ArticlePDF Available

A bond-order analysis of the mechanism for hydrated proton mobility in liquid water

Authors:

Abstract and Figures

Bond-order analysis is introduced to facilitate the study of cooperative many-molecule effects on proton mobility in liquid water, as simulated using the multistate empirical valence-bond methodology. We calculate the temperature dependence for proton mobility and the total effective bond orders in the first two solvation shells surrounding the H(5)O(2) (+) proton-transferring complex. We find that proton-hopping between adjacent water molecules proceeds via this intermediate, but couples to hydrogen-bond dynamics in larger water clusters than previously anticipated. A two-color classification of these hydrogen bonds leads to an extended mechanism for proton mobility.
Content may be subject to copyright.
A bond-order analysis of the mechanism for hydrated proton mobility
in liquid water
Hadas Lapid and Noam Agmon
a)
Department of Physical Chemistry and the Fritz Haber Research Center, The Hebrew University,
Jerusalem 91904, Israel
Matt K. Petersen and Gregory A. Voth
b)
Department of Chemistry and the Center for Biophysical Modeling and Simulation, University of Utah,
Salt Lake City, Utah 84112-0850
Received 8 July 2004; accepted 20 September 2004; published online 14 December 2004
Bond-order analysis is introduced to facilitate the study of cooperative many-molecule effects on
proton mobility in liquid water, as simulated using the multistate empirical valence-bond
methodology. We calculate the temperature dependence for proton mobility and the total effective
bond orders in the first two solvation shells surrounding the H
5
O
2
proton-transferring complex. We
find that proton-hopping between adjacent water molecules proceeds via this intermediate, but
couples to hydrogen-bond dynamics in larger water clusters than previously anticipated. A two-color
classification of these hydrogen bonds leads to an extended mechanism for proton
mobility. © 2005 American Institute of Physics. DOI: 10.1063/1.1814973
I. INTRODUCTION
Proton mobility in liquid water has attracted much atten-
tion in the last century,
1
with efforts intensifying during the
last decade. This is demonstrated, for example, by the num-
ber of papers on this subject in one special issue.
2–5
The
efforts to elucidate the mechanism of proton mobility in wa-
ter are motivated by the role protons play in acid-base reac-
tions in aqueous solutions, in environmental chemistry, and
in bioenergetics, where energy is transiently stored as trans-
membranal proton gradients.
6
A mechanism for proton mobility was suggested to com-
prises the following ingredients.
1,7,8
i Cyclic isomerization between the two forms of pro-
tonated water: a The more stable H
3
O
is transiently con-
verted into H
5
O
2
and back;
1
or else b one H
5
O
2
converts
directly into another.
7
ii This interconversion is coupled to hydrogen-bond
HB dynamics in the second solvation shell of the H
3
O
.
Since the coordination number of H
3
O
is 3 whereas
that of liquid water is close to 4, it was conjectured that the
transfer event is preceded by HB cleavage to the acceptor
water molecule, and followed by HB formation to the donor
molecule.
1
While this mechanism has found its way into
physical chemistry textbooks,
9
the issue is still under active
investigation through molecular dynamics MDsimulations.
MD simulations of protonated water are complicated by
the need to find a good representation for the interaction
potential. Two major approaches have been invoked: calcu-
lation of the potential ‘on the fly,’’ at every time step, using
density functional theory,
7,8
and use of multistate empirical
valence bond EVB potentials.
10–20
The second method is less costly computationally, hence,
allows one to run sufficiently long trajectories for gathering
the required statistics. The initially implemented two-state
EVB Refs. 11 and 12 has been extended into multistate
MSEVB, with parameters calibrated to reproduce ab initio
data on small protonated water clusters and for the proton
solvated in bulk water.
15–20
The MS-EVB potential allows
for proton delocalization among several water molecules in a
deterministic fashion. There are important differences be-
tween the MS-EVB approaches of Refs. 1315 and 1620
so the reader is referred to these papers for more detail. Ad-
ditional methodologies are rapidly accumulating,
21–23
but
they are of a more phenomenological nature.
Early simulations actually indicated that proton migra-
tion involves a concerted double proton transfer PT event,
which converts one H
5
O
2
moiety into another directly, with-
out a special role for the H
3
O
cation.
2,7,14,23
However, more
recent Car-Parrinello CP path-integral simulations
8
indi-
cated the dominance of single PT events. These convert a
H
3
O
into a H
5
O
2
moiety and back, as first suggested in
Ref. 1.
The question of single vs double PT events (H
3
O
to
H
5
O
2
vs H
5
O
2
to H
5
O
2
) depends on the relative stability of
the two cations. Whereas in Ref. 7 they are nearly isoener-
getic, in the MS-EVB algorithm used below
17,18,20
H
3
O
is
more stable than H
5
O
2
by about 1 kcal/mol. One expects
that in trajectories applying this potential, H
5
O
2
features as a
transient intermediate structure between two more stable
H
3
O
structures.
This difference in stability, in turn, may depend on dif-
ferences in HB strengths for the different protonated water
potentials, with stronger HBs favoring the H
3
O
ion. For
example, in a recent analysis of the CP methodology for
liquid water,
24
it has been shown that use of a larger fictitious
electronic mass
in the CP Lagrangian artificially produces
a
Author to whom correspondence should be addressed. Electronic mail:
agmon@fh.huji.ac.il
b
Electronic mail: voth@chem.utah.edu
THE JOURNAL OF CHEMICAL PHYSICS 122, 014506 2005
122, 014506-10021-9606/2005/122(1)/014506/11/$22.50 © 2005 American Institute of Physics
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
good self-diffusion coefficients D
water
and radial distribution
functions g(r). When a smaller
is used to better ensure
that the trajectory remains near the Born-Oppenheimer sur-
face, g(r) from the CPMD simulations of water, using two
popular generalized gradient approximations for the density
functional, become too structured compared to experiment
and D
water
decreases significantly relative to the correct
value. These calculations indicate that the HBs produced by
using larger
values are evidently accidentally weaker and
this reduces the structure in g(r) and enhances D
water
]. A
large
value 1100 a.u. was used in the CP simulation of
Ref. 7.
It remains to consider the more difficult question,
whether and how the PT events couple to HB dynamics. The
mechanisms discussed in the literature
1,7,8
suggested that the
rate determining step is cleavage of the HB donated from a
second-shell water molecule to the first-shell proton accep-
tor. A variant of this scenario was indeed observed in simu-
lations of proton mobility in ice.
25
Instead of complete cleav-
age and formation of HBs, impeded by the rigidity of the ice
structure, these simulations showed an interplay between
weakening increased length followed by the strengthening
decreased length of the two red HBs in Fig. 1b.
In protonated liquid water simulations, efforts were
made to follow the coordination number of the acceptor and
donor water molecules,
2
or else the angle between the donor,
acceptor, and the HB donated to it.
19,20
While some evidence
for the suggested role of this HB was detected, the average
change in the coordination number was much smaller than
unity. These results indicate that the rate-limiting step for PT
does not reside solely in the specified HB, as previously
suggested.
1,7–9
The assumption that may break down here is that the
first-shell water ligands behave like bulk water, possessing a
coordination number of 3.9. Indeed, femtosecond pump-
probe near-IR measurements suggest that the first-shell water
molecules around cations or anions exhibit slower reorienta-
tional times than bulk water,
26
indicating stronger HBs. Simi-
larly, the three first-shell neighbors of a H
3
O
ion must form
extra-strong HBs to it.
27
This helps in delocalizing 20%
30% of the positive protonic charge on the three first-shell
ligands.
20
Consequently, it becomes electrostatically unfavor-
able to donate a HB to these oxygen atoms.
28
The missing
HB leads to the reduction in coordination number, from 3.9
in bulk water to about 3.6 for the first-shell ligands.
