomposed interaction energies, are presented in Figure 4 and Table S7.
335
The contribution of polarization to the interaction energies is larger in the systems that
336
involve imidazole (S1-S6) compared to those that do not (S7-S10). The contribution from
337
polarization appears even more significant for S1, S4 and S6 where all ligands are neutral
338
and the total interaction energies are smaller. The electrostatic contribution is roughly as
339
large as all the other stabilizing interactions together for the neutral systems, whereas it
340
is more dominant in all the others. The contribution of dispersion to the total interaction
341
is somewhat more significant for all the imidazole-containing complexes. This appears to 17 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
342
play a role in biology since dispersion interactions are relevant for nature’s fine-tuning of
343
biochemical reactions, even if the electrostatic forces are much stronger. 76
344
The interaction energies for S8 and S9 in vacuum are both about -550 kcal/mol, and
345
EDA revealed only minor differences between the systems. The addition of a water ligand
346
is apparently offset by the acetate coordinating in a mono-dentate instead of a bidentate
347
fashion.
348
Turning now to the solvent-embedded (in THF) systems, we find that the electrostatic
349
contributions were larger compared to the calculations in vacuum, whereas the polariza-
350
tion interaction energies became a little bit less significant (Figure 4b). The difference in
351
interaction free energies between the different systems levels off somewhat when they
352
are embedded in a continuum due to the considerably large desolvation energies (Fig-
353
ure S1 in Supporting Information). We note that, the desolvation energy considered by
354
LMOEDA 58 refers only to the electrostatic interaction between solute and solvent in a
355
cavity defined by the complex. Complexes with a negatively charged ligand (S2, S3, S5,
356
S7, S8 and S9) were characterized by desolvation energies of 170–250 kcal/mol. The
357
desolvation energies of complexes with only neutral ligands (S4, S1, S6 and S10) were
358
much lower, between 10 and 80 kcal/mol.
359
The complex with the least favorable interaction energy in PCM is the tetrahedral com-
360
plex S7 with one acetate and two water ligands (Table S8). The high energy cost due
361
to the desolvation for this system resulted in a higher ∆Gint,shell value compared to S8
362
and S9, with their large numbers of water ligands (Figure 4b). It had previously been
363
suggested that a tetrahedral coordination of ligands was preferred for Zn2+ complexes
364
containing one negative ligand. 77 Here we found that hexacoordinated system (S9) had
365
more favorable interaction energy than the tetracoordinated (S7) by 40 kcal/mol in vacuum
366
and by 100 kcal/mol in PCM.
367
Examination of the atomic charge on the Zn ion can yield a qualitative indication of the
368
amount of the electronic charge transfer. In the following, we refer to charges calculated
18 ACS Paragon Plus Environment
Page 18 of 39
Page 19 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
369
by the CHELPG method, which are shown in Table 5 (MKS and Mulliken charges are
370
shown in Table S8). The water molecules apparently polarized the Zn ion, which resulted
371
in a +2.2 au charge for Zn in model system S10, consistent with similar calculations
372
(personal communication with authors of 78 ). In the systems with a single acetate and
373
water ligands (S7-S9) the Zn charge was just over +1 au, whereas in the complexes that
374
contained imidazoles (S1-S6) the charge of the Zn ion ranged between +0.5 and +1 au.
375
The results reveal that charge transfer is evident in all systems. This may indicate that
376
even polarizable force fields are liable to inaccuracies. Inclusions of charge transfer in
377
the calculation of force-field interaction energies 25 may yield better agreement with high-
378
level ab initio results. Several methods have been suggested to overcome this problem
379
by charge fitting. 11,17,79 However, this may not be an adequate solution for systems that
380
contain several His ligands, because (1) the charge transfer was very substantial, (2)
381
there was an evident disagreement even between the two ESP-based methods, and (3)
382
any movement of the ion or ligands is liable to modify the amount of charge transfer.
383
Simulations of the dynamics of such complexes thus remain a challenge.
19 ACS Paragon Plus Environment
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
Zn systems Name Ligands S1 Imi4 S2 Imi3 Ace S3 Imi3 Meo S4 Imi3 Wa S5 Imi2 Ace S6 Imi2 Wa2 S7 AceWa2 S8 AceWa4 S9 AceWa5 S10 Wa6
∆Eint -428 -586 -600 -399 -568 -366 -496 -537 -553 -342
∆Eint,shell -438 -606 -619 -409 -571 -376 -513 -555 -554 -357
QM vacuum
QM PCM ∆Gint,shell Water THF -371 -379 -368 -397 -388 -416 -330 -339 -297 -330 -290 -300 -223 -258 -332 -358 -349 -373 -339 -340
MM vacuum (∆Eint,shell ) OPLS-AA Amber99SB CHARMM27 CHARMM-Drude 18 18 21 18 11 A A B A C Polarizable FF -286 -271 -325 -347 -404 -509 -559 -550 -630 -588 -663 -702 -533 -513 -588 -570 -642 N/A -287 -271 -328 -324 -380 -471 -493 -485 -558 -510 -576 -647 -287 -271 -330 -302 -356 -425 -493 -485 -561 -468 -530 -567 -613 -601 -686 -580 -654 -631 -619 -630 -676 -613 -652 -600 -414 -392 -461 -374 -435 -397
Table 2: Interaction energies, ∆Eint and ∆Eint,shell (kcal/mol), calculated with QM (MP2/aug-cc-pTZV) in vacuum and PCM (water and THF), with five combinations of FF and LJ parameters for Zn2+ , see Table 1, and the CHARMM-Drude polarizable force field 22,68 in vacuum for the Zn model systems (S1-S10). N/A - not available.
