J Mol Model (2011) 17:3117–3128
DOI 10.1007/s00894-011-1000-0
ORIGINAL PAPER
Effect of metal Ions (Ni2+, Cu2+ and Zn2+) and water
coordination on the structure of L-phenylalanine, L-tyrosine,
L-tryptophan and their zwitterionic forms
Milan Remko & Daniel Fitz & Ria Broer &
Bernd Michael Rode
Received: 19 October 2010 / Accepted: 28 January 2011 / Published online: 2 March 2011
# The Author(s) 2011. This article is published with open access at Springerlink.com
Abstract Methods of quantum chemistry have been applied
to double-charged complexes involving the transition metals
Ni2+, Cu2+ and Zn2+ with the aromatic amino acids (AAA)
phenylalanine, tyrosine and tryptophan. The effect of
hydration on the relative stability and geometry of the
individual species studied has been evaluated within the
supermolecule approach. The interaction enthalpies, entropies and Gibbs energies of nine complexes Phe•M, Tyr•M,
Trp•M, (M=Ni2+, Cu2+ and Zn2+) were determined at the
Becke3LYP density functional level of theory. Of the
transition metals studied the bivalent copper cation forms
the strongest complexes with AAAs. For Ni2+and Cu2+ the
most stable species are the NO coordinated cations in the
AAA metal complexes, Zn2+cation prefers a binding to the
aromatic part of the AAA (complex II). Some complexes
energetically unfavored in the gas-phase are stabilized upon
microsolvation.
M. Remko (*)
Department of Pharmaceutical Chemistry,
Comenius University Bratislava,
Odbojarov 10,
SK-832 32 Bratislava, Slovakia
e-mail:
[email protected]
D. Fitz : B. M. Rode
Institute of General, Inorganic and Theoretical Chemistry,
University of Innsbruck,
Innrain 52a,
6020 Innsbruck, Austria
R. Broer
Department of Theoretical Chemistry,
Zernike Institute for Advanced Materials,
University of Groningen,
Nijenborgh 4,
9747 AG Groningen, The Netherlands
Keywords Aromatic amino acids . Complexes involving
the transition metals Ni2+, Cu2+ and Zn2+ . DFT
calculations . Molecular structure . Solvent effect
Introduction
Aromatic amino acids (AAA) are essential to higher
animals including humans. Phenylalanine and tryptophan
contain phenyl and indole rings and are classified as nonpolar. Due to its hydrophobicity L-phenylalanine is nearly
always found buried within a protein. The indole functional
group of L-tryptophan contains the N-H hydrogen bonding
group, which can interact with solvent in folded proteins.
L-tyrosine is a hydroxylated derivative of L-phenylalanine.
The phenolic hydroxyl group of this amino acid is
substantially more acidic than are the hydroxyls of aliphatic
amino acids (serine, threonine) and its acidity vary depending
on its environment. The body needs the essential amino acid
L-phenylalanine (and L-tyrosine) to make melanine, dopamine, noradrenaline, adrenaline and thyroxine. A shortage of
either of these two aromatic amino acids could cause mental
disorders, including anxienty, depression, low libido, and
chronic fatigue. Tryptophan is an essential amino acid that is
necessary for normal growth in infants and for nitrogen
balance in adults. All three aromatic amino acids can also be
administered as nutraceuticals.
Both experimental and theoretical studies demonstrate
that amino acids exist in the gas phase in the neutral form
[1]. However, in aqueous solution the zwitterions are the
predominant species [2–4]. The zwitterionic form of amino
acids (+H3N–CH(R)–COO–) can be stabilized by the
presence of counter ions and/or water molecules [1, 3]. In
order to investigate the nature of the interaction between
metal cations and aromatic amino acids, a number of
3118
theoretical and experimental studies have been conducted
[5–14]. The effect of water molecules on the stability of the
zwitterionic structure of phenylalanine, tyrosine and tryptophan was also investigated [3, 4]. Aromatic amino acids
may, besides interaction through their polar heads, also
interfere via their aromatic part of the molecule. The
aromatic groups of these amino acids can be involved in
non-covalent cation – π interactions with different cationic
species [11, 14–17]. This kind of interaction in biological
systems may play an important part, e.g., in biological
recognition processes [17], selectivity and functioning of
ion channels [18], or in drug-receptor interaction [19].
Despite their importance in biology and chemistry, the
interactions of aromatic amino acids with metallic cations
were prevailingly focused on the metallic complexes
containing alkali and alkaline earth metallic cations such
as Na+, K+, Mg2+ and Ca2+, which are essential for animal
and human life [1, 3, 11–14, 20]. On the other hand,
divalent cations of the transition metal elements, besides their
important role in functioning of enzymes, proteins and other
biomacromolecules, may play an important part for the
formation of peptides from simpler compounds under
conditions modeling those of the primordial earth [21, 22].
Divalent cations (e.g., Ni2+, Cu2+, Zn2+, etc.) can enhance
such a formation of peptides [22–26]. Few of the previous
studies considered metalation of aromatic amino acids by
transition metals. They are limited to the gas-phase complexes of phenylalanine with Ag+ and Zn2+ [8]. Rimola et al.
[9, 10] carried out theoretical calculations on Cu+ and Cu2+
gas-phase complexes of phenylalanine, tyrosine and tryptophan. Very recently, Larrucea et al. [7] carried out systematic
calculations of microsolvated complexes of aromatic amino
acids with Al3+. However, to the best of our knowledge, a
systematic study of the aromatic amino acid – transition
metal cation complexes has not been realized yet.
In this work we use several methods of theoretical
chemistry to investigate the conformations and energetics of
the complexes formed by transition metal cations Ni2+, Cu2+
and Zn2+ with phenylalanine, tyrosine and tryptophan. The
solvent effect on the systems studied was evaluated within a
supermolecule approach considering microsolvation of both
neutral and zwitterionic species. These metal cations have
numerous and important biochemical functions in living
systems. Hydrated complexes with aromatic amino acids
should serve as simpler models for more complex protein –
metal systems in an aqueous environment
Computational details
The geometries of the Phe•M, Tyr•M, Trp•M, Phe•M
(H2O)5, Tyr•M(H2O)6, and Trp•M(H2O)6 (M=Ni2+, Cu2+
and Zn2+) systems (Fig. 1) have been completely optimized
J Mol Model (2011) 17:3117–3128
Fig. 1 Structure and atom numbering of the aromatic amino acid – metal
ion complexes
with the Gaussian 03 program [27], using the density
functional theory methods [28–32] Becke3LYP/6–311+G
(d,p) [31, 32] and BHandHLYP/6–311+G(d,p) (note that
these are the half-and-half functionals implemented in
Gaussian 03 which are not the same as the ones proposed
by Becke [33]. The formation of metal – AAA complexes
can be described by reaction (A):
MðgÞ þ AAAðgÞ $ AAA MðgÞ
ð1Þ
M=Ni2+, Cu2+ and Zn2+. The gas-phase interaction
enthalpy (cation binding affinity), entropy and Gibbs energy
were calculated by the same way as in our previous
publication [34]. The equilibrium geometries and tautomeric
equilibria of AAA•M, and AAA•M(H2O)n, n=5, 6 (M=
Ni2+, Cu2+ and Zn2+) and corresponding zwitterionic forms
were determined at the Becke3LYP level of theory (for Cu2+
complexes also BHandHLYP level of theory) and the tripleξ basis set 6–311+G(d,p). For Ni, Cu and Zn, we used the
6–311+G(d,p) Wachters-Hay all electron basis set [35, 36].