3,20
Therefore, cleavage of a HB donated to the acceptor
water molecule cannot be rate limiting, because it simply
does not exist for 40% of the time. Rather, evidence for such
behavior has been detected one water layer further away.
20
Since there are four first-shell neighbors to a H
5
O
2
, and
these are further engaged in up to 12 additional HBs, this
means that a full description of proton mobility in liquid
water involves the participation of larger water clusters than
previously anticipated, at least as large as depicted in Fig.
1c. This agrees with earlier observations of Ohmine and
collaborators
29,30
that dynamic processes in water are driven
by large-scale collective motions.
The problem is to follow many HBs simultaneously and
average their effect in an appropriate manner. This is ad-
dressed by the ‘bond order analysis’ BOAproposed in the
present work. Utilizing it we are able to obtain concrete in-
sight into the microscopic mechanism of proton mobility, at
least within the MS-EVB2 model.
The present work is structured as follows: After briefly
reviewing the simulation methodology Sec. II, we present
simulation results on the temperature dependence of the pro-
ton diffusion coefficient Sec. III. From it we conclude that
within a limited range of temperatures say 280310 K,
there is little change in the mechanism of proton mobility. To
reduce thermal noise, we choose to work at 280 K. The prin-
ciples of BOA are then outlined in Sec. IV. Its main results
are described in Sec. V. These are based predominantly on
the notion of the total effective bond order TEBO, which
invokes a two-color classification of HBs. We conclude Sec.
VIby suggesting an extended version for the mechanism of
proton mobility in water, which involves the collective reor-
ganization of both types of HBs in the first- and second-
solvation layers of the transferring H
5
O
2
complex.
FIG. 1. Color The proton-transferring complex H
5
O
2
, and its first two
solvation shells. In the first shell panel b兲兴, six HBs are tracked see Fig. 3
for their color codes. In the second shell panel c兲兴, 12 HBs are tracked.
The two unfavorable HBs from the first shell redare not followed onto the
second shell.
014506-2 Lapid
et al.
J. Chem. Phys. 122, 014506 (2005)
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
II. SIMULATION METHODOLOGY
Classical MD simulations of a single proton in a cubic
box of 125 water molecules were run using the Schmitt and
Voth MS-EVB program,
4,16,17
version 2.
20
The box linear di-
mension was 15.6 Å corresponding to a density of 1.0 gr/cc
at 300 K, and periodic boundary conditions were imposed
on its walls. Using time steps of 0.5 fs, a trajectory was
equilibrated for at least 150 ps 300000 time steps at the
desired temperature NVT ensemble. Five different tem-
peratures were used for the calculation of the proton diffu-
sion coefficient D
H
, whereas the mechanistic study was
performed at a single temperature, T 280 K. After setting
the temperature, the thermostat was turned off and the trajec-
tory was continued at constant energy NVE ensemble.
In order to quantify D
H
, the center-of-excess-charge
CEC coordinate was utilized in the MS-EVB methodology
as outlined in Ref. 20. The coordinates of the CEC were
tracked from time step to time step. If, at any given time
step, the coordinates varied from the previous step by the
approximate length of a periodic image they were accord-
ingly adjusted. This constructed a trajectory where the CEC
effectively diffused through an infinite space of periodic im-
ages relative to an origin in the original MD cell.
D
H
was calculated at the five temperatures using simi-
lar procedures. For example, at 275 K, 30 starting configu-
rations were collected from a 300 ps NVT trajectory every
10 ps. These configurations were then used to start 500 ps
NVE runs. The mean squared displacement was calculated
for each NVE trajectory and a least squares fit was obtained
between 10 and 50 ps. This interval was chosen so the mea-
surement would be above the nonlinear regime and below
the noisiest portion of the line. The slope was used to esti-
mate D
H
. This value was then averaged over the 30 runs.
For the mechanistic study, 13 NVE trajectories, of length
30 ps each, were run after separate NVT equilibrations. The
atomic coordinates were saved every 25 fs. These were
remapped onto the central unit cell with the H
3
O
at the
origin. Any water molecule that got broken across the peri-
odic boundaries was reconnected. This gave a corresponding
trajectory file which was suitable for visualization and
analysis.
Using a visualization program gOpenMol version 2.2,
by Leif Laaksonen, PT events between adjacent water mol-
ecules were identified. To avoid possible correlations be-
tween events, only the first few events were considered from
each 30 ps trajectory which was followed by an equilibra-
tion period. Aborted or incomplete transfers were not in-
cluded in this study. We have thus collected an arbitrary set
of 25 clear-cut PT events for analysis.
For each PT event, a trajectory segment was rerun start-
ing at 12 ps before and ending 12 ps after the event. The
coordinates were then saved at 5 fs intervals, to provide a
more detailed picture of the dynamics, and remapped onto
the central cell as above. These refined trajectories form the
data base for our mechanistic study.
Each of the refined trajectories is characterized by one
H
5
O
2
moiety composed of the donor and acceptor water
molecules. Its first- and second-shell neighboring water
molecules were identified from the atomic Cartesian coordi-
nates via a minimal distance criterion. This procedure was
repeated every time step, so that if a solvent and bulk water
molecules got interchanged the new solvent molecule was
followed. As depicted schematically in Fig. 1, the coordi-
nates of a total of 20 water molecules are tracked every time
step. From them, we calculate the 20 HB distances shown in
the figure as dashed lines. The HBs are separated into donor
and acceptor types, as indicated by the blue and red colors in
Figs. 1b and 1c.
III. THE TEMPERATURE DEPENDENCE
OF PROTON MOBILITY
Previous MS-EVB simulations have calculated the protic
CEC mean-square displacement, and hence D
H
,at
T300 K e.g., Fig. 4 in Ref. 2 and Fig. 8 in Ref. 20. Here
we extend these studies to obtain the temperature depen-
dence of D
H
for the MS-EVB2 model over a range of tem-
peratures, 260320 K.
Figure 2 shows our results for five temperature values
squares. Due to the classical nuclear dynamics imple-
mented here, the absolute value of D
H
is too small in com-
parison with experiment. It should be noted that quantization
of the MS-EVB model, using the path integral centroid mo-
lecular dynamics method, reproduces this quantum effect
very well, which is mainly due to simple mode quantization
of the hydrogen bond.
17–19
The ratio of classical to quantum
proton hopping rates was shown in earlier calculations to be
0.56 Table III in Ref. 17. Additionally, there may be some
small contribution to the diffusion rate from correlated pro-
ton hops over more than two water molecules that are not
completely captured by the MS-EVB2 model. Here we find
that when we scale the experimental data
31
by a factor of
0.43, they coincide with our calculated diffusion constants
see dashed curve. Interestingly, the temperature depen-
dence is curved, and this curvature appears to be reproduced
by our calculation, even though the parameters of the MS-
EVB model
20
were adjusted at 300 K.
FIG. 2. Temperature dependence of the proton diffusion coefficient in water
in units of Å
2
/ps) as calculated from classical MS-EVB2 trajectories, with
their corresponding error bars indicated. The dashed line represents a fit to
the experimental data Ref. 31, multiplied by 0.43. The slope of the straight
dotted line gives the activation energy of 2.70.1 kcal/mol.
014506-3 Hydrated proton mobility in water J. Chem. Phys. 122, 014506 (2005)
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
A straight line through the calculated points around room
temperature give an Arrhenius activation energy of 2.7 kcal/
mol, slightly larger than the experimental value of about 2.5
kcal/mol.
27
With the inclusion of quantum effects, the calcu-
lated value is expected to decrease by up to 0.4 kcal/mol see
Fig. 10 in Ref. 17.
It is interesting that a simplified EVB implementation
gets a similar activation energy.
22
This could indicate, as ar-
gued below, that the activation energy for proton mobility
reflects the strength of the HB between bulk water
molecules.