The Journal of Physical Chemistry
ACS Paragon Plus Environment
20
Page 20 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
Zn systems Name Ligands S1 Imi4 S2 Imi3 Ace S3 Imi3 Meo S4 Imi3 Wa S5 Imi2 Ace S6 Imi2 Wa2 S7 AceWa2 S8 AceWa4 S9 AceWa5 S10 Wa6
∆Eint -86 -244 -258 -57 -226 -24 -154 -195 -211 0
∆Eint,shell -80 -249 -262 -52 -214 -19 -156 -198 -197 0
QM vacuum
QM PCM ∆Gint,shell Water THF -32 -39 -29 -57 -49 -76 9 1 42 10 49 40 116 82 7 -18 -10 -33 0 0
MM vacuum (∆Eint,shell ) OPLS-AA Amber99SB CHARMM27 CHARMM-Drude 18 18 21 18 11 A A B A C Polarizable FF 128 121 135 28 31 -112 -145 -158 -169 -214 -228 -305 -119 -121 -127 -195 -207 N/A 127 121 133 50 55 -74 -79 -93 -97 -135 -141 -250 127 121 131 73 79 -28 -79 -93 -100 -93 -95 -170 -199 -209 -225 -206 -218 -234 -205 -238 -215 -239 -218 -203 0 0 0 0 0 0
int,shell int,shell Table 3: Interaction energies for complexes (S1-S9) with respect to hydrated Zn2+ (S10), i.e. ∆ESX − ∆ES10 , 2+ calculated with MP2 in vacuum and in PCM, and with five combinations of FF and LJ parameters for Zn (Table 1) in vacuum. All values are in kcal/mol. N/A - not available.
Page 21 of 39 The Journal of Physical Chemistry
ACS Paragon Plus Environment
21
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
(a)
Page 22 of 39
(b)
Figure 3: Comparison of the MM and QM (MP2) interaction energies (∆Eint,shell ) for complexes S1-S10. The MM energies were calculated with the OPLS-AA, CHARMM27 and Amber99SB for the ligands combined with Zn parameters from Stote and Karplus, 18 Sakaharov and Lim, 11 and Li et al., 21 and with the CHARMM-Drude polarizable force field 22,68 . (a) Correlation plot. (b) Comparison of MM interaction energies relative to QM (MP2) where the reference ("0") is the MP2 energy. All values are in kcal/mol.
384
385
Calculations of Zn-ligand interaction energies with a polarizable force-
386
field
387
Some of the limitations of classical force fields when it comes to interactions between
388
metal ions and proteins may be overcome by the use of polarizable force-field. For this
389
reason, we calculated the interaction energies of complexes S1, S2 and S4–S10 with the
390
CHARMM-Drude polarizable force field. The CHARMM-Drude force field was chosen as
391
it is available in several different MD packages and because it included parameters for
392
all the ligands included in this study except methoxide. The results (Table 2), revealed a
393
marked improvement over all of the non-polarizable force-fields tested here. The MM to
394
QM difference is more consistent compared to the non-polarizable force fields (Figure 3b).
395
Moreover, there is an agreement on the ordering of the interaction energies between the 22 ACS Paragon Plus Environment
Page 23 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
Table 4: Ligand exchange energies calculated as the difference in interaction energies between two complexes SX and SY, ∆∆E int,shell (SX → SY ), (Eq. 3) denoted diff in the table, calculated with MP2, with OPLS-AA and CHARMM27 force fields with Stote and Karplus parameters for Zn 18 and with a polarizable force field 22,68 in vacuum. The interaction energies between the ion and individual ligands with ligands at their lowest energy structure, ∆∆E int (SX → SY ), (see Eq. 1) for the MP2-method in vacuum were calculated for several complexes and are given in parenthesis. The difference between ∆∆E int (SX → SY ) and ∆∆E int,shell (SX → SY ) is the contribution of ligand-ligand interactions within the complex. MP2 ∆∆Gint,shell (SX → SY ) values are calculated with PCM (water and THF). All energies are in kcal/mol. Wa to Imi
Wa to Imi
S6 Imi2 Wa2 ∆E int,shell -376 (-366) -287 -302 -425
S4 Imi3 Wa ∆E int,shell -409 (-399) -287 -324 -471
diff -33 (-33) 0 -22 -46
MP2 vacuum OPLS-AA CHARMM27 CHARMM-Drude
S7 AceWa2 ∆E int,shell -513 (-496) -493 -468 -567
S5 Imi2 Ace ∆E int,shell -571 (-568) -493 -510 -647
diff -58 (-72) 0 -42 -80
MP2 vacuum OPLS-AA CHARMM27 CHARMM-Drude
S6 Imi2 Wa2 ∆E int,shell -376 (-366) -287 -302 -425
S1 Imi4 ∆E int,shell -438(-428) -286 -347 -509
diff -62 (-62) +1 -45 -84
MP2 vacuum OPLS-AA CHARMM27 CHARMM-Drude
S10 Wa6 ∆E int,shell -357 (-342) -414 -374 -397
S8 AceWa4 ∆E int,shell -555 (-537) -613 -580 -631