Open shell calculations (Cu2+ complexes) have been carried
out using a spin-unrestricted formalism. Ni2+ may exist in its
-1.7
-38.7
-16.6
2.9
8.4
10.4
-18.4
0.4
-2.2
-36.6
-17.9
-0.2
7.4
15.8
-18.2
3.2
3.9
20.2
-17.2
1.2
8.2
8.9
13.3
2.3
-1.8
-38.0
16.1
4.8
6.3
1.8
-1.9
-0.5
Φ[C(1)–C(2)–O(3)···M(5)]
-0.9
-35.8
-17.8
Φ[C(2)–C(1)–N(6)···M(5)]
-20.3
79.8
28.4
10.0
-12.8
12.8
27.5
12.9
-16.3
10.9
-16.7
11.4
-11.6
-11.2
28.1
-15.2
-5.3
26.7
13.1
Φ[O(3)–C(2)–C(1)–N(6)]
15.2
113.3
114.6
81.8
82.4
115.3
113.0
81.8
81.9
113.5
113.6
84.3
83.8
113.6
115.2
80.0
81.8
115.1
112.7
82.1
83.6
115.2
113.2
84.6
79.7
113.8
110.8
87.7
83.5
115.7
81.8
84.8
Θ[O(3)···M(5)···N(6)]
81.6
113.8
113.7
Θ[C(2)–O(3)···M(5)]
113.4
107.3
106.0
109.5
110.9
109.4
105.6
108.4
108.9
109.9
110.1
109.4
110.3
108.8
106.0
109.7
111.5
105.7
108.2
Θ[C(1)–N(6)···M(5)]
109.3
119.5
122.8
108.5
108.7
121.5
123.3
109.3
107.0
119.2
122.3
108.3
109.5
121.8
122.9
108.6
108.1
121.2
123.5
109.4
108.6
121.2
123.1
108.7
107.3
119.5
123.3
110.0
108.1
120.9
107.1
108.1
Θ[C(2)–C(1)–N(6)]
109.0
119.3
122.0
Θ[C(1)–C(2)–O(3)]
122.9
2.087
2.093
2.117
2.027
2.011
2.066
2.103
2.090
2.013
2.050
2.044
1.982
2.011
2.064
2.157
2.067
2.017
2.065
2.097
2.087
1.972
2.043
2.048
1.990
2.080
2.104
2.036
1.963
1.961
2.019
2.008
1.965
d[O(3)···M(5)]
2.082
2.057
2.035
d[N(6)···M(5)]
2.106
1.495
1.283
1.295
1.491
1.495
1.289
1.313
1.487
1.496
1.282
1.306
1.495
1.495
1.283
1.299
1.486
1.491
1.289
1.313
1.488
1.501
1.282
1.309
1.494
1.494
1.282
1.287
1.512
1.501
1.279
1.312
1.485
1.281
1.496
1.300
1.500
d[C(2)–O(4)]
d[C(1)–N(6)]
1.247
1.529
1.535
1.239
1.243
1.521
1.546
1.219
1.251
1.525
1.538
1.227
1.249
1.520
1.535
1.233
1.241
1.521
1.544
1.220
1.251
1.516
1.534
1.224
1.247
1.530
1.524
1.253
1.253
n=0
n=5
n=0
Zn2+
n=5
1.516
1.547
1.219
1.526
1.250
1.536
1.233
d[C(1)–C(2)]
d[C(2)–O(3)]
n=6
n=0
n=0
n=0
n=0
n=5
n=0
n=6
Ni2+
Cu2+
Cu2+
n=6
Zn2+
n=6
n=0
n=6
n=0
n=6
Zn2+
Cu2+
Ni2+
(Tyr···M2+) x (H2O)n
Ni2+
According to the experimental evidence [46–48] the αamino acids and their derivatives prefer to form fivemembered chelate rings with metal cations. Previous
investigations [48–51] have shown that the most stable
structures for a metal cation and an aliphatic amino acid are
two, namely the one where the cation binds to both the N6
and the O3 sites of the neutral amino acid, and where it
binds to both oxygen ends of the zwitterionic amino acid.
However, aromatic amino acids can form three types of
complexes with metallic cations illustrated in Fig. 1.
Complex (I) represents charge-solvated systems. In those
complexes the M2+ ion is coordinated to nitrogen and
carbonyl oxygen (NO coordination) of AAAs. The aromatic
ring may take part in the cation – π interaction, complex
(II). The zwitterionic form of AAAs is involved in a
bifurcated interaction of the negatively charged carboxyl
group with the metal cation, complex (III), Fig. 1. Selected
structural parameters of the metallic complexes are given in
Tables 1, 2, and 3. The respective atoms are numbered
according to the scheme presented in Fig. 1. An analysis of
the harmonic vibrational frequencies of the optimized
species proved that all of them are minima (zero number
of imaginary frequencies). In the case of the isolated
Phe•Ni2+complexes I –III the calculations were performed
for the Ni2+cation present in both, high-spin and low-spin
(Trp···M2+) x (H2O)n
Molecular structures
(Phe···M2+) x (H2O)n
Results and discussion
Parameters
complexes both in high-spin and low-spin states (triplet
and singlet, respectively). Thus, for Ni2+ complexes the
initial calculations were carried out for both low-spin and
high-spin states. Cu2+ is an open shell system with a d9
(2D) ground state. Zn2+ is a d10 ion, so its complex is a
closed-shell system with a singlet ground state. Previous
investigations have shown [37–39] that the density
functional theory method yields results, which compare
favorably with the corresponding results obtained using
the high level ab initio coupled-cluster method. Hence,
DFT can sometimes be an economic alternative to ab initio
methods for studying larger systems. The values of metal
affinities computed using DFT are comparable to ab initio
SCF results and mostly in good agreement with the
corresponding experimental data [40–43]. The question of
the performance of different theoretical methods to correctly
describe the cation−pi interaction was discussed in several
publications [9, 10, 44, 45]. It was shown, that in the case of
aromatic amino acids−metal cation complexes, the hybrid
(B3LYP, BHLYP) DFT model chemistries perform well
compared to the highly correlated CCSD(T) and MP2
methods, respectively [9, 10, 45].