1
As such, the underlying water potential may be
the most crucial element in determining an accurate value for
the activation energy. It should be noted, however, that other
important quantities, such as the actual value of D
H
i.e., its
pre-exponential factor as well as the binding and spectro-
scopic properties of the excess proton, are likely to be more
sensitive to the overall physical accuracy of the model.
IV. BOND-ORDER ANALYSIS
We introduce bond-order BOanalysis in order to quan-
tify the HB environment around each oxygen atom. The BO
provides a ‘gray scale’ description of HBs, which replaces
their conventional all-or-none definition in terms of cutoff
distances and angles. In addition, it allows us to sum the
contribution from several HBs within a solvation shell, gen-
erating a small number of parameters which we use to de-
scribe the PT process.
Following Pauling,
32
the BO n is related exponentially
to the bond length r,
nexp
rr
eq
/a
, 1
where r
eq
is ‘the’ equilibrium bond length, which we take
as the value for OH in gas-phase water, 0.956 Å. The param-
eter a, according to Ref. 33, is 0.35 Å the exact value is
immaterial for the qualitative analysis described herein. The
Pauling BO varies smoothly between covalent and hydrogen
bonds, with stronger bonds having larger BOs. Typical val-
ues are given in Table I. It has been observed that in adjacent
covalent/hydrogen bond pairs, OH¯ O, the total BO is
conserved.
33–36
Here we generalize the definition of BO to give the total
effective BO TEBO m around an oxygen center. Since we
are interested in protonated water clusters, such as shown in
Fig. 1, we view the HBs as emanating from the protonated
center. As we ‘walk’ out from this center along the HB
network, HBs that stabilize it are directed from hydrogen to
oxygen. Consider, for example, the hydrogen atom H
*
in
Fig. 3, which is hydrogen bonded to oxygen O
*
in the water
molecule H
2
O
*
. We wish to characterize the effective coor-
dination number of this H
2
O
*
due to all other HBs i.e.,
excluding H
*
¯ O
*
H
2
).
Typically there are up to three such bonds, two of which
are donated by the hydrogens of H
2
O
*
their BOs are de-
noted n
1
and n
2
), whereas the third denoted n
3
) is accepted
by O
*
. From the perspective of transferring the proton H
*
to
the nearest oxygen atom O
*
, n
1
and n
2
represent favorable
interactions, which stabilize the transferring proton. In con-
trast, n
3
is an unfavorable, destabilizing interaction. The cu-
mulative effect of these three HBs is thus depicted by the
weighted sum
mn
1
n
2
n
3
, 2
in which n
3
receives a negative weight. Note that if H
*
is
positively charged, a bond of type n
3
is less likely to exist.
However, n
3
plays an increasing role as one moves further
away from the positively charged center. Thus, the magni-
tude of m describes how receptive the HB environment
around an oxygen atom is toward accepting a proton. We use
m to characterize the first and second shells around a trans-
ferring proton as follows.
Suppose we ‘sit’’ on the transferring proton as depicted
in Fig. 1. This proton is flanked by two water molecules, to
its left, (H
2
O)
l
, and to its right, (H
2
O)
r
. The BOs of the
proton to the two corresponding oxygen atoms are denoted
by n
l
and n
r
, respectively Fig. 1a兲兴. These constitute the
PT coordinates. The first solvation shell of the central H
5
O
2
is subsequently characterized by the total effective BOs, m
1l
TABLE I. Typical bond orders in water and protonated water. Covalent
bonds are denoted by full lines whereas HBs are dotted.
Bond BO
Gas-phase OH 1
Liquid-phase OH 0.9
H
2
O–H
0.8
H
2
O¯ H
¯ OH
2
0.55
H
2
OH
¯ O 0.3
H
2
O¯ H
2
O 0.15
FIG. 3. ColorThe three HBs that enter into the definition of the TEBO in
Eq. 2. n
1
and n
2
, which are donated from the central water molecule, are
depicted in blue. n
3
, which is accepted at the central molecule, is depicted
in red. When H
*
is positively charged, blue and red correspond to favorable
and unfavorable interactions, respectively.
014506-4 Lapid
et al.
J. Chem. Phys. 122, 014506 (2005)
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
and m
1r
, respectively Fig. 1b兲兴. These are defined as in Eq.
2. We note that even this first-shell cluster, depicted in Fig.
1b, is larger than previously considered
1,8
because we con-
sider explicitly the dynamics of the two HBs donated by
(H
2
O)
r
and not only of the one accepted by it.
Moving on to the second solvation shell, let us consider
the oxygen atoms l
and l
to which O
l
donates HBs. They
are characterized by total effective BOs m
l
and m
l
, respec-
tively. Similarly on the acceptor right side, see Fig. 1c.
Thus the average m values for the second shell on the left
and right sides are defined by
m
2l
m
l
m
l
/2,
3
m
2r
m
r
m
r
/2.
The 20 HBs generated by our tracking routines have conse-
quently been concatenated into six parameters. These are n
l
,
m
1l
, and m
2l
on the left side of the transferring proton, and
analogously n
r
, m
1r
, and m
2r
on its right. We shall monitor
these parameters during PT events in the MS-EVB simula-
tions.
It is also helpful to consider the differences between the
left and right TEBOs,
m
i
i
m
il
m
ir
, 4
where
i
is a scaling factor which puts the different layers on
the same scale. One may expect to observe PT when m
i
0, which we verify below.
V. ANALYSIS OF PROTON-TRANSFER EVENTS
Analysis of proton-hopping events was performed at a
single temperature 280 K. While the mechanism of proton
mobility is not expected to change much from 300 K, re-
duced HB fluctuations may make it easier to detect. As Fig.
2 indicates, the MS-EVB model should be applicable over a
whole temperature range around room temperature including
280 K.
We focus first on the PT event within the protonated
water dimer H
5
O
2
with somewhat more detail than previ-
ously presented. Consequently, we utilize the TEBO vari-
ables to demonstrate how PT within this complex correlates
with the dynamics in the first two solvation shells. All of our
25 PT events are collected as supplementary material,
37
from
which only four are utilized in demonstrating the results be-
low. Because our m
i
values are already averages of 3i HBs
each, the examples presented in the sequel are indeed char-
acteristic.
A. Proton transfer within the central H
5
O
2
¿
moiety
We have monitored the five OH distances within the
central H
5
O
2
complex for each PT event. These distances
are depicted in Fig. 4. r
1
and r
2
are the PT coordinates
within this complex, whereas the other four are the covalent
OH bonds of the two participating water molecules.
Figure 5 shows one of the observed PT events trajectory
7a, denoted T7a. It is clearly indicative of an initial H
3
O
cation which is converted into another via an intermediate
H
5
O
2
. Initially, r
1
in panel a, and r
3
and r
4
in panel b
assume an average value of 1.055 Å, which is typical to an
FIG. 4. ColorFive OH distances within the H
5
O
2
complex. Color codes
correspond to Fig. 5 below.
FIG. 5. Color Proton-transfer dynamics within the H
5
O
2
complex in one
sample trajectory PT event T7a. The five distances shown correspond to
those depicted in Fig. 4. Dotted horizontal arrows mark typical bond lengths
within the H
5
O
2
complex, whereas the vertical dashed lines delimit its
existence epoch.
014506-5 Hydrated proton mobility in water J. Chem. Phys. 122, 014506 (2005)
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
OH bond within H
3
O
. At the same time, r
2
is a HB formed
between H
3
O
and water, having a typical value of 1.6 Å.
The two OH bond lengths of the acceptor water molecule
are then around 0.98 Å, as seen in panel c. Due to water-
water interactions in the condensed phase, this value is
slightly larger than the gas-phase bond length of 0.96 Å.