diff -198 (-195) -199 -206 -234
MP2 vacuum OPLS-AA CHARMM27 CHARMM-Drude
S6 Imi2 Wa2 ∆E int,shell -376 (-366) -287 -302 -425
S5 Imi2 Ace ∆E int,shell -571 (-568) -493 -510 -647
diff -195 (-202) -206 -208 -222
S6 Imi2 Wa2 ∆E int,shell -376 (-366) -287 -302 -425
S7 AceWa2 ∆E int,shell -513 (-496) -493 -468 -567
diff -137 (-130) -206 -166 -142
MP2 vacuum OPLS-AA CHARMM27 CHARMM-Drude 2 Wa to 2 Imi
2 Wa to Ace
2 Imi to Ace MP2 vacuum OPLS-AA CHARMM27 CHARMM-Drude
S6 Imi2 Wa2 ∆Gint,shell -290 -300
S4 Imi3 Wa ∆Gint,shell -330 -339
diff -40 -39
MP2 PCM water MP2 PCM THF
S7 AceWa2 ∆Gint,shell -223 -258
S5 Imi2 Ace ∆Gint,shell -297 -330
diff -74 -72
MP2 PCM water MP2 PCM THF
S6 Imi2 Wa2 ∆Gint,shell -290 -300
S1 Imi4 ∆Gint,shell -371 -379
diff -81 -79
MP2 PCM water MP2 PCM THF
S10 Wa6 ∆Gint,shell -339 -340
S8 AceWa4 ∆Gint,shell -332 -358
diff -7 -18
MP2 PCM water MP2 PCM THF
S6 Imi2 Wa2 ∆Gint,shell -290 -300
S5 Imi2 Ace ∆Gint,shell -297 -330
diff -7 -30
S6 Imi2 Wa2 ∆Gint,shell -290 -300
S7 AceWa2 ∆Gint,shell -223 -258
diff +67 +42
MP2 PCM water MP2 PCM THF
2 Wa to 2 Imi
2 Wa to Ace
2 Imi to Ace MP2 PCM water MP2 PCM THF
23 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
S1
S2
211 -219
-84
-72
S5
Electrostatic Exchange
-194 -22
-69
S8
-453
-79
S9
Desolvation
-22
145 -59
-164
-15
a
Dispersion Repulsion
-288
144 -59
-149
Polarization
-186
-478
172
197
S6
-23 -316
S7
-20 -526
203 -81
-208 -92
-218
-18 -481
235
251 -99
-208
-26 -317
S4
S3
181
Page 24 of 39
-168
129
-54
-154
-9
-11 -466
S10
-462
24 ACS Paragon Plus Environment
-8 -270
Page 25 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
252
185 213
204
53
S1
S2
-74
-195 -85
197
S3
-25
-179
-20
-21
-530
-339
-99
-161
-565
234 239
S4
S5
-184
-94
S6
-93
-151
-23
-314
145
251
S8
-59
177
S9
-59
-143
-13 -484
146
194 -119
b
-18
-21
180
-72
-160
-81
-538
-341
S7
203 71
237
65
-9 -484
S10
129 11
-54 -147
-152
-5
-8 -476
-274
Figure 4: Interaction energies for Zn model systems S1-S10 divided into their energy components according to the LMOEDA (in frame a) and EDAPCM (in frame b) schemes with the area of the pie relative to the total interaction energies, a) in vacuum and, b) in PCM THF. Negative values represent a stabilizing interaction, positive values a destabilizing interaction.
25 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
Table 5: Charges calculated with the CHELPG method at the MP2/aug-cc-pVTZ level for the optimized Zn systems S1-S10. For MKS and Mulliken charges, see Table S8. Zn systems Name Ligands Charge S1 Imi4 +2 S2 Imi3 Ace +1 S3 Imi3 Meo +1 S4 Imi3 Wa +2 S5 Imi2 Ace +1 S6 Imi2 Wa2 +2 S7 AceWa2 +1 S8 AceWa4 +1 S9 AceWa5 +1 S10 Wa6 +2
Zn 0.54 0.77 0.72 0.75 0.98 0.99 1.18 1.20 1.08 2.22
-0.30 -0.33 -0.31 -0.19 -0.32 -0.39 -0.71 -0.74 -0.76 -1.15
CHELPG charges Ligands -0.22 -0.13 -0.06 -0.30 -0.08 -0.76 -0.76 -0.74 -0.21 -0.20 -0.68 -0.24 -0.27 -0.76 -0.35 -0.71 -0.75 -0.27 -0.74 -0.86 -0.69 -0.72 -0.74 -0.75 -0.73 -0.74 -0.75 -0.77 -0.72 -0.78 -0.71 -0.68 -0.69 -0.75 -1.13 -1.11 -1.13 -1.11 -1.15
Color codes: imidazole N, acetate O, methoxide O, water O. 396
MM and QM calculations, which suggests that simulations involving ligand exchange may
397
be possible. Examination of the different complexes reveals that the calculated energies
398
are most similar for complexes S10, S9, S6 and S7 that involve a high degree of hydration
399
(at least two water molecules). Complex S8 is an outlier in this respect - it involves four
400
water molecules and its energy is too favorable by 76 kcal/mol relative to MP2. The rea-
401
son for this is that in the optimized structure the aspartate in complex S8 was bidentate,
402
whereas in S9 it was monodentate and hydrogen-bonded to the water. The bidentate in-
403
teraction was stronger (due to electrostatics) and hence the interaction energy was more
404
favorable for complex S8. Examination of the EDA for complexes S8 and S9 (Figure 4A)
405
reveals that despite the different coordination with respect to the aspartate the contribu-
406
tions of electrostatics, exchange, repulsion, polarization and dispersion are almost the
407
same between those systems. The same is apparently not true in MM, where the two
408
oxygens have the same fixed charge.