3119
Table 1 B3LYP/6-311+G(d,p) optimized relevant bond lengths (Å), bond angles (deg) and dihedral angles (deg) for the metal-coordinated and hydrated complexes of L-phenylalanine, L-tryptophan,
and L-tyrosine species
J Mol Model (2011) 17:3117–3128
108.6
115.1
-18.4
130.0
105.0
115.3
-15.1
146.8
108.8
114.3
-10.3
124.0
107.6
113.6
-12.6
155.4
109.7
112.2
-20.4
135.3
106.8
111.5
-15.3
152.4
108.7
121.5
0.1
105.9
104.7
115.0
-9.2
132.1
109.1
113.2
-12.9
127.3
Θ[C(2)–C(1)–N(6)]
Θ[C(1)–N(6)···M(5)]
Φ[O(3)–C(2)–C(1)–N(6)]
Φ[C(2)–C(1)–N(6)···M(5)]
107.1
113.4
-10.1
155.9
108.7
117.0
-18.2
137.7
105.3
114.4
-17.1
155.1
107.4
121.5
-12,6
110.1
107.0
117.2
-15.2
152.3
109.5
117.5
-4.5
113.7
109.3
116.6
-21.3
138.4
106.5
112.5
-15.9
154.6
107.5
119.9
-13.4
147.1
1.536
1.220
1.301
1.502
2.063
121.1
1.542
1.205
1.317
1.532
1.989
120.6
1.538
1.217
1.304
1.496
1.994
121.3
1.542
1.202
1.323
1.508
1.991
121.1
1.536
1.219
1.303
1.500
2.042
121.4
1.540
1.203
1.320
1.517
1.998
120.8
1.534
1.220
1.304
1.508
2.085
121.8
1.542
1.206
1.319
1.530
1.974
121.0
1.537
1.217
1.305
1.493
2.008
121.4
1.543
1.203
1.324
1.503
1.986
121.2
1.536
1.224
1.297
1.510
2.033
120.6
1.543
1.204
1.318
1.529
1.996
120.7
1.541
1.218
1.299
1.498
1.995
120.8
1.539
1.204
1.318
1.518
1.979
120.4
d[C(1)–C(2)]
d[C(2)–O(3)]
d[C(2)–O(4)]
d[C(1)–N(6)]
d[N(6)···M(5)]
Θ[C(1)–C(2)–O(3)]
n=6
n=0
n=5
n=0
n=5
Zn2+
1.535
1.219
1.305
1.504
2.058
122.1
1.539
1.220
1.300
1.498
2.054
120.9
1.541
1.203
1.319
1.519
1.990
120.7
1.546
1.201
1.326
1.496
2.003
121.6
n=0
n=0
n=0
n=0
n=5
n=0
n=0
Ni2+
Cu2+
Ni2+
Cu2+
(Trp_pi···M2+) x (H2O)n
(Phe_pi···M2+) x (H2O)n
n=6
Zn2+
n=6
n=0
n=6
Cu2+
Ni2+
(Tyr_pi···M2+) x (H2O)n
n=6
Zn2+
n=6
J Mol Model (2011) 17:3117–3128
Parameters
Table 2 B3LYP/6-311+G(d,p) optimized relevant bond lengths (Å), bond angles (deg) and dihedral angles (deg) for metal-coordinated and hydrated pi-complexes of the L-phenylalanine, L-tryptophan,
and L-tyrosine species
3120
states (triplet and singlet, respectively). Because the singlet
Ni2+ion and its phenylalanine complexes are higher in
energy than the corresponding triplet species, the results
obtained for the triplet Ni2+cation are discussed only. The
initial conformations of the AAA•M (M=Ni2+, Cu2+ and
Zn2+) complexes were constructed on the basis of previous
investigations of similar systems (phenylalanine - Zn2+,
Polfer et al. [8], Cu2+ gas-phase complexes of phenylalanine, tyrosine and tryptophan determined by Rimola et al.
[9]). In the neutral complex I the cation binds to both the
O3 and N6 sites of the neutral AAA. This complex has
been found by Rimola et al. [9] to be the most stable one.
The existence of π-complexes of AAAs with monovalent
metal cations was reported by several authors [1, 4, 9–14,
20]. Rimola et al. [9] examined the ability of interaction of
the transition metal dication Cu2+ with AAAs. However,
they were not able to determine a complex in which the Cu2
+
cation interacts with the π-system of the aromatic ring. In
order to check this finding, we decided to model this kind
of interaction using complex II, Fig. 1. In this complex,
besides π-interaction, the metal dication is stabilized via the
coordination through lone pair of the N6 nitrogen atom.
The conformation of this structure is further stabilized by
the hydrogen bond of the NH2•••O=C type (Fig. 1). The
salt-bridge complex III represents preferable coordination
of the zwitterionic AAA histidine [34].
The solvent effect on the geometry and stability of
individual complexes is studied making use of a supermolecule modeling of hydration. The solvation of the parent
systems was investigated using the Phe•M(H2O)5, Tyr•M
(H2O)6 and Trp•M(H2O)6 (M=Ni2+, Cu2+ and Zn2+) species.
The water molecules were placed at the polar parts of the
AAAs. Metal cations are coordinated, like in our previous
work [40], by two water molecules. The optimized complexes Phe•M(H2O)5, Tyr•M(H2O)6 and Trp•M(H2O)6, (M=
Ni2+, Cu2+ and Zn2+) converged to the energetically stable
species stabilized by both, strong electrostatic interaction and
intermolecular hydrogen bonds. For an illustration of the
molecular structure of these complexes the fully optimized
metal coordinated and solvated systems represented by the
zinc species are shown in Figs. 2, 3 and 4. Generally, water
molecules form intermolecular hydrogen bonds with polar
proton donor (OH, NH, NH2, NH3) and proton acceptor
C=O groups of neutral and zwitterionic forms of AAAs.
Neutral complexes AAA•M (M = Ni2+, Cu2+ and Zn2+)
In these complexes the M2+ ion is bicoordinated to nitrogen
and carbonyl oxygen (NO coordination) of AAAs. Selected
bond lengths, bond angles and dihedral angles of the fully
optimized neutral metal complexes and their hydrated
systems are given in Table 1. The hydration results in
different changes of the molecular structure of the parent
Parameters
(Phe_zwitt···M2+) x (H2O)n
Ni2+
(Trp_ zwitt ···M2+) x (H2O)n
Cu2+
Zn2+
Ni2+
(Tyr_ zwitt ···M2+) x (H2O)n
Cu2+
Zn2+
Ni2+
Cu2+
Zn2+
n=0
n=5
n=0
n=5
n=0
n=5
n=0
n=6
n=0
n=6
n=0
n=6
n=0
n=6
n=0
n=6
n=0
n=6
d[C(1)–C(2)]
1.525
1.513
1.568
1.508
1.527
1.515
1.539
1.523
1.566
1.557
1.536
1.511
1.532
1.514
1.569
1.520
1.531
1.512
d[C(2)–O(3)]
1.268
1.273
1.227
1.273
1.269
1.271
1.265
1.265
1.232
1.233
1.262
1.272
1.266
1.269
1.230
1.261
1.265
1.272
d[C(2)–O(4)]
1.256
1.271
1.271
1.274
1.263
1.271
1.247
1.263
1.267
1.272
1.256
1.272
1.252
1.271
1.267
1.268
1.259
1.272
d[C(1)–N(6)]
1.524
1.495
1.520
1.495
1.520
1.495
1.529
1.497
1.520
1.500
1.521
1.493
1.526
1.496
1.520
1.497
1.522
1.495
d[O(4)···M(5)]
2.063
2.024
1.921
1.963
2.082
2.015
2.056
2.083
1.924
1.884
2.083
2.021
2.080
2.049
1.926
2.018
2.062
2.016
d[O(3)···M(5)]
2.016
1.996
3.023
2.006
2.027
2.067
2.116
2.034
3.018
3.136
2.152
2.054
2.036
2.002
3.038
2.132
2.120
2.061
Θ[C(1)–C(2)–O(3)]
119.3
122.8
116.5
123.8
120.2
122.3
117.8
121.1
116.2
119.1
118.9
122.1
118.7
122.4
116.0
121.9