At about 1 ps, a rather tight H
5
O
2
complex is formed,
with both r
1
and r
2
fluctuating around 1.2 Å dotted arrow in
panel a兲兴. Interestingly, the four covalent bonds also assume
an intermediate value, around 1.015 Å dotted arrows in pan-
els b and c兲兴. Thus all the OH bonds give testimony to
the formation of the complex. The complex attempts to sepa-
rate several times, but succeeds in doing so only at about 3
ps. It is thus rather long lived ca.2ps. This lifetime varies
from event to event, see Sec. V B below.
In the second half of the transfer event, the complex
dissociates to form the product H
3
O
. Then r
2
becomes a
covalent bond whereas r
1
is converted into a HB. At the
same time, r
3
and r
4
further reduce to the characteristic wa-
ter value of 0.98 Å, whereas r
4
and r
5
increase to around
1.05 Å. The PT act is then completed. As previously
suggested,
1
it can indeed be regarded as isomerization from a
donor H
3
O
, via an intermediate H
5
O
2
, to the acceptor
H
3
O
.
We mention that in the 25 PT events monitored in this
study there was no indication of a concerted double-proton
transfer event that converts a donor H
5
O
2
directly into an
acceptor H
5
O
2
, as suggested in some MD work.
2,7,14,23
As
discussed in the Introduction, this may depend on the poten-
tial used in the simulations. For MS-EVB2 potential utilized
here,
20
particularly when the trajectories are classical, H
3
O
is more stable hence longer living than H
5
O
2
, and this
makes a double PT event less probable. Nevertheless, we do
observe that after PT the OH bonds to the product H
3
O
remain excited and may participate in further transfer at-
tempts see, for example, the jump in r
5
in Fig. 5c,at4.2
ps. At least for the relatively low temperature considered
here 280 K, such secondary PT attempts occur only well
after the main PT event has terminated.
Finally, it is interesting to consider the amplitude of the
vibrations in the various bonds as time proceeds. The general
anticipation is that the longer bonds are weaker and hence
fluctuate more readily. Panel a shows very large fluctua-
tions in the ‘soft’ HBs, which become considerably more
restricted as a HB is converted into a ‘rigid’ covalent bond.
A similar trend is seen in the covalent bonds. For example,
the OH bonds in the donor H
3
O
molecule, panel b, fluc-
tuate wildly. Their fluctuations become more tamed at long
times, after the PT event. Possibly, this may also be due to
the fact that the pyramidal H
3
O
disturbs the tetrahedral
water structure around it. This disturbance is mostly allevi-
ated once it is converted into a first-shell water molecule.
B. Lifetime distribution of the protonated dimer
We can use the (n
l
,n
r
) data collected in the supplemen-
tary material
37
upper panelsto compute a lifetime distribu-
tion for the H
5
O
2
complex. This complex was assumed to
exist between the first and last time that n
l
n
r
. The 25
lifetimes
thus determined average to
375 fs. They
were binned into five equal intervals between 01 ps, as
shown in Fig. 6. Two trajectories had
1 ps and three had
0. In the latter case H
5
O
2
is better described as a transi-
tion state of a direct reaction.
Although the statistics generated by a small number of
events is not particularly good, the lifetime distribution p(
)
does seem to obey an exponential law,
p
A exp
/
, 5
and the best fit gives
367 fs, very close to its numerical
value. Exponential lifetime distribution is what one expects
from first order kinetics A B, where A and B are the two
forms of protonated water. The average lifetime of the
H
5
O
2
, about 370 fs, is around one order of magnitude
shorter than that of the H
3
O
cation, which is in accordance
with their energy difference ca. 1 kcal/mol Ref. 20兲兴. This
ratio may diminish if the nuclear coordinates are propagated
by quantal rather than classical MD.
C. Correlated dynamics within the solvation shells
The main mechanistic result of the present study is the
correlation between the PT dynamics observed in Fig. 5, and
the HB dynamics in the first- and second-solvation shells of
the protonated dimer. These are best depicted by the TEBO
parameters m
1
for first shell and m
2
for second shell de-
fined above. In order to eliminate the fast hydrogen vibra-
tions, these curves were smoothed using a three-point mov-
ing average filter, through which the data were run three
consecutive times. What is left is the ‘backbone’ oxygen
fluctuations, so that the crossing of corresponding ‘left’ and
‘right’’ curves is a better indication of an attempted transfer
event. Results are presented as supplementary material for all
the 25 trajectories,
37
three of which are discussed below.
Corresponding to the three panels in Fig. 1, we show in
the three panels of Fig. 7 the smoothed TEBO parameters for
FIG. 6. Lifetime distribution for the H
5
O
2
complex histogram, with the
exponential fit line of Eq. 5.
014506-6 Lapid
et al.
J. Chem. Phys. 122, 014506 (2005)
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
the PT coordinates and the first-two solvation layers, for the
trajectory denoted T9b2. Panel a shows time evolution of
the PT coordinates. This is the same as Fig. 5a in terms of
BOs but smoothed and for a different PT event. The for-
mation and cleavage of the protonated dimer is identified as
the first and last crossing of the two curves. The correspond-
ing times are indicated by the dashed vertical lines. During
the epoch between these lines, there are periods of a tightly
bound dimer the lines depicting n
l
and n
r
lie close together
and other periods of wide fluctuations which may almost be
considered as involving an intervening backward PT step.
This constitutes a ‘fluxional complex,’
8
which samples
many substates during its lifetime.
The new feature revealed by Fig. 7 is the correlated mo-
tion of the different hydration layers. The figure shows that
an analogous behavior to that in panel ais observed for the
m
i
values of layers 1 and 2. Hence one concludes that PT
occurs when all water layers respond in concert. As Ohmine
and collaborators have observed in simulated water
dynamics,
29,30
processes in water are driven by large-scale
collective motions. In this paper a cooperative behavior is
demonstrated for proton mobility in water.
A more careful inspection reveals that the curves for the
first layer in panel b approach and separate more slowly
than the PT coordinates enter or exit their first and last cross-
ings in panel a, respectively. In comparison, the second
layer in panel cshows an even more sluggish response, the
two lines appearing to ‘stick together beyond the epoch
defining the protonated dimer complex. In a sense, the sur-
rounding solvent is ‘‘preparing itself to the PT reaction first,
much as Marcus has envisioned the occurrence of a charge-
transfer reaction. As Onsager once commented, reactive
events in water take place from the outside in, as an ‘in-
verted snow ball’second shell first, inner core last.
Additional insight on how the different water-layers par-
ticipate in the PT event may be gleaned by overlaying the
bottom two panels in Fig. 7. Maintaining the same color
code, Fig. 8 shows that before the complex is formed the two
green curves roughly coincide, m
2l
m
1r
encircled. This
corresponds to a symmetric solvent environment around the
donor H
3
O
. After the complex disintegrates, the two blue
curves roughly coincide, m
1l
m
2r
encircled. This indi-
cates the formation of a symmetric solvent environment
around the acceptor H
3
O
moiety. The following relation
donor: m
2l
m
1r
,
dimer: m
1l
m
1r
, m
2l
m
2r
, 6
acceptor: m
1l
m
2r
,
summarizes our observation. It is best appreciated with ref-
erence to Fig. 1.
The isomerization times namely, the time to actually
convert H
3
O
into H
5
O
2
or vice versa can be estimated
qualitatively from Fig. 8. It is seen that the donor environ-
ment loses its threefold symmetry about 50100 fs before
the protonated H
5
O
2
is formed circle. Similarly, the accep-
tor environment gains its threefold symmetry at about 50
100 fs after the H
5
O
2
dissociates circle. These times are
appreciably faster than the 370 fs dimer lifetime or the few
FIG. 7. Color Proton-transfer dynamics correlates with the HB dynamics
within the first two solvation layers surrounding the H
5
O
2
complex see
Fig. 1 for definitions and color codes. Panel a depicts PT event T9b2,
which is analogous to PT event T7a in Fig. 5a, only in BO coordinates.