26 ACS Paragon Plus Environment
Page 26 of 39
Page 27 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
409
Conclusions
410
In this study of small systems that mimic Zn binding sites in proteins three force fields
411
(OPLS-AA, Amber99SB and CHARMM27) were compared with calculations of the inter-
412
action energies using the MP2 method and a large basis set (which was previously shown
413
to yield good agreement with high level CCSD(T) calculations). In vacuum, all systems
414
with protein ligands (S1-S9) were more stable than the hydrated ion (S10), i.e., they were
415
characterized by more negative interaction energies, using MP2. Complexes with neutral
416
ligands only (imidazole and water) were characterized by less favorable total interaction
417
energies and high amount of polarization according to energy decomposition analysis
418
(LMOEDA), compared to complexes with an acetate ligand. When the systems were em-
419
bedded in PCM the differences in interaction energies between the systems almost dis-
420
appeared due to a large desolvation term for the charged acetate ligand. The Zn-acetate
421
interaction was estimated to be about 200 kcal/mol in vacuum (based on calculations of
422
ligand exchange), whereas it was only 10-30 kcal/mol in water or THF because of the
423
large desolvation term. No such differences were observed for the interaction between
424
Zn and imidazole. Half of the attractive interaction energies in the imidazole containing
425
systems could be attributed to electrostatics and more than one fourth of the interaction
426
to polarization. The interaction free energies for neutral complexes containing imidazole
427
were quite similar regardless of the solvent.
428
The imidazole-containing systems (S1-S6) were characterized by less favorable in-
429
teraction energies in non polarizable force field based (MM) calculations than those cal-
430
culated with MP2, while the three systems without imidazoles result in more favorable
431
interaction energies for MM than MP2. The MM energies are off by at least 20 kcal/mol
432
and up to 150 kcal/mol. Interestingly, the free energy of binding the ion in complexes S1,
433
S4, and S6, that included only neutral ligands, were lower than that of the hydrated ion.
434
Classical, non-polarizable force fields represent the interactions between Zn2+ and its lig-
27 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
435
ands as a combination of electrostatics and LJ interactions (when using the more flexible
436
non-bonded model). Systems S1, S4, S6 and S10 are those where the electrostatic con-
437
tribution to the binding interaction was lower (Figure 4a). In complexes S1, S4 and S6,
438
the contribution from polarization, exchange and repulsion was rather large. Polarization
439
is only implicitly accounted for by classical force fields (e.g., by tailoring partial charges).
440
Exchange is neglected by force fields, whereas repulsion is represented to some extent
441
by a repulsive term in the LJ interactions, which applies at short interatomic distances.
442
Electrostatics is by far the most dominant contribution in non-polarizable MM force fields,
443
and when it is lower the interactions are less favorable than those calculated by Ab initio
444
methods. Force-field parameters for ions were historically developed for use in aqueous
445
solutions, which is likely to be the cause for the over-stability of highly hydrated complexes
446
(S8, S9 and S10) in the non-polarizable force fields. Unfortunately, a simple shift of the
447
potential so that ion-water interactions are not overly stable is not likely to be the key for
448
simulating Zn2+ dynamically together with peptides while accounting for ligand exchange,
449
and the contribution from exchange and repulsion makes it difficult for polarization alone
450
to fix this issue. It was shown here that the order of the interaction energies is similar
451
when calculated with a polarisable force field and MP2, but that the differences in their
452
magnitude that depend on the system. Ion parameters in the CHARMM-Drude force field
453
were developed by simulating the ions in SWM4-NDP water. 68 One may expect that the
454
generation of a polarizable force-field, e.g. of the Drude type, where special attention
455
during the parametrization procedure is paid to the residues in focus in this study, as well
456
as to methanethiol (for cysteine), would lead to a good agreement between the MM and
457
QM inteaction energies, and thus yield fairly realistic simulations.
458
Supporting Information Available
459
The following files are available free of charge.
Supplementary Information: Comple-
28 ACS Paragon Plus Environment
Page 28 of 39
Page 29 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
460
mentary data including force field input, bond lengths for optimized systems, interaction
461
energies with different basis sets, LMOEDA values and atomic charges with the MKS and
462
Mulliken methods.