119.5
122.2
Θ[C(2)–C(1)–N(6)]
107.1
111.0
103.8
111.1
109.3
111.1
105.6
109.6
103.6
109.1
108.2
111.2
106.6
110.6
103.5
110.1
108.6
111.1
Θ[C(1)–O(4)···M(5)]
85.8
87.9
117.3
89.9
86.0
89.1
88.2
86.5
117.1
123.3
89.1
88.7
85.7
87.2
117.7
90.7
88.8
88.9
Θ[C(2)–O(3)···M(5)]
88.1
89.2
65.3
88.1
88.6
86.8
85.1
88.8
65.4
62.6
85.9
87.2
88.0
89.2
65.0
85.8
86.0
86.9
Θ[O(3)···M(5)···O(4)]
65.1
65.5
48.2
66.1
64.7
64.9
63.8
64.3
38.3
45.5
62.7
65.1
64.6
65.2
47.9
63.9
63.6
65.1
Φ[O(3)–C(2)–C(1)–N
(6)]
Φ[C(1)–C(2)–O(4)···M
(5)]
Φ[C(1)–C(2)–O(3)···M(5)]
-12.1
-17.2
7.9
-17.9
-9.1
-16.1
-2.5
-4.1
2.4
-8.6
-10.8
-13.3
-9.4
-20.5
4.9
-10.9
-5.2
-17.3
179.9
-177.9
-177.0
177.8
-177.9
-178.2
-179.1
-179.2
-175.1
-178.1
177.4
177.3
-179.8
-178.6
177.9
179.8
178.5
177.7
-179.8
178.1
176.6
-177.7
177.9
178.2
179.1
179.3
176.6
179.1
-177.5
-177.2
179.8
178.7
-177.3
-179.2
-178.5
-177.6
J Mol Model (2011) 17:3117–3128
Table 3 B3LYP/6-311+G(d,p) optimized relevant bond lengths (Å), bond angles (deg) and dihedral angles (deg) for metal-coordinated and hydrated complexes of the zwitterionic L-phenylalanine, Ltryptophan, and L-tyrosine species
3121
3122
J Mol Model (2011) 17:3117–3128
Fig. 2 Overall structure of the
B3LYP/6311+G(d,p) optimized
neutral complexes of chelated
phenylalanine, tyrosine and
tryptophan and their hydrates
complexes. In general, the covalent bond lengths C(1)–C
(2), C(2)–O(3), C(2)–O(4) and C(1)–N(6) of the AAAs are
only slightly changed upon hydration. However, microsolvation results in most cases in a slight prolongation of
the M2+···O and M2+···N bonds by about 0.02 – 0.05Å. The
M···O bonds are always shorter (about 0.02 – 0.07Å) than
the analogous M···N bonds (M=Ni2+, Cu2+ and Zn2+),
indicating that these metal cations form stronger bonds to
the oxygen atoms of AAAs. The equilibrium distances
M···O and M···N increase in the order: Ni2+ < Cu2+ ≤ Zn2+.
With regard to valence angles, upon hydration larger
changes (by about 1 – 4o) were observed only in the
values of the valence angle C(1)–N(6)···M(5). The O
(3)···M(5)···N(6) angle in the AAA complexes extends
over a relatively tiny interval (80 – 85o). The chelate rings
of the metallic complexes are not planar [dihedral angles
O(3)–C(2)–C(1)–N(6) and C(2)–C(1)–N(6)···M(5)], and
the deviation from planarity considerably increases with
microsolvation (Table 1).
Fig. 3 Overall structure of the
B3LYP/6311+G(d,p) optimized
π-complexes of chelated
phenylalanine, tyrosine,
tryptophan and their hydrates
Π-complexes AAA•M (M = Ni2+, Cu2+ and Zn2+)
The role of the cation – π side chain AAA interaction was
modeled using the complex in which the cation is bound to
both, the Phe, Tyr or Trp π face and to the –NH2 group
(Figs. 1 and 3). Ma et al. [16] has assumed the existence of
such interactions in their study of the alkali metal cation –
phenylalanine complexes. We have located these conformers as minimum for all structures except for the
BeckeHandHLYP calculations of the system Tyr•Cu2+,
which collapsed to the neutral conformer I. However, the
cation – π side chain interaction was preserved in the
hydrated complex Tyr•Cu2+(H2O)6. Important geometric
characteristics of these complexes are given in Table 2.
Like in the neutral system (I), the covalent bond lengths C
(1)–C(2), C(2)–O(3), C(2)–O(4) and C(1)–N(6) of the
AAAs are only slightly changed upon hydration. The
equilibrium distance M(5)···N(6) slightly increases in the
hydrated complexes (about 0.05Å). The valence angle C
J Mol Model (2011) 17:3117–3128
3123
Fig. 4 Overall structure of the
B3LYP/6-311+G(d,p)
optimized complexes of the
chelated zwitterionic form of
phenylalanine, tyrosine,
tryptophan and their hydrates
(1)–N(6)···M(5) and dihedral angles O(3)–C(2)–C(1)–N(6)
and C(2)–C(1)–N(6)···M(5) change appreciably in the
hydrated systems (by about 3 – 300), thus solvation of πcomplexes is connected with an appreciable structural
rearrangement.
Complexes zwitterionic AAA•M (M = Ni2+, Cu2+ and Zn2+)
For transition metal – AAA complexes III, Figs. 1 and 3,
represent the interaction of the zwitterionic phenylalanine,
tyrosine and tryptophan with metal cations. In these
complexes metal cations are bicoordinated to the carboxylate group of AAAs. Recently, Rimola et al. [9] computationally predicted zwitterionic complexes of Phe•Cu2+ and
Tyr•Cu2+ as the second most stable structure. Of the nine
AAA•M (M=Ni2+, Cu2+ and Zn2+) zwitterionic complexes
studied, the hydrated Trp•Cu2+(H2O)6 complex represents
the energetically most stable structure. The metal cation
coordination to the carboxylate group of zwitterionic AAAs
is almost symmetric. The only exception is the Cu2+cation
which in the isolated systems, due to electrostatic repulsion
between the copper dication and the positively charged –
NH3 group of AAAs, is coordinated unsymmetrically with
the shortest Cu2+•••O distance of about 1.9Å to the oxygen
Table 4 B3LYP/6-311+G(d,p)
calculated relative stability
( ΔG, in kJ mol-1) of neutral
complexes (I, II) and
zwitterionic (III) metal ion
complexes of the AAAs studied
Cation
Ni2+
Cu2+
a
BhandHLYP method
b
Collapsed to the conformer I
Zn2+
System
Without water
With water
Without water
With water
Without watera
With watera
Without water
With water
O(4) that is not acting as proton acceptor (Table 3). The
Cu2+•••O(3) distances were found to be considerably longer
(about 3Å). This highly asymmetric coordination of the
copper dication is also preserved in the hydrated complex
Trp(zwitt)•Cu2+(H2O)6. The metal cations lie in both,
isolated and hydrated complexes, in the plane of the
carboxylate group (dihedral angles Φ[C(2)–C(1)–O(4)···M
(5)] and Φ[C(1)–C(2)–O(3)···M(5)], Table 3). The O(3)···M
(5)···O(4) angle of the bifurcated metal bond is within a
relatively large interval of 40 – 65o. Thus, the position of
the metal cation at the binding site of zwitterionic AAAs
depends on both, the kind of acid and environment.