The first and last crossings of n
l
and n
r
delimit the existence of the complex
vertical dashed lines. The zero of time is set at the middle of this interval.
The two BO parameters panel a兲兴 and four TEBO parameters panels b
and c兲兴 have been smoothed to eliminate fast hydrogen atom vibrations.
FIG. 8. ColorOverlay of the bottom two panels in Fig. 7. Bold and dashed
lines correspond to first and second layers, respectively. See text for further
discussion.
014506-7 Hydrated proton mobility in water J. Chem. Phys. 122, 014506 (2005)
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
picoseconds which elapse between proton hops. Thus the
rate limiting process is the concerted reorganization of the
HB environment as revealed by the TEBO analysis.
To see that the above conclusions are not trajectory spe-
cific, we present the TEBO parameters along two other tra-
jectories. Figure 9 event T10c presents an extreme case of
a single crossing event at t 0), corresponding to a ‘‘direct’
H
3
O
to H
3
O
transition without a long-lived H
5
O
2
inter-
mediate. Figure 10 event T13ais one of the ‘worst’ cases
in our collection in terms of demonstrating the correlation
between the solvation layers. It is similar to Fig. 9 in exhib-
iting a direct transition, but the curves depicting m
2l
and m
2r
fail to separate after this transition. A closer inspection of the
upper panel reveals that this is probably connected with re-
peated nonreactive re-encounters, when n
l
and n
r
in the
upper panel nearly touch.
The two cases are compared in Fig. 11, which depicts the
difference m, between the two curves in each panel, see
Eq. 4. More precisely, we show m
0
n
l
n
r
inner core,
black, m
1
1.5(m
1l
m
1r
) first layer, red, and m
2
3(m
2l
m
2r
) second layer, green. The scaling factors
1.5 and 3 were applied to put the data from all layers on a
similar scale. PT is represented by a crossing of the m
0 line.
In both cases, the first layer follows the inner core very
closely: even the backbone fluctuations in m
1
and m
0
are
nearly identical. Thus the correlation between these two is so
strong that their response is essentially in concert. The cor-
relation with the second layer is only in the average trend,
having m
2
0 before the PT, decreasing to m
2
0 after-
ward. This seems to hold on the average even in the worst
case of T13a. When m
2
lingers around 0, we obtain re-
peated transfer events or a period with a tight H
5
O
2
com-
plex.
Figure 12 shows the average of m
i
over our 25 trajec-
tory data bases. There is some arbitrariness in averaging over
different trajectories. For example, the result depends on the
choice of t 0 for each trajectory. Here we maintain the
assignment of t 0 at the middle of each interval defining
the H
5
O
2
complex. The resulting curves in this figure dem-
onstrate that the collective behavior of the core and first two
layers holds also on the average. Again the PT event is seen
to correlate more strongly with the dynamics in the first layer
than with the second the curves for i 0 and 1 nearly over-
lap. On the other hand, the averaging procedure has com-
pletely obliterated the H
5
O
2
intermediate. Since in each tra-
jectory it lives for a different time duration
, the flat step at
m
i
0 has been replaced by a gradual slope. The resulting
apparently slow (1 ps) conversion of reactant to product
H
3
O
conceals the much faster (100 fs) interconversion
between H
3
O
and H
5
O
2
which is driven by the slower
reorganization of the environment. The figure thus demon-
FIG. 9. ColorSame as Fig. 7 for PT event T10c, which shows a very short
H
5
O
2
lifetime.
FIG. 10. Color Same as Fig. 7 for PT event T13a, which shows some of
the worst correlations with second-shell dynamics.
014506-8 Lapid
et al.
J. Chem. Phys. 122, 014506 (2005)
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
strates both the utility and drawback of applying averaging
procedures in mechanistic studies.
D. Contribution from individual hydrogen bonds
The above analysis focused on the average BO contribu-
tion to the first two solvation shells surrounding the H
5
O
2
intermediate. It is instructive to bisect this into typical con-
tributions from the different types of HBs in the two hydra-
tion layers. Doing so, we lose the self-averaging property of
the m
i
so that the results discussed below show much larger
variability from one trajectory to another.
The behavior of the first layer, Fig. 1b, resembles to
some extent the behavior observed for PT in ice.
25
The
‘good’ HBs (n
1
and n
2
, see Fig. 3 seldom break. As the
proton migrates from left to right, they expand on the left
and shrink on the right.
1
This leads to the observed decrease
of m
1l
predominantly before PT and the increase in m
1r
afterwards. In ice, the two ‘bad’’ HBs of type n
3
) show the
opposite trend: The one on the left lengthens whereas that on
the right contracts.
25
In liquid water, due to the delocalization
of the positive charge, these two bonds are not frequently
observed.
20
When they do exist, we often find that the one on
the right cleaves before the transfer and the one on the left
forms afterwards, as suggested in Ref. 1.
In the second layer, one may break down m
2
into the
contribution from n
1
n
2
and n
3
. Figure 13 shows the result
for event T10c on the donor side for each of the two water
molecules involved (l
and l
in Fig. 1. It can be seen that
before the PT event vertical dashed line, one or more of the
‘good’ HBs cleave, reducing n
1
n
2
abruptly blue line.
After the PT event, the ‘negative’ HB, n
3
becomes active
red line. A mirror image of this scenario often holds on the
acceptor side.
FIG. 11. ColorTEBO parameter difference for the transferring proton and
the two solvation layers, for the trajectories shown in Figs. 9 top and 10
bottom. These represent favorable and unfavorable cases respectivelyin
terms of the correlation with the second shell.
FIG. 12. ColorAveraged TEBO difference for the core and two solvation
layers. Same as Fig. 11, only averaged over the 25 trajectories in the supple-
mentary material Ref. 37. For the second shell green, a scaling of
2
4.5 was used.
FIG. 13. ColorThe contribution from individual HBs to m
2l
for PT event
T10c see Fig. 9c兲兴. The TEBO m
2l
is half the sum of the four lines in this
figure.
014506-9 Hydrated proton mobility in water J. Chem. Phys. 122, 014506 (2005)
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
The contribution of n
3
to m
2
is very much in line with
the observations of Day et al.
19,20
By monitoring the HO¯ H
angle of the red HBs in Fig. 1c, they have observed that the
ones on the right break, on the average, concomitant with the
PT event. Thus the HB cleavage event suggested in Ref. 1 as
a rate limiting step for proton mobility does occur, only one
water molecule further away from the protonated center.
What was not previously anticipated is the contribution
of the good HBs, n
1
and n
2
. These, on the average, tend to
cleave on the donor side left and reform on the acceptor
side right. Evidently, their behavior is just the opposite of
that of n
3
. Consequently, the H
3
O
first-shell coordination
numbers fluctuate over the larger range of 14, rather than
just between 3 and 4 as assumed before. Clearly, not all these
individual HB cleavage and formation events occur in each
trajectory. What is required for PT is just enough of them to
tilt the balance from the donor to the acceptor side.
E. The extended picture of proton mobility
in liquid water
From the above discussion, an extended picture of pro-
ton mobility emerges. A schematic summary of HB rear-
rangements coupled to the proton hopping act is given in Fig.
14. HBs break and form predominantly within the second
hydration shell curly arrows, whereas in the first shell they
typically only extend or contract straight arrows, therefore
showing much stronger correlation with the inner core. The
rate limiting step is likely to be the HB cleavage events
which occur within the second shell. This conclusion is
qualitatively as suggested in Refs. 1 and 7, except that it is
not possible to implicate one single HB as the key player in
the mechanism.