463
Acknowledgement
464
The calculations were performed on resources provided by the Swedish National Infras-
465
tructure for Computing (SNIC) at Lunarc, centre for scientific and technical computing for
466
research at Lund University and at PDC center for high performance computing at KTH
467
Royal Institute of Technology (project numbers SNIC 2016/1-55, SNIC 2016/1-222, and
468
SNIC 2016/1-518). This work was supported by the Linnæus Centre of Excellence “Bio-
469
materials Chemistry” (R.F), and by the eSSENCE national strategic research program in
470
e-science (K.H).
471
References
472
473
(1) Maret, W.; Li, Y. Coordination dynamics of zinc in proteins. Chem. Rev. 2009, 109, 4682–4707.
474
(2) Burgess, J. Metal ions in solution; Ellis Horwood: Chichester, 1978.
475
(3) Ohtaki, H.; Radnai, T. Structure and dynamics of hydrated ions. Chemical Reviews
476
477
478
479
480
1993, 93, 1157–1204. ˘ (4) Roe, R. R.; Pang, Y.-P. ZincâAšs Exclusive Tetrahedral Coordination Governed by Its Electronic Structure. Molec. Model. Ann. 1999, 5, 134–140. (5) Harding, M. M. Geometry of metal–ligand interactions in proteins. Acta Cryst. D 2001, 57, 401–411.
29 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
481
482
(6) Sousa, S. F.; Lopes, A. B.; Fernandes, P. A.; Ramos, M. J. The zinc proteome: A tale of stability and functionality. Dalton Trans. 2009, 38, 7946–7956.
483
(7) Peters, M. B.; Yang, Y.; Wang, B.; Füsti-Molnár, L.; Weaver, M. N.; Merz, K. M. Struc-
484
tural Survey of Zinc Containing Proteins and the Development of the Zinc AMBER
485
Force Field (ZAFF). J. Chem. Theory Comput. 2010, 6, 2935–2947.
486
487
(8) Friedman, R. Structural and computational insights into the versatility of cadmium binding to proteins. Dalton Trans. 2014, 43, 2878–2887.
488
(9) Zhang, J.; Yang, W.; Piquemal, J.-P.; Ren, P. Modeling Structural Coordination and
489
Ligand Binding in Zinc Proteins with a Polarizable Potential. J. Chem. Theory Com-
490
put. 2012, 8, 1314–1324.
491
492
493
494
(10) Hoops, S. C.; Anderson, K. W.; Merz, K. M. Force field design for metalloproteins. J. Am. Chem. Soc. 1991, 113, 8262–8270. (11) Sakharov, D. V.; Lim, C. Zn Protein Simulations Including Charge Transfer and Local Polarization Effects. J. Am. Chem. Soc. 2005, 127, 4921–4929.
495
(12) Corminboeuf, C.; Hu, P.; Tuckerman, M. E.; Zhang, Y. Unexpected Deacetylation
496
Mechanism Suggested by a Density Functional Theory QM/MM Study of Histone-
497
Deacetylase-Like Protein. J Am. Chem. Soc. 2006, 128, 4530–4531.
498
(13) Wu, R.; Hu, P.; Wang, S.; Cao, Z.; Zhang, Y. Flexibility of Catalytic Zinc Coordina-
499
tion in Thermolysin and HDAC8: A Born-Oppenheimer ab Initio QM/MM Molecular
500
Dynamics Study. J. Chem. Theory Comput. 2010, 6, 337–343.
501
502
503
(14) Friedman, R. Aggregation of amyloids in a cellular context: modelling and experiment. Biochem. J. 2011, 438, 415–426. (15) Friedman, R.; Boye, K.; Flatmark, K. Molecular modelling and simulations in cancer
30 ACS Paragon Plus Environment
Page 30 of 39
Page 31 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
504
research. Biochimica et Biophysica Acta (BBA)-Reviews on Cancer 2013, 1836, 1–
505
14.
506
(16) Becconi, O.; Ahlstrand, S.; Salis, A.; Friedman, R. Protein-ion interactions: simula-
507
tions of bovine serum albumin in physiological solutions of NaCl, KCl and LiCl. Isr J
508
Chem 2017, In Press, 0.
509
510
511
512
513
514
(17) Lin, F.; Wang, R. Systematic Derivation of AMBER Force Field Parameters Applicable to Zinc-Containing Systems. J. Chem. Theory Comput. 2010, 6, 1852–1870. (18) Stote, R. H.; Karplus, M. Zinc binding in proteins and solution: a simple but accurate nonbonded representation. Proteins 1995, 23, 12–31. (19) Babu, C. S.; Lim, C. Empirical Force Fields for Biologically Active Divalent Metal Cations in Water. J. Phys. Chem. A 2006, 110, 691–699.
515
(20) Wu, R.; Lu, Z.; Cao, Z.; Zhang, Y. A Transferable Non-bonded Pairwise Force Field
516
to Model Zinc Interactions in Metalloproteins. J. Chem. Theory Comput. 2011, 7,
517
433–443.