Relative energies
The relative energies with reference to the most stable
species of the 54 gas-phase and hydrated complexes studied
are given in Table 4. As it is evident from this Table, the
relative energies of the species studied vary in a relatively
large energy interval (by about 5 – 80 kJ mol-1) indicating
that in most cases gas-phase complexes should exist in one
most stable form. Similar net preference for one most stable
structure was theoretically found for histidine metallic
complexes [34]. Hydration changes the structure of the
(Phe···M2+) x (H2O)5
(Trp···M2+) x (H2O)6
(Tyr···M2+) x (H2O)6
I
II
III
I
II
III
I
II
III
0
0
0
0
0
0
35.6
0
6.8
43.6
52.6
48.3
87.6
63.4
0
42.4
44.6
56.3
26.5
19.9
36
42.7
27.5
33.4
0
0
0
21.2
0
0
66.1
0
39.9
4.8
47.1
23.0
65.4
22.1
0
13.3
51.3
46.0
28.6
0
35.7
8.6
65.2
72.9
0
0
0
0
0
0
26.4
5.7
49.2
23.1
64.4
11.2
0b
33.6
0
0
39.5
39.2
28.7
14.7
37.1
40.9
14.8
41.8
3124
J Mol Model (2011) 17:3117–3128
AAA•M (M=Ni2+, Cu2+ and Zn2+) complexes significantly.
Of the three types of systems investigated (I, II and III,
Fig. 1), the relative stability of the gas-phase complexes
computed with DFT methods suggest, with exception of
the Zn2+system, metallic complexes I of the neutral AAAs
to be the most stable species. In the most stable complex I
the Ni2+ and Cu2+ cations are bicoordinated to the AAA N
(6) and O(3) atoms, Figs. 1 and 2. In the gas-phase the
energetically unfavored zwitterionic Trp•Cu2+ complex is
stabilized upon solvation and represents the most stable
copper complex of tryptophane in aqueous solution
(Becke3LYP method, Table 4). It was recently demonstrated [9, 52] that functionals (e.g., BhandHLYP) with a
larger percentage of exact exchange (50%) than the most
commonly used Becke3LYP for open-shell systems compare better to the highly correlated CCSD(T) method.
Thus, for the AAA – Cu2+ complexes, in addition to
Becke3LYP we also carried out calculations with the
BHandHLYP functional. BHandHLYP relative Gibbs energies are for most cases by about 10 – 30 kJ mol-1 larger
(Table 4). The discrepancy in the results of the different DFT
methods used for the prediction of the relative stability of
open shell systems containing Cu2+ cations may be
explained by the shortcomings of the B3LYP method to
correctly describe the delocalized nature of the Cu2+
complexes [43]. The cation Zn2+ behaves differently. Due
to the large Gibbs energy gap between the most stable
complex II and the other systems (I and III) the most stable
complex II is dominant in the gas phase. Coordinated water
shifts the absolute stability towards complex I of the
Phe•Zn2+ and Trp•Zn2+ systems (Table 4). An interesting
situation is observed for the Tyr•Zn2+ complex. In both, the
gas phase and solvated state the π complex II has been
found to be the global minimum (Table 4). However, due to
the low energy barrier of about 5 kJ mol-1 both species (I
and II, respectively) may coexist in water.
Table 5 B3LYP/6-311+G(d,p)
calculated enthalpies ΔH,
entropies, ΔS, and Gibbs
energies ΔG, of the ion – AAA
systemsa
a
Data in parentheses were taken
from the work by Rimola et al. [9].
b
BHandHLYP method
Interaction enthalpies, entropies and Gibbs energies
Table 5 summarizes the interaction enthalpies, entropies and
Gibbs energies of the AAA•M (M=Ni2+, Cu2+ and Zn2+)
complexes computed at the Becke3LYP level of DFT
together with the data for copper complexes computed
using the BHandHLYP model and the published values
of interaction energies for copper complexes taken from
the work by Rimola et al. [9]. As it is evident from our
calculations of 54 AAA-metal cation species, the stability
of the different conformers and forms of AAAs can be
substantially modified by specific metal ion AAA interactions as well as by hydration. The interaction energies
can be computed for individual complexes of varying
relative energetic stability. However, only the results
presented for the thermodynamically most stable structures have a physical meaning. The interaction energies
were computed as the difference between the most stable
species (Table 4). For AAAs as reference state conformers
we used the conformers stabilized by means of the
OH•••NH2 hydrogen bond and an additional stabilizing
interaction between the amino group and the π-electron
system of the aromatic ring. These conformers have been
found previously as the most stable gas phase structures of
phenylalanine [53] and tryptophan [54]. The formation of
AAA – metal cation complexes as described in Eq. 1 is an
exothermic process. The association reaction is connected
with a large and negative, destabilizing effect of entropy
(Table 5). In the complexes studied this entropy effect
amounts from -105 to -150 J mol-1 K-1. The largest entropy
changes (some -150 J mol-1 K-1) are observed for AAA
complexes with Zn2+. In this case, however, the Zn2+ – π
interaction forces the polar amino group of the amino
acid to bend toward the aromatic ring (Fig. 3). The
importance of entropy effects for the relative stability of
individual conformers of phenylalanine was recently
System
Complex
ΔH298, kJ/mol
ΔS298, J/mol K
ΔG298, kJ/mol
I
I
Ib
II
I
I
Ib
II
I
I
Ib
II
Phe···Ni2+
Phe···Cu2+
Phe···Cu2+
Phe···Zn2+
Trp···Ni2+
Trp···Cu2+
Trp···Cu2+
Trp···Zn2+
Tyr···Ni2+
Tyr···Cu2+
Tyr···Cu2+
Tyr···Zn2+
-1051.8
-1199.9
-1104.1
-984.1
-1184.6
-1349.9
-1249.7
-1098.1
-1117.2
-1277.5
-1179.1
-1004.9
-120.6
-109.9
-105.4
-144.6
-112.1
-108.7
-107.6
-142.9
-124.2
-116.4
-115.2
-154.7
-1016.1
-1167.1
-1072.6
-940.9
-1151.2
-1317.5
-1217.6
-1055.4
-1080.1
-1242.8
-1144.8
-958.9
(-1196.6)
(-1096.5)
(-1346.0)
(-1243.1)
(-1272.4)
(-1171.1)
(-1165.7)
(-1066.1)
(-1311.7)
(-1209.2)
(-1238.0)
(-1137.2)
J Mol Model (2011) 17:3117–3128
3125
stressed by Kaczor et al. [6]. To take the entropy effects
into consideration, Gibbs energies, which describe the
tendency to associate in real molecular complexes, will be
discussed. The computed Gibbs energies ΔG298 are
negative and span a relatively narrow energy interval from
-940 to -1320 kJ mol-1. The energetic differentiation of the
complexes studied based on the type of the aromatic
substituent of AAAs shows an increasing interaction
Gibbs energy in the order Phe<Tyr<Trp for all three
metal cations studied. This ordering of stability correlates
very well with the experimental gas-phase basicities and
proton affinities of these aromatic amino acids [55]. For
reason of comparison, Table 5 also contains computed
interaction enthalpies and Gibbs energies for Cu2+ complexes of AAAs by Rimola et al. [9] using a slightly
different basis set. The results of Rimola et al. [9] are in
very good agreement with data obtained for the same
complexes by us. The Gibbs interaction energies show for
all three AAAs studied a decreasing binding affinity in the
order Cu2+ > Ni2+ > Zn2+. Thus, of the transition metal
cations studied, the copper dication is the most effective
Lewis acid in the interaction with the basic center of
AAAs. The largest interaction Gibbs energy was found for
the tryptophan - Cu 2+ complex (-1317.5 kJ mol -1 ,
Becke3LYP method). According to Marino et al. [41]
charge transfer plays an important role in the complexation
of the transition metal cations Ni2+, Cu2+, and Zn2+.