The two types of HBs show opposite behaviors within
the second shell. Prior to PT Fig. 14a兲兴, good HBs cleave
on the donor side and bad HBs cleave on the acceptor side. A
sufficient number of these bonds should remain simulta-
neously broken in order to break the threefold symmetry
around the donor H
3
O
. The cleavage events themselves are
fast say, 50150 fs, so that they occur consecutively rather
than simultaneously. The last of these finally tilts the balance
from reactants to products. This explains why the activation
energy for proton mobility Sec. III is so close to the HB
strength in liquid water 2.6 kcal/mol,
38
although it is a col-
lective breaking of HBs and not the cleavage of a single HB
that drives the PT.
Finally, following the PT event, good HBs form on the
acceptor side, whereas the bad ones reform on the donor side
see Fig. 14b兲兴. This terminates the fluctuations of the pro-
ton within the H
5
O
2
complex. It localizes the proton on the
acceptor H
3
O
moiety and establishes a new threefold sym-
metry around it. The temporal division may be less sharp
than depicted, as some of the HB dynamics occurs also dur-
ing the lifetime of the complex.
VI. CONCLUSIONS
The elucidation of the mechanism of proton mobility in
water is a basic problem of physical chemistry,
9
with far
reaching consequences, for example, in fuel-cell technology
and biology. Previous discussions were based on the combi-
nation of experiment and chemical intuition,
1
on qualitative
visualization of short trajectories,
7,8
or focused on a single
HB.
2,19,20
A method for systematic analysis of the coopera-
tive behavior of many HBs was lacking. The present work
introduced BOA, a method which is both intuitive and easy
to implement.
FIG. 14. Color HB dynamics couples to proton mobility in water. Before
PT, panel a, HBs mainly break second shell, curly orange arrows or
expand first shell, straight orange arrows. Good HBs may break on the
donor side leftwhereas bad ones break on the acceptor side right.Around
the donor H
3
O
, two bonds expand, whereas the central OO bond con-
tracts to form the H
5
O
2
complex not shown. As the complex disintegrates
to form the product H
3
O
, panel b, this central bond expands, whereas the
two other HBs in the first shell contract straight green arrows. HBs now
form in the second solvation shell curly green arrows, predominantly the
good ones on the acceptor side and the bad ones on the donor side. Some
bonds may have reformed during the lifetime of the H
5
O
2
intermediate
bottom left.
014506-10 Lapid
et al.
J. Chem. Phys. 122, 014506 (2005)
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
Utilizing BOA, we could track the cooperative behavior
of the 18 HBs surrounding the transferring H
5
O
2
complex.
We find that this complex is almost always an intermediate
separating the donor and acceptor H
3
O
moieties. The tran-
sition between these structures indeed couples to HB dynam-
ics, but within considerably larger water clusters than previ-
ously anticipated.
As PT progresses, the TEBO diminishes on the donor
side and increases on the acceptor side. This appears to occur
in concert within the first and second solvation shells. In the
first shell one observes predominantly stretching and con-
traction of HBs, and the correlation with the PT act is very
strong. In the second shell there are more HB cleavage and
formation events, and the correlation is only on average. One
may nevertheless conjecture that the ‘rate limiting step’ lies
in the second shell. The collective accumulation of several
consecutive HB cleavage events there eventually tilts the
balance from one form of protonated water to the other.
The TEBO is constructed by dividing the HBs into two
types: The blue bonds emanate from the protonated center
and thus stabilize it, whereas the red bonds which are di-
rected toward the protonated center destabilize it. The anal-
ogy with Moses parting the Red Sea
1
is now two colored:
The red sea parts in front of the proton and closes behind its
back, whereas the blue sea parts in his rear and forms up
front.
The above conclusions may depend on several aspects of
our simulation methodology. One concern may be the utili-
zation of one particular MS-EVB potential. Judging from the
excellent value obtained for the activation energy of proton
mobility 2.7 kcal/mol at room temperature, this potential
appears to yield quite realistic results. A second concern is
the neglect of quantum effects on the nuclear dynamics. We
anticipate that these could reduce the activation energy by
about 0.3 kcal/mol but, as previously noted,
7
the influence on
the mechanism itself may not be dramatic. An increase in
temperature from the value of 280 K considered herein
could also make a difference in the details of the mechanism.
Given the above reservations, the conclusions in this
work are not expected to be the final word on the proton
mobility mechanism. The classical MS-EVB2 calculations
serve to demonstrate the utility of our TEBO approach. Since
it is easily applied to trajectories of any origin, it may be
interesting to apply this method to different temperatures and
simulation methodologies in the future. Finally, the BOA
may be useful in analyzing proton mobility near and through
biological membranes and channels, topics of great interest
for proton-driven bioenergetics.
6
ACKNOWLEDGMENTS
This research was supported by Grant No. 98-00083
from the United States-Israel Binational Science Foundation
BSF, Jerusalem, Israel. G.A.V. acknowledges support from
the United States National Science Foundation NSF Grant
No. CHE-0317132.
1
N. Agmon, Chem. Phys. Lett. 244, 456 1995.
2
R. Vuilleumier and D. Borgis, Isr. J. Chem. 39,4571999.
3
D. Zahn and J. Brickmann, Isr. J. Chem. 39, 469 1999.
4
U. W. Schmitt and G. A. Voth, Isr. J. Chem. 39, 483 1999.
5
N. Agmon, Isr. J. Chem. 39, 493 1999.
6
T. E. Decoursey, Physiol. Rev. 83, 475 2003.
7
M. Tuckerman, K. Laasonen, M. Sprik, and M. Parrinello, J. Phys. Chem.
99, 5749 1995.
8
D. Marx, M. E. Tuckerman, J. Hutter, and M. Parrinello, Nature London
397, 601 1999.
9
P. W. Atkins, Physical Chemistry, 6th ed. W. H. Freeman, New York,
1998.
10
A. Warshel and R. M. Weiss, J. Am. Chem. Soc. 102,62181980.
11
A. Warshel, J. Phys. Chem. 86, 2218 1982.
12
J. Lobaugh and G. A. Voth, J. Chem. Phys. 104, 2056 1996.
13
R. Vuilleumier and D. Borgis, J. Mol. Struct. 436437,5551997.
14
R. Vuilleumier and D. Borgis, J. Phys. Chem. B 102, 4261 1998.
15
R. Vuilleumier and D. Borgis, J. Chem. Phys. 111, 4251 1999.
16
U. W. Schmitt and G. A. Voth, J. Phys. Chem. B 102, 5547 1998.
17
U. W. Schmitt and G. A. Voth, J. Chem. Phys. 111, 9361 1999.
18
U. W. Schmitt and G. A. Voth, Chem. Phys. Lett. 329,362000.
19
T. J. F. Day, U. W. Schmitt, and G. A. Voth, J. Am. Chem. Soc. 122, 12027
2000.
20
T. J. F. Day, A. V. Soudackov, M. C
ˇ
uma, U. W. Schmitt, and G. A. Voth,
J. Chem. Phys. 117, 5839 2002.
21
M. A. Lill and V. Helms, J. Chem. Phys. 115,79932001.
22
S. Walbran and A. A. Kornyshev, J. Chem. Phys. 114, 10039 2001.
23
A. A. Kornyshev, A. M. Kuznetsov, E. Spohr, and J. Ulstrup, J. Phys.
Chem. B 107,33512003.
24
J. C. Grossman, E. Schwegler, E. W. Draeger, F. Gygi, and G. Galli, J.
Chem. Phys. 120,3002004.
25
C. Kobayashi, S. Saito, and I. Ohmine, J. Chem. Phys. 113, 9090 2000.
26
A. W. Omta, M. F. Kropman, S. Woutersen, and H. J. Bakker, Science
301, 301 2003.
27
N. Agmon, J. Chim. Phys. Phys.-Chim. Biol. 93, 1714 1996.
28
It should be noted that the simplified EVB potentials Ref. 22 and phe-
nomenological stochastic proton hopping algorithms Ref. 21do not fully
include this charge delocalization effect, so their underlying HB dynamics
will be different.