518
(21) Li, P.; Roberts, B. P.; Chakravorty, D. K.; Merz, K. M. Rational Design of Particle
519
Mesh Ewald Compatible Lennard-Jones Parameters for +2 Metal Cations in Explicit
520
Solvent. J. Chem. Theory Comput. 2013, 9, 2733–2748.
521
(22) Lopes, P. E.; Huang, J.; Shim, J.; Luo, Y.; Li, H.; Roux, B.; Mackerell, A. D. Force
522
Field for Peptides and Proteins based on the Classical Drude Oscillator. J Chem
523
Theory Comput 2013, 5430–5449.
524
525
(23) Li, P.; Kenneth M. Merz, J. Taking into Account the Ion-Induced Dipole Interaction in the Nonbonded Model of Ions. J. Chem. Theory Comput. 2014, 10, 289–297.
526
(24) Ren, P.; Ponder, J. W. Consistent treatment of inter- and intramolecular polarization
527
in molecular mechanics calculations. J. Comput. Chem. 2002, 23, 1497–1506. 31 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
528
(25) Gresh, N.; El Hage, K.; Perahia, D.; Piquemal, J.-P.; Berthomieu, C.; Berthomieu, D.
529
Polarizable molecular mechanics studies of Cu(I)/Zn(II) superoxide dismutase:
530
Bimetallic binding site and structured waters. J. Comput. Chem. 2014, 35, 2096–
531
2106.
532
533
(26) Soniat, M.; Hartman, L.; Rick, S. W. Charge Transfer Models of Zinc and Magnesium in Water. J Chem Theory Comput 2015, 11, 1658–1667.
534
(27) Moroz, O. V.; Blagova, E. V.; Wilkinson, A. J.; Wilson, K. S.; Bronstein, I. B. The
535
Crystal Structures of Human S100A12 in Apo Form and in Complex with Zinc: New
536
Insights into S100A12 Oligomerisation. J. Mol. Biol. 2009, 391, 536 – 551.
537
538
539
540
(28) Project, E.; Nachliel, E.; Gutman, M. Parameterization of Ca+2 -protein interactions for molecular dynamics simulations. J. Comput. Chem. 2008, 29, 1163–1169. (29) Hess, B.; van der Vegt, N. F. Cation specific binding with protein surface charges. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 13296–13300.
541
(30) Ahlstrand, E.; Spångberg, D.; Hermansson, K.; Friedman, R. Interaction energies
542
between metal ions (Zn2+ and Cd2+ ) and biologically relevant ligands. Int. J. Quant.
543
Chem. 2013, 113, 2554 – 2562.
544
(31) Minicozzi, V.; Stellato, F.; Comai, M.; Serra, M. D.; Potrich, C.; Meyer-Klaucke, W.;
545
Morante, S. Identifying the Minimal Copper- and Zinc-binding Site Sequence in
546
Amyloid-β Peptides. J. Biol. Chem. 2008, 283, 10784–10792.
547
548
549
550
(32) Pelmenschikov, V.; Siegbahn, P. E. M. Catalytic Mechanism of Matrix Metalloproteinases: Two-Layered ONIOM Study. Inorg. Chem. 2002, 41, 5659–5666. (33) Stradal, T. B.; Troxler, H.; Heizmann, C. W.; Gimona, M. Mapping the zinc ligands of S100A2 by site-directed mutagenesis. J. Biol. Chem. 2000, 275, 13219–13227.
32 ACS Paragon Plus Environment
Page 32 of 39
Page 33 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
551
(34) Eriksson, A. E.; Jones, T. A.; Liljas, A. Refined structure of human carbonic an-
552
hydrase II at 2.0 Å resolution. Proteins: Struct., Funct., and Bioinform. 1988, 4,
553
274–282.
554
(35) Engel, C. K.; Pirard, B.; Schimanski, S.; Kirsch, R.; Habermann, J.; Klingler, O.;
555
Schlotte, V.; Weithmann, K. U.; Wendt, K. U. Structural basis for the highly selective
556
inhibition of MMP-13. Chem. Biol. 2005, 12, 181–189.
557
(36) Ryde, U. The coordination chemistry of the structural zinc ion in alcohol dehydroge-
558
nase studied by ab initio quantum chemical calculations. Eur. Biophys. J. 1996, 24,
559
213–221.
560
(37) Tiraboschi, G.; Gresh, N.; Giessner-Prettre, C.; Pedersen, L. G.; Deerfield, D. W.
561
Parallel ab initio and molecular mechanics investigation of polycoordinated Zn(II)
562
complexes with model hard and soft ligands: Variations of binding energy and of
563
its components with number and charges of ligands. J. Comput. Chem. 2000, 21,
564
1011–1039.
565
566
(38) Dudev, T.; Lim, C. Factors Governing the Protonation State of Cysteines in Pro˘ ’ An Ab Initio/CDM Study. J. Am. Chem. Soc. 2002, 124, 6759–6766. teins:âAL
567
(39) De Courcy, B.; Gresh, N.; Piquemal, J.-P. Importance of lone pair interac-
568
tions/redistribution in hard and soft ligands within the active site of alcohol dehydro-
569
genase Zn-metalloenzyme: Insights from electron localization function. Interdiscip.
570
Sci.: Comput. Life Sci. 2009, 1, 55–60.