Moreover, the Cu2+ cation is an open shell system, and it
has been assumed that the different internal polarizability
of the whole Cu2+ complexes is probably one of the main
reasons for this difference. Rimola et al. [9] examined
complexes of monovalent and divalent copper with AAAs.
They showed that functionals such as BeckeHLYP with larger
percentages of exact exchange (50%) than the commonly used
Becke3LYP (20%) compare better to the highly correlated ab
initio CCSD(T) method for these open-shell systems. We also
did calculations of all open-shell Cu2+ complexes with the
half-and-half functional BeckeHandHLYP implemented in
the G03 computer code. BeckeHandHLYP relative Gibbs
energies for AAA•Cu2+ complexes were computed to be by
about 100 kJ mol-1 lower than with B3LYP (Table 5).
According to Rimola et al. [9] the variation in interaction
energy using two functionals may be at least partly explained
by the different results of calculations of the second
ionization energy for copper using Becke3LYP and BeckeHandHLYP methods, respectively. Sousa et al. [43] explain
the discrepancy in the results of the different DFT methods
used for the prediction of the relative stability of open shell
systems containing the Cu2+ cation by the shortcomings of
the Becke3LYP method to correctly describe the delocalized
nature of the Cu2+ complexes.
Electronic structures
The atomic charges of the complexes studied were
evaluated by natural population analysis [56, 57] using
the NBO program [58]. The natural charges (NAC), Wiberg
bond indices (WBI) and amount of charge transfer (CT) for
the most stable species are shown in Table 6. The
coordination of these transition metal cations that are
borderline hard acids [59] appreciably increases polarization of AAAs studied. This polarizing effect of bivalent
cations is much higher in the case of Lewis acids Ni2+ and
Cu2+, respectively. These Lewis acids accept very large
Table 6 Natural atomic charges (NAC), Wiberg bond indices (WBI) and amount of charge transfer (CT) from the B3LYP/6-311+G(d,p) NBO
analysis
System
Complex
NAC/e
WBI
CT/e
O(3)
N(6)
M2+
O(3)···M2+
N(6)···M2+
I
I
Phe···Ni2+
Phe···Cu2+
-0.679
-0.663
-0.912
-0.917
1.089
0.936
0.1448
0.0874
0.1927
0.1419
0.911
1.064
Ia
II
I
I
Ia
II
I
I
Ia
II
Phe···Cu2+
Phe···Zn2+
Trp···Ni2+
Trp···Cu2+
Trp···Cu2+
Trp···Zn2+
Tyr···Ni2+
Tyr···Cu2+
Tyr···Cu2+
Tyr···Zn2+
-0.700
-0.933
-0.982
-0.908
-0.928
-0.936
-0.998
-0.907
-0.928
-0.935
-0.993
0.937
1.585
0.851
0.905
0.934
1.459
0.963
0.907
0.935
1.532
0.0611
0.0942
0.3259
0.1736
0.1495
0.1013
0.2856
0.1757
0.1450
0.0982
0.3128
1.063
0.415
1.149
1.095
1.066
0.545
1.037
1.093
1.065
0.468
a
BHandHLYP method
-0.660
-0.668
-0.709
-0.667
-0.663
-0.705
0.1264
0.0871
0.0634
0.1313
0.0858
0.0624
3126
J Mol Model (2011) 17:3117–3128
amount of negative charge (0.9 – 1.1e). Thus the role of
charge transfer for bonding of the AAAs studied to nickel
and copper dications is of particular importance.
More interesting situation is with the Zn2+ complexes.
Most stable species correspond to the π complex II. The
cation−π interactions are electrostatic in nature, accounting
for 60 – 90% of the overall binding [16]. This fact is also
manifested in our calculations of the zinc-coordinated
complexes. The magnitude of charge transfer contribution
is much lower (ca. 0.4 – 0.5e). However, zinc is sometimes
considered a transition metal and sometimes a main group
element. According to Amin and Truhlar computational
chemistry requirements of zinc coordination resemble those
of the main group rather than the first transition row [60]. Its
interaction with aromatic amino acids through cation−π
interaction resembles similar interaction of main group metal
cations with aromatic motifs of amino acids [11, 13, 14].
As is evident from Table 6 the quantitative importance
of CT term varies considerably among the ions. It is
frequently argued that metal ion interaction energies well
correlate with electron affinity of metal cation [61]. Hoyau
et al. in their theoretical study of the interaction of glycine
wit 15 metal cations have shown that the magnitude of
charge transfer correlates well with the electron affinity
(EA) of the metal ion [62]. In the case of metal cations
studied by us their electron affinities (18.17, 20.29 and
17.96 eV for s Ni2+, Cu2+, and Zn2+, respectively [63])
exhibit irregularity with respect to CT. For transition metal
cations Ni2+ and Cu2+, the larger the metal ion EA the
larger the CT. Although Zn2+ electron affinity is large and
comparable in magnitude with that for Ni2+ cation, its
magnitude does not quantitatively correlate with charge
transfer. Moreover, the amount of CT from the amino acid
toward the M2+ ions also does not correlate with the
interaction energies of the complexes investigated
(Tables 5 and 6).
interaction of the Zn2+ cation with the AAAs studied
results in the π − complex II as the most stable
structure.
3. Solvation changes the relative stability of individual
species in several cases. Coordinated water shifts the
absolute stability toward the complex I of the Phe•Zn2+
and Trp•Zn2+ systems. On the other hand, the π−
complex II of the Tyr•Zn2+ system has been found to
be the global minimum in both, gas phase and solvated
state.
4. Among the Lewis acids investigated, the strongest
affinity to the aromatic amino acids studied resulted for
the Cu2+ cation. The largest interaction Gibbs energy
was found for the tryptophan – Cu2+ complex.
Summary and conclusions
References
This theoretical study was set out to determine stable
conformers for the 54 complexes of phenylalanine, tyrosine
and tryptophan with the transition metal cations Ni2+, Cu2+
and Zn2+ and their hydrated forms. From the results the
following conclusions can be drawn.
1. For phenylalanine, tyrosine and tryptophan complexes
with Ni2+ and Cu2+ the most stable structure corresponds to the complex in which the metal cation binds
to the oxygen of the carbonyl group and to the nitrogen
of the amine (system I).
2. The π−complexes of AAAs with metal cations studied
represent a new kind of low−energy conformers. The
This work yields quantities that may be inaccessible or
complementary to experiments and represents the first
quantum chemical approach in which both the gas-phase
and solvated phase complexation between aromatic amino
acids (phenylalanine, tyrosine and tryptophan) and transition metal cations Ni2+, Cu2+ and Zn2+ are considered and
absolute metal ion affinities were evaluated. The results are
also relevant for attempts to design peptides selectively
binding different metal ions [64].
Acknowledgments The authors thank the Austrian Science
Foundation (FWF), and the Slovak Ministry of Education (Grant
No. 1/0084/10) for financial support. The European Union HPC−
Europa Transnational Access Program under the Project HPC−
Europa2 (Project No 458) at SARA Amsterdam (M.R.) has also
supported this work. The authors acknowledge with thanks the
Stichting Academisch Rekencentrum Amsterdam (SARA) for the
use of its resources and for excellent support.
Open Access This article is distributed under the terms of the
Creative Commons Attribution Noncommercial License which permits any noncommercial use, distribution, and reproduction in any
medium, provided the original author(s) and source are credited.