29
I. Ohmine, J. Phys. Chem. 99, 6767 1995.
30
I. Ohmine and S. Saito, Acc. Chem. Res. 32, 741 1999.
31
B. D. Cornish and R. J. Speedy, J. Phys. Chem. 88, 1888 1984.
32
L. Pauling, J. Am. Chem. Soc. 69,5421947.
33
N. Agmon, J. Mol. Liq. 73Õ74, 513 1997.
34
N. Agmon, Chem. Phys. Lett. 45,3431977.
35
H. Alig, J. Lo
¨
sel, and M. Tro
¨
mel, Z. Kristallogr. 209,181994.
36
T. Steiner and W. Saenger, Acta Crystallogr., Sect. B: Struct. Sci. 50,348
1994.
37
See EPAPS Document No. E-JCPSA6-121-510446 for figures depicting
all of the 25 PT events in a format similar to Fig. 7. A direct link to this
document may be found in the online article’s HTML reference section.
The document may also be reached via the EPAPS homepage http://
www.aip.org/pubservs/epaps.html or from ftp.aip.org in the directory
/epaps/. See the EPAPS homepage for more information.
38
E. W. Castner, Jr., Y. J. Chang, Y. C. Chu, and G. E. Walrafen, J. Chem.
Phys. 102, 653 1995.
014506-11 Hydrated proton mobility in water J. Chem. Phys. 122, 014506 (2005)
Downloaded 16 Dec 2004 to 132.64.1.37. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
... 28 A lot of our modern understanding of the proton structure in water has emerged from several decades of atomistic molecular dynamics simulations using both first-principles Density Functional Theory (DFT), [29][30][31][32][33][34][35] as well as reactive potentials such as empirical valence bond. [36][37][38][39][40][41] Quantum chemistry-type calculations in the gas phase and in cluster environments, which typically ignore thermal and entropic effects, have also played an important role in understanding the structure of the proton and how it is affected by hydration water. [42][43][44][45][46][47][48][49] Different conclusions have been reached regarding the dominant structure of the proton including, besides the canonical Eigen and Zundel complexes, descriptions such as the distorted Eigen cation 50 and the distorted Zundel structure. ...
... 54,55 Since the identity of the proton changes as it migrates through the HBN, various types of chemically-inspired criteria need to be used to identify where the proton is instantaneously localized. 33,34,38 Within this context, chemical bias requires some form of dimensionality reduction which can lead to an uncontrolled information loss when examining fluctuations of the system in a reduced space of coordinates. In the last decade there has been a tremendous spurt in the use of data-science techniques to study complex molecular systems. ...
... The picture that emerges from the current analysis is in good agreement with previous characterizations of the PT mechanisms associated with the special-pair dance, where the excess proton dynamically fluctuates between forming H-bonds between three different atomic environments. 38 Finally, Figure 6d ...
Preprint
Full-text available
The structure of the excess proton in liquid water has been the subject of lively debate from both experimental and theoretical fronts for the last century. Fluctuations of the proton are typically interpreted in terms of limiting states referred to as the Eigen and Zundel species. Here we put these ideas under the microscope taking advantage of recent advances in unsupervised learning that use local atomic descriptors to characterize environments of acidic water combined with advanced clustering techniques. Our agnostic approach leads to the observation of only a single charged cluster and two neutral ones. We demonstrate that the charged cluster involving the excess proton, is best seen as an ionic topological defect in water's hydrogen bond network forming a single local minimum on the global free-energy landscape. This charged defect is a highly fluxional moiety where the idealized Eigen and Zundel species are neither limiting configurations nor distinct thermodynamic states. Instead, the ionic defect enhances the presence of neutral water defects through strong interactions with the network. We dub the combination of the charged and neutral defect clusters as ZundEig demonstrating that the fluctuations between these local environments provide a general framework for rationalizing more descriptive notions of the proton in the existing literature.
... While proton transfer and proton transport occur in a variety of environments, from solutions to membrane proteins and fuel-cell membranes, the protonated water hexamer is one of the smallest clusters to incorporate most of the PT experimental features and solvation effects at the leading order. According to ref. 50, one more hydration layer is needed to reach the water bulk limit. From this viewpoint, the hexamer is close to that limit, and some relevant effects, emerging already at this size, can be transferred to larger systems. ...
Article
Full-text available
Water is a key ingredient for life and plays a central role as solvent in many biochemical reactions. However, the intrinsically quantum nature of the hydrogen nucleus, revealing itself in a large variety of physical manifestations, including proton transfer, gives rise to unexpected phenomena whose description is still elusive. Here we study, by a combination of state-of-the-art quantum Monte Carlo methods and path-integral molecular dynamics, the structure and hydrogen-bond dynamics of the protonated water hexamer, the fundamental unit for the hydrated proton. We report a remarkably low thermal expansion of the hydrogen bond from zero temperature up to 300 K, owing to the presence of short-Zundel configurations, characterised by proton delocalisation and favoured by the synergy of nuclear quantum effects and thermal activation. The hydrogen bond strength progressively weakens above 300 K, when localised Eigen-like configurations become relevant. Our analysis, supported by the instanton statistics of shuttling protons, reveals that the near-room-temperature range from 250 K to 300 K is optimal for proton transfer in the protonated water hexamer.
... As largely discussed in the article concerning the relationship between raised conductivity and dynamic order in water, the correlation of proton hopping over long hydrogen-bonded water chains may arise [64]. Thus, Lapid and colleagues [65] propose that proton mobility is characterized by a cooperative phenomenon, resulting in facilitated proton conduction when water molecules are more ordered [66,67]. The higher degree of orderliness of water is also confirmed through theoretical estimates by various authors based on quantum electrodynamic theory (QED). ...
Article
Full-text available
Physicochemical investigations of (UHD) solutions subjected to certain physical factors (like shaking) are becoming more frequent and increasingly yielding convincing results. A much less studied phenomenon is the transfer of molecular information (UHD signals) from one fluid to another without an intermediate liquid phase. The purpose of this study was to investigate the possibility of such a UHD signal transfer from UHD solutions into the receiver fluid, especially when the molecular source used in solutions was a biologically active molecule of antibodies to interferon-gamma. We used physicochemical measurements and UV spectroscopy for this purpose. The results of this large pilot study confirm the possibility of such a transfer and a rough similarity to the original UHD signal donors, the weaker signal detection relative to the original donor fluids, and that exposure time improves the effect.
Article
The proton transport in one-dimensional (1D) confined water chains has been extensively studied as a model for ion channels in cell membrane and fuel cell. However, the mechanistic understanding of the proton transfer (PT) process in 1D water chains remains incomplete. In this study, we demonstrate that the two limiting structures of the hydrated excess proton, H5O2+ (Zundel) and H3O+ (linear H7O3+), undergo a change in dominance as the water chain grows, causing two co-existing and opposing PT mechanisms. Specifically, H5O2+ is stable in the middle of the chain, whereas H3O+ serves as a transition state (TS). Except for this region, H3O+ is stabilized while H5O2+ serves as a TS. The interaction analysis shows that the electrostatic interaction plays a crucial role in the difference in PT mechanisms. Our work fills a knowledge gap between the various PT mechanisms reported in bulk water and long 1D water chains, contributing to a deeper understanding of biological ion channels at the atomic level.
Preprint
Full-text available
The investigation of cooperative dynamics in H2O, visible in the coherent neutron scattering, has been hindered up to now due to the very small signal. Using a novel approach with employing of neutron polarization analysis we were able for the first time to directly measure the coherent neutron scattering signal in light water with unprecedented accuracy. The observed coherent signal is enhanced at intermediate Q range of 0.2-1 Å -1 which is clear evidence that intermolecular interactions in water spread beyond distances between two nearest neighbours. Our study reveals the existence of a picosecond cooperative process in water, whose nature could be related to the cooperative rearrangements between several water molecules. This process serves as a precursor to the large-scale transport related to the viscosity. Our results help to understand the general transport mechanism at nanoscale which can be useful for biomedical technologies or the development of nanofluidic devices.