571
(40) Sakharov, D. V.; Lim, C. Force fields including charge transfer and local polarization
572
effects: Application to proteins containing multi/heavy metal ions. J. Comput. Chem.
573
2009, 30, 191–202.
574
(41) Zhu, X.; Bora, R. P.; Barman, A.; Singh, R.; Prabhakar, R. Dimerization of the Full-
33 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
575
Length Alzheimer Amyloid β-Peptide (Aβ42) in Explicit Aqueous Solution: A Molec-
576
ular Dynamics Study. J. Phys. Chem. B 2012, 116, 4405–4416.
577
578
579
580
(42) Zhu, T.; Xiao, X.; Ji, C.; Zhang, J. Z. H. A New Quantum Calibrated Force Field for Zinc-Protein Complex. J. Chem. Theory Comput. 2013, 9, 1788–1798. (43) Friedman, R.; Nachliel, E.; Gutman, M. Molecular Dynamics of a Protein Surface: Ion-Residues Interactions. Biophys. J. 2005, 89, 768–781.
581
(44) Cini, R.; Musaev, D. G.; Marzilli, L. G.; Morokuma, K. Molecular orbital study of
582
complexes of zinc(II) with imidazole and water molecules. J. Mol. Struc.: Theochem
583
1997, 392, 55 – 64.
584
(45) Friedman, R.; Nachliel, E.; Gutman, M. Application of Classical Molecular Dynamics
585
for Evaluation of Proton Transfer Mechanism on a Protein. Bioch. Bioph. Acta. 2005,
586
1710, 67–77.
587
(46) Trzaskowski, B.; Adamowicz, L.; Deymier, P. A. A theoretical study of zinc(II) interac-
588
tions with amino acid models and peptide fragments. J. Biol. Inorg. Chem. 2008, 13,
589
133–137.
590
(47) Berthon, G. Critical evaluation of the stability constants of metal complexes of amino
591
acids with polar side chains (Technical Report). Pure & Appl. Chem. 1995, 67, 1117–
592
1240.
593
594
(48) Ryde, U. Carboxylate Binding Modes in Zinc Proteins: A Theoretical Study. Biophys. J. 1999, 77, 2777 – 2787.
595
(49) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S.; Gordon, M.; Jensen, J. H.;
596
Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, S. et al. General Atomic and Molecular
597
Electronic Structure System. J. Comput. Chem. 1993, 14, 1347–63.
34 ACS Paragon Plus Environment
Page 34 of 39
Page 35 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
598
599
The Journal of Physical Chemistry
(50) Zhao, Y.; Truhlar, D. G. Density functionals with broad applicability in chemistry. Acc. Chem. Res. 2008, 41, 157–167.
600
(51) Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and
601
quadruple zeta valence quality for H to Rn: Design and assessment of accuracy.
602
Phys. Chem. Chem. Phys. 2005, 7, 3297–3305.
603
604
(52) Møller, C.; Plesset, M. Note on an approximation treatment for many-electron systems. Phys. Rev. 1934, 46, 0618–0622.
605
(53) Balabanov, N. B.; Peterson, K. A. Systematically convergent basis sets for transition
606
metals. I. All-electron correlation consistent basis sets for the 3d elements Sc–Zn. J.
607
Chem. Phys. 2005, 123, 064107.
608
(54) Li, H.; Jensen, J. H. Improving the efficiency and convergence of geometry opti-
609
mization with the polarizable continuum model: New energy gradients and molecular
610
surface tessellation. J. Comput. Chem. 2004, 25, 1449–1462.
611
612
(55) Feller, D. The role of databases in support of computational chemistry calculations. J. Comput. Chem. 1996, 17, 1571–1586.
613
(56) Boys, S.; Bernardi, F. The calculation of small molecular interactions by the differ-
614
ences of separate total energies. Some procedures with reduced errors. Mol. Phys.
615
1970, 19, 553–566.
616
617
618
619
620
621
(57) Peifeng, S.; Hui, L. Energy decomposition analysis of covalent bonds and intermolecular interactions. J. Chem. Phys. 2009, 131, 014102. (58) Su, P.; Liu, H.; Wu, W. Free energy decomposition analysis of bonding and nonbonding interactions in solution. J. Chem. Phys. 2012, 137, 034111. (59) Mulliken, R. S. Electronic Population Analysis on LCAO-MO Molecular Wave Functions. I. J. Chem. Phys. 1955, 23, 1833–1840. 35 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
622
623
(60) Singh, U. C.; Kollman, P. A. An approach to computing electrostatic charges for molecules. J. Comput. Chem. 1984, 5, 129–145.
624
(61) Breneman, C. M.; Wiberg, K. B. Determining atom-centered monopoles from molec-
625
ular electrostatic potentials. The need for high sampling density in formamide con-
626
formational analysis. J. Comput. Chem. 1990, 11, 361–373.
627
628
629
630
(62) Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A. E.; Berendsen, H. J. C. GROMACS: Fast, flexible, and free. J. Comput. Chem. 2005, 26, 1701–1718. (63) Berendsen, H. J. C.; Grigera, J. R.; Straatsma, T. P. The missing term in effective pair potentials. J. Phys. Chem. 1987, 91, 6269–6271.