1. Bush MF, Oomens J, Saykally RJ, Williams ER (2008) Effects of
alkaline earth metal ion complexation on amino acid zwitterion
stability: results from infrared action spectroscopy. J Am Chem
Soc 130:6463–6471. doi:10.1021/ja711343q
2. Olsztynska-Janus S, Szymborska K, Komorowska M, Lipinski J
(2009) Conformational changes of L–phenylalanine – Near infrared–
induced mechanism of dimerization: B3LYP studies. J Mol Struct
THEOCHEM 911:1–7. doi:10.1016/j.theochem.2009.06.046
3. Sobolewski AL, Dorit Shemesh D, Domcke W (2009) Computational
studies of the photophysics of neutral and zwitterionic amino acids in
an aqueous environment: tyrosine−(H2O)2 and tryptophan−(H2O)2
clusters. J Phys Chem A 113:542–550. doi:10.1021/jp8091754
4. Costanzo F, Valle RG, Barone V (2005) MD simulation of the
Na+−phenylalanine complex in water: competition between
cation− π interaction and aqueous solvation. J Phys Chem B
109:23016–23023. doi:10.1021/jp055271g
J Mol Model (2011) 17:3117–3128
5. Jover J, Bosque R, Sales J (2008) A comparison of the binding
affinity of the common amino acids with different metal cations.
Dalton Trans 6441–6453. doi:10.1039/b805860a
6. Kaczor A, Reva ID, Proniewicz LM, Fausto R (2006)
Importance of entropy in the conformational equilibrium of
phenylalanine: a matrix−isolation infrared spectroscopy and
density functional theory study. J Phys Chem A 110:2360–
2370. doi:10.1021/jp0550715
7. Larrucea J, Rezabal E, Marino T, Russo N, Ugalde JM (2010) Ab
initio study of microsolvated Al3+−aromatic amino acid complexes. J Phys Chem B 114:9017–9022. doi:10.1021/jp101874p
8. Polfer NC, Oomens J, Moore DT, Helden G, Meijer G, Dunbar
RC (2006) Infrared spectroscopy of phenylalanine Ag(I) and Zn
(II) complexes in the gas phase. J Am Chem Soc 128:517–525.
doi:10.1021/ja0549291
9. Rimola A, Rodríguez-Santiago L, Sodupe M (2006) Cation–π
Interactions and Oxidative Effects on Cu+ and Cu2+ Binding to
Phe, Tyr, Trp, and His Amino Acids in the Gas Phase. Insights
from first–principles calculations. J Phys Chem B 110:24189–
24199. doi:10.1021/jp064957l
10. Rimola A, Sodupe M, Tortajata J, Rodríguez-Santiago L (2006)
Gas phase reactivity of Cu+−aromatic amino acids: An experimental and theoretical study. Int J Mass Spectrom 257:60–69.
doi:10.1016/j.ijms.2006.06.017
11. Ryzhov V, Dunbar RC, Cerda B, Wesdemiotis C (2000) Cation–π
effects in the complexation of Na+ and K+ with Phe, Tyr, and Trp
in the gas phase. J Am Soc Mass Spectrom 11:1037–1046
12. Ruan C, Rodgers MT (2004) Cation−π interactions: structures
and energetics of complexation of Na+ and K+ with the aromatic
amino acids, phenylalanine, tyrosine, and tryptophan. J Am Chem
Soc 126:14600–14610. doi:10.1021/ja048297e
13. Siu FM, Ma NL, Tsang CW (2001) Cation−π interactions in
sodiated phenylalanine complexes: is phenylalanine in the
charge − solvated or zwitterionic form? J Am Chem Soc
123:3397–3398. doi:10.1021/ja0056872
14. Siu FM, Ma NL, Tsang CW (2004) Competition between π and
non−π cation−binding sites in aromatic amino acids: a theoretical
study of alkali metal cation (Li+, Na+, K+)−phenylalanine complexes. Chem Eur J 10:1966–1976. doi:10.1002/chem.200305519
15. Gallivan JP, Dougherty DA (1999) Cation−π interactions in
structural biology. PNAS 96:9459–9464
16. Ma JC, Gougherty DA (1997) The cation−π interaction. Chem
Rev 97:1303–1324. doi:10.1021/cr9603744
17. Michael LA, Chenault JA, Miller BR III, Knolhoff AM, Nagan MC
(2009) Water, shape recognition, salt bridges, and cation–pi interactions differentiate peptide recognition of the HIV rev−responsive
element. J Mol Biol 392:774–786. doi:10.1016/j.jmb.2009.07.047
18. Wright SN, Wang SY, Wang GK (1998) Lysine point mutations in
Na+ channel D4−S6 reduce inactivated channel block by local
anesthetics. Mol Pharmacol 54:733–739
19. Torrice MM, Bower KS, Lester HA, Dougherty DA (2009)
Probing the role of the cation–π interaction in the binding sites of
GPCRs using unnatural amino acids. PNAS 106:11919–11924.
doi:10.1073/pnas.0903260106
20. Dudev T, Lim C (2003) Principles governing Mg, Ca, and Zn
binding and selectivity in proteins. Chem Rev 103:773–787.
doi:10.1021/cr020467n
21. Rode BM (1999) Peptides and the origin of life. Peptides 20:773–
786. doi:10.1016/S0196−9781(99)00062−5
22. Rode BM, Bujdák J, Eder AH (1993) The role of inorganic
substances in the chemical evolution of peptides on earth. Trends
Inorg Chem 3:45–62
23. Plankensteiner K, Reiner H, Rode BM (2005) Stereoselective
differentiation in the salt−induced peptide formation reaction and
its relevance for the origin of life. Peptides 26:535–541.
doi:10.1016/j.peptides.2004.11.019
3127
24. Remko M, Rode BM (2000) Bivalent cation binding effect on
formation of the peptide bond. Chem Phys Lett 316:489–494.
doi:10.1016/S0009−2614(99)01322−6
25. Remko M, Rode BM (2001) Catalyzed peptide bond formation in
the gas phase. Phys Chem Chem Phys 3:4667–4673. doi:10.1039/
b105623a
26. Remko M, Rode BM (2004) Catalyzed peptide bond formation
in the gas phase. Role of bivalent cations and water in
formation of 2−aminoacetamide from ammonia and glycine
and in dimerization of glycine. Struct Chem 15:223–232.
doi:10.1023/B:STUC.0000021531.69736.00
27. Frisch MJ et al. (2003) Gaussian 03, Revision B.04. Gaussian Inc,
Pittsburgh PA
28. Hehre WJ, Radom L, Schleyer PvR, Pople JA (1986) Ab initio
molecular orbital theory. Wiley, New York
29. Kohn W, Sham LJ (1965) Self−consistent equations including
exchange and correlation effects. Phys Rev 140:A1133–A1138.
doi:10.1103/PhysRev.140.A1133
30. Bickelhaupt FM, Baerends EJ (2000) Kohn −Sham density
functional theory: predicting and understanding chemistry. In:
Lipkowitz KB, Boyd DB (eds) Rev Comput Chem. Wiley, New
York, 15: pp 1–86
31. Becke AD (1993) Density−functional thermochemistry. III. The
role of exact exchange. J Chem Phys 98:5648–5652. doi:10.1063/
1.464913
32. Lee C, Yang W, Parr RG (1988) Development of the Colle−Salvetti
correlation−energy formula into a functional of the electron density.
Phys Rev B 37:785–789. doi:10.1103/PhysRevB.37.785
33. Becke AD (1993) A new mixing of Hartree–Fock and local
density − functional theories. J Chem Phys 98:1372–1377.
doi:10.1063/1.464304
34. Remko M, Fitz D, Rode BM (2010) Effect of metal ions (Li+,
Na+, K+, Mg2+, Ca2+, Ni2+, Cu2+ and Zn2+) and water coordination on the structure and properties of L−histidine and zwitterionic L−histidine. Amino Acids 39:1309–1319. doi:10.1007/
s00726−010−0573−8
35. Wachers AJH (1970) Gaussian Basis Set for Molecular Wavefunctions Containing Third−Row Atoms. J Chem Phys 52:1033–
1036. doi:10.1063/1.1673095
36. Hay PJ (1977) Gaussian basis sets for molecular calculations. The
representation of 3d orbitals in transition metal atoms. J Chem
Phys 66:4377–4384. doi:10.1063/1.433731
37. Johnson BG, Gill PMW, Pople JA (1993) The performance of a
family of density functional methods. J Chem Phys 98:5612–
5626. doi:10.1063/1.464906
38. Oliphant N, Bartlett RJ (1994) A systematic comparison of
molecular properties obtained using Hartree–Fock, a hybrid
Hartree–Fock density−functional−theory, and coupled−cluster
methods. J Chem Phys 100:6550–6561. doi:10.1063/1.467064
39. Bernardi F, Bottoni A, Garavelli M (2002) Exploring organic
chemistry with DFT: radical, organo−metallic, and Bio−organic
applications. Quant Struct Act Rel 21:128–148
40. Remko M, Rode BM (2006) Effect of Metal Ions (Li+, Na+, K+,
Mg2+, Ca2+, Ni2+, Cu2+, and Zn2+) and water coordination on the
structure of glycine and zwitterionic glycine. J Phys Chem A
110:1960–1967. doi:10.1021/jp054119b
41. Marino T, Russo N, Toscano M (2000) Gas−phase metal ion (Li+,
Na+, Cu+) affinities of glycine and alanine. J Inorg Biochem
79:179–185. doi:10.1016/S0162−0134(99)00242−1
42. Sousa SF, Fernandes PA, Ramos MJ (2007) Comparative
assessment of theoretical methods for the determination of
geometrical properties in biological zinc complexes. J Phys Chem
B 111:9146–9152. doi:10.1021/jp072538y
43. Sousa SF, Fernandes PA, Ramos MJ (2007) General performance
of density functionals. J Phys Chem A 111:10439–10452.
doi:10.1021/jp0734474
3128
44. Wu R, McMahon TB (2008) Investigation of cation−π interactions in biological systems. J Am Chem Soc 130:12554–12555.
doi:10.1021/ja802117s
45. Reddy AS, Sastry NG (2005) Cation [M=H+, Li+, Na+, K+, Ca2+,
Mg2+, NH4+, and NMe4+] interactions with the aromatic motifs of
naturally occurring amino acids: a theoretical study. J Phys Chem
A 109:8893–8903. doi:10.1021/jp0525179
46. Freeman HC (1967) Crystal structures of metal−peptide complexes. Adv Protein Chem 22:257–424. doi:10.1016/S0065−3233
(08)60043−1
47. Yamada S, Terashima S, Wagatsuma M (1970) A simple and
novel peptide bond formation by copper(II) complex. Tetrahedron
Lett 11:1501–1504. doi:10.1016/S0040−4039(01)98006−2
48. Siegel H, Martin RB (1982) Coordinating properties of the amide
bond. Stability and structure of metal ion complexes of peptides
and related ligands. Chem Rev 82:385–426. doi:10.1021/
cr00050a003
49. Jockusch RA, Price WD, Williams ER (1999) Structure of
cationized arginine (Arg·M+, M=H, Li, Na, K, Rb, and Cs) in
the gas phase: further evidence for zwitterionic arginine. J Phys
Chem A 103:9266–9274. doi:10.1021/jp9931307
50. Cerda BA, Wesdemiotis CH (2000) Zwitterionic vs. charge−
solvated structures in the binding of arginine to alkali metal
ions in the gas phase. Analyst 125:657–660. doi:10.1039/
a909220j
51. O’Brien JT, Prell JS, Steill JD, Oomens J, Williams ER (2008)
Interactions of mono−and divalent metal ions with aspartic and
glutamic acid investigated with IR photodissociation spectroscopy
and theory. J Phys Chem A 112:10823–10830. doi:10.1021/
jp805787e
52. Rimola A, Rodríguez-Santiago L, Ugliengo P, Sodupe M (2007)
Is the pepride bond formation activated by Cu2+ interactions?
Insights from density functional calculations. J Phys Chem B
111:5740–5747. doi:10.1021/jp071071o
53. Snoek LC, Robertson EG, Kroemer RT, Simons JP (2000)
Conformational landscapes in amino acids: infrared and ultraviolet
J Mol Model (2011) 17:3117–3128
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
ion−dip spectroscopy of phenylalanine in the gas phase. Chem Phys
Lett 321:49–56. doi:10.1016/S0009−2614(00)00320−1
Snoek LC, Kroemer RT, Hockridge MR, Simons JP (2001)
Conformational landscapes of aromatic amino acids in the gas
phase: Infrared and ultraviolet ion dip spectroscopy of tryptophan.
Phys Chem Chem Phys 3:1819–1826. doi:10.1039/B101296G
Harrison AG (1997) The gas−phase basicities and proton
affinities of amino acids and peptides. Mass Spectrom Rev 16:201–
217
Reed AE, Weinstock RB, Weinhold FA (1985) Natural population
analysis. J Chem Phys 83:735–746. doi:10.1063/1.449486
Reed AE, Weinhold FA (1985) Natural localized molecular
orbitals. J Chem Phys 83:1736–1740. doi:10.1063/1.449360
Glendening ED, Reed AE, Carpenter JE, Weinhold F. The NBO
program, Version 3.1
Shriver DF, Atkins PW, Langford CH (1996) Inorganic Chemistry
Second Ed. Oxford University Press, Oxford, pp 212–215
Amin EA, Truhlar DG (2008) Zn coordination chemistry:
development of benchmark suites for geometries, dipole
moments, and bond dissociation energies and their use to test
and validate density functionals and molecular orbital theory. J
Chem Theory Comput 4:75–85. doi:10.1021/ct700205n
Martin RB (1998) Metal ion stabilities correlate with electron
affinity rather than hardness or softness. Inorg Chim Acta 283:30–
36. doi:10.1016/S0020-1693(98)00087-5
Hoyau S, Pélicier JP, Rogalewicz F, Hoppilliard Y, Ohanessian G
(2001) Complexation of glycine by atomic metal cations in the
gas phase. Eur J Mass Spectrom 7:303–311. doi:10.1255/ejms.440
Pearson RG (1988) Absolute electronegativity and hardness:
application to inorganic chemistry. Inorg Chem 27:734–740.
doi:10.1021/ic00277a030
Rulíšek L, Havlas Z (2003) Theoretical studies of metal ion
selectivity. 3. A theoretical design of the most specific combinations of functional groups representing amino acid side chains for
the selected metal ions (Co2+, Ni2+, Cu2+, Zn2+, Cd2+, and Hg2+).
J Phys Chem B 107:2376–2385. doi:10.1021/jp026951b