631
(64) Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. Development and Testing of the
632
OPLS All-Atom Force Field on Conformational Energetics and Properties of Organic
633
Liquids. J. Am. Chem. Soc. 1996, 118, 11225–11236.
634
(65) Kaminski, G. A.; Friesner, R. A.; Tirado-Rives, J.; Jorgensen, W. L. Evaluation and
635
Reparametrization of the OPLS-AA Force Field for Proteins via Comparison with
636
Accurate Quantum Chemical Calculations on Peptides. J. Phys. Chem. B 2001, 105,
637
6474–6487.
638
(66) Hornak, V.; Abel, R.; Okur, A.; Strockbine, B.; Roitberg, A.; Simmerling, C. Compar-
639
ison of multiple Amber force fields and development of improved protein backbone
640
parameters. Proteins: Structure, Function, and Bioinformatics 2006, 65, 712–725.
641
642
(67) MacKerell, A. D.; Banavali, N.; Foloppe, N. Development and current status of the CHARMM force field for nucleic acids. Biopolymers 2000, 56, 257–265.
643
(68) Yu, H.; Whitfield, T. W.; Harder, E.; Lamoureux, G.; Vorobyov, I.; Anisimov, V. M.;
644
Mackerell, A. D.; Roux, B. Simulating monovalent and divalent ions in aqueous solu-
645
tion using a Drude polarizable force field. J Chem Theory Comput 2010, 6, 774–786. 36 ACS Paragon Plus Environment
Page 36 of 39
Page 37 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
The Journal of Physical Chemistry
646
(69) Lamoureux, G.; Harder, E.; Vorobyov, I. V.; Roux, B.; MacKerell Jr, A. D. A polariz-
647
able model of water for molecular dynamics simulations of biomolecules. Chemical
648
Physics Letters 2006, 418, 245 – 249.
649
(70) Brooks, B. R.; Bruccoleri, R. E.; Olafson, B. D.; States, D. J.; Swaminathan, S.;
650
Karplus, M. CHARMM: A Program for Macromolecular Energy, Minimization, and
651
Dynamics Calculations. J. Comp. Chem. 1983, 4, 187–217.
652
(71) Brooks, B. R.; Brooks, C. L.; MacKerell, A. D.; Nilsson, L.; Roux, B.; Won, Y.; Ar-
653
chontis, G.; Bartels, C.; Boresch, S.; Caflisch, A. et al. CHARMM: The Biomolecular
654
Simulation Program. J. Comp. Chem. 2009, 30, 1545–614.
655
656
(72) Jo, S.; Kim, T.; Iyer, V. G.; Im, W. CHARMM-GUI: a web-based graphical user interface for CHARMM. J Comput Chem 2008, 29, 1859–1865.
657
(73) Lee, J.; Cheng, X.; Swails, J. M.; Yeom, M. S.; Eastman, P. K.; Lemkul, J. A.;
658
Wei, S.; Buckner, J.; Jeong, J. C.; Qi, Y. et al. CHARMM-GUI Input Generator for
659
NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using
660
the CHARMM36 Additive Force Field. J Chem Theory Comput 2016, 12, 405–413.
661
(74) Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.;
662
Skeel, R. D.; Kalé, L.; Schulten, K. Scalable molecular dynamics with NAMD. J Com-
663
put Chem 2005, 26, 1781–1802.
664
(75) Jiang, W.; Hardy, D. J.; Phillips, J. C.; Mackerell, A. D.; Schulten, K.; Roux, B. High-
665
performance scalable molecular dynamics simulations of a polarizable force field
666
based on classical Drude oscillators in NAMD. J Phys Chem Lett 2011, 2, 87–92.
667
(76) Tsfadia, Y.; Friedman, R.; Kadmon, J.; Selzer, A.; Nachliel, E.; Gutman, M. Molecular
668
dynamics simulations of palmitate entry into the hydrophobic pocket of the fatty acid
669
binding protein. FEBS Lett. 2007, 581, 1243–1247.
37 ACS Paragon Plus Environment
The Journal of Physical Chemistry
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
670
(77) Dudev, T.; Lim, C. Tetrahedral vs Octahedral Zinc Complexes with Ligands of Biolog-
671
ical Interest: A DFT/CDM Study. J. Am. Chem. Soc. 2000, 122, 11146–11153.
672
(78) Asthagiri, D.; Pratt, L. R.; Paulaitis, M. E.; Rempe, S. B. Hydration Structure and
673
Free Energy of Biomolecularly Specific Aqueous Dications, Including Zn2+ and First
674
Transition Row Metals. J. Am. Chem. Soc. 2004, 126, 1285–1289.
675
(79) Duarte, F.; Bauer, P.; Barrozo, A.; Amrein, B. A.; Purg, M.; Åqvist, J.; Kamerlin, S.
676
C. L. Force Field Independent Metal Parameters Using a Nonbonded Dummy Model.
677
J. Phys. Chem. B 2014, 118, 4351–4362.
38 ACS Paragon Plus Environment
Page 38 of 39
Page 39 of 39
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 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 60
678
The Journal of Physical Chemistry
Graphical TOC Entry
679
39 ACS Paragon Plus Environment