Preprint
Full-text available
The design of proton-exchange membranes (PEMs) for high-performance, durable fuel 1 cells and related electrochemical devices requires a delicate balance between high ion-exchange 2 capacity and proton conductivity, while ensuring robust mechanical properties and preserving 3 dimensional, chemical and thermal stability. In addition, low species crossover is desirable to reduce 4 hydrogen peroxide formation. Ionomers used in PEMs can be classified into two main groups: 5 (i) perfluorosulfonic acid (PFSA) polymers, and (ii) aromatic hydrocarbon (HC) polymers. In this 6 work, an analysis of key characteristics of both PEM types is presented, including water uptake, 7 proton conductivity, water transport properties, thermal conductivity, permeability, mechanical 8 properties and chemical and thermal stability, among others. Comparatively, PFSA-based PEMs 9 are undoubtedly the commercial standard due to its proven high proton conductivity and good 10 chemical stability, even though PTFE-reinforced aromatic HC-based PEMs have also started to be 11 commercialized recently. In the last decades, a growing trend is identified toward the development 12 of hybrid and composite ultra-thin PEMs (5 − 20 µm in thickness) with tailored properties, e.g., 13 incorporating microporous fillers to enhance water uptake and layered reinforced microstructures to 14 improve mechanical properties and chemical stability. PEM design is to be accomplished within an 15 environmentally friendly circular economy with facilitated recycling and re-utilization. 16
Article
Full-text available
Proton mobility is traditionally thought to be governed by water molecule rotation. Water rotation times from D2O NMR spinlattice relaxation measurements are compared with proton hopping times from mobility data with and without subtraction of the estimated hydrodynamic mobility. In the latter case the two data agree nicely at high temperatures. It is concluded that the hydrodynamic proton mobility is considerably smaller than previously believed because H3O+ is nearly immobilized by extra-strong hydrogen-bonds to first-shell water ligands, estimated to be about 2 kcal/mol stronger than bulk hydrogen-bonds. Water rotation is slower than proton hopping below 20°C and has a hydrodynamic component to its activation energy. Therefore, proton mobility is not governed by water rotation but rather both processes are controlled by hydrogen-bond dynamics. Comparison with hydrogen-bond lifetimes from depolarized light scattering suggests that two consecutive cleavage events of ordinary hydrogen bonds constitute a single proton hop. This agrees with a recent molecular model for the Grotthuss mechanism.
Article
Full-text available
Hydrogen-bond lengths to protonated water monomer and dimer in liquid water are difficult to estimate from quantum chemistry calculations of protonated water clusters due to cluster size effects. It appears that bulk water exhibits shorter hydrogen-bonds as compared with gas-phase hydrates. Here the empirical Pauling-BEBO relation is used to correlate the intermolecular O⋯H bond length with the intramolecular HO bond length at a common H-atom. Comparison of the correlation for clusters and bulk water produces refined estimates for the bond lengths in question.
Article
Measurements of the electrical resistivity of 1, 0.1, and 0.01 M solutions of HCl and KCl in water to-32°C are reported. Values of the proton conductivity λH+ in the HCl solutions are estimated. λH+ is a linear function of temperature in the range -32 to +45°C, λH+ = A(T/Ts - 1), and extrapolates to zero at Ts = 227 K. Implications concerning the structure of water at Ts are discussed.
Article
Proton transport in water is treated within a mixed quantum–classical molecular dynamics scheme. Therein the migration of a positive charge is treated as a two-step process: (i) the displacement of a hydronium ion, followed by (ii) proton transfer between a hydronium ion and an adjacent water molecule. Both the hydronium ion and the water molecules are treated as flexible molecules. In the H5O2+-complex, the proton no longer belongs to a single water molecule, but is delocalized between the two oxygen atoms. The proton transfer is treated quantum mechanically as a function of a transfer coordinate within a single complex. The quantum description is changed to another pair of water molecules after strong proton localization to one of the water molecules in the complex and according to distance criteria with respect to adjacent water molecules. It is demonstrated that the proposed formalism is well suited for an effective simulation of the proton migration in water. In particular, the lifetimes of Eigen complexes and the proton diffusion coefficient (as obtained from simulations) are in very good agreement with corresponding experimental data.
Article
In order to study the microscopic nature of the hydrated proton and its transport mechanism, we have introduced a multistate empirical valence bond model, fitted to ab initio results. This model was applied to the study, at low computational cost, of the structure and dynamics of an excess proton in liquid water. The quantum character of the proton is included by means of an effective parametrization of the model using preliminary path-integral calculations. The mechanism of proton transfer is interpreted as the translocation of a special O–H+–O bond along the hydrogen network, i.e., a series of reactions of the form H5O2+ + H2O ⇌ H2O + H5O2+, rather than H3O+ + H2O → H2O + H3O+ as usually described. The translocation of the special bond can be described as a diffusion process with a jump time of 1 ps. A time-dependent correlation function analysis of the special pair relaxation yields two timescales, 0.3 and 3.5 ps. The first time is attributed to the interconversion between a delocalized (H5O2+-like) and a localized (H9O4+-like) form of the hydrated proton within a given special pair. The second one is the relaxation time of the special pair, including return trajectories. The computed diffusion constant, as well as the isotopic substitution effect, are in good agreement with experiment. The hydration structure around the excess proton is discussed in terms of various radial distribution functions around the water molecules involved in the special pair and those in the first solvation shell. The hydrogen-bond-dynamics which accompanies the translocation process is studied statistically. The “Moses mechanism” proposed by Noam Agmon for proton mobility in water is partially verified by our simulations.
Article
The classical and quantum equilibrium properties of an excess proton in bulk phase water are examined computationally with a special emphasis on the influence of an explicit quantum dynamical treatment of the nuclei on the calculated observables. The potential model used, our recently developed multistate empirical valence bond (MS-EVB) approach is described. The MS-EVB model takes into account the interaction of an exchange charge distribution of the charge-transfer complex with the polar solvent, which qualitatively changes the nature of the solvated complex. The impact and importance of the exchange term on the stability of the solvated H5O2+ (Zundel) cation relative to the H9O4+ (Eigen) cation in the liquid phase is demonstrated. Classical and quantum path-integral molecular dynamics (PIMD) simulations of an excess proton in bulk phase water reveal that quantization of the nuclear degrees of freedom results in an increased stabilization of the solvated Zundel cation relative to the Eigen cation, and that species intermediate between the two are also probable. Quantum effects lead to a significant broadening of the probability distributions used to characterize the two species, and a definite differentiation and sharp characterization of the species connected to the excess proton in liquid water is found to be difficult.
Article
We present the complete intermolecular dynamical spectrum of liquid water, by merging the data sets from femtosecond nonlinear‐optical polarization spectroscopy with the depolarized, Bose–Einstein corrected Raman spectrum to cover the frequency range from 0–1200 cm−1. The impulse response function for liquid water at room temperature is calculated, including all of the intermolecular motions.
Article
The lengthening of the covalent O-H bond in O-H ... O hydrogen bonds is re-examined from high-precision low-temperature neutron diffraction studies of organic molecules (32 crystal structures, 136 hydrogen bonds, T < 130 K, R < 0.06, H atoms on symmetry elements excluded). Accuracies are around or better than 0.002 angstrom. The dependencies of the covalent O-H bond length on the H ... O and O ... O separations are smooth with no sign of discontinuities. For all types of O-H and O-D hydrogen-bond donors and in combination with all types of O acceptors, the 0-H bond length follows the same function of the H...O separation (within the experimental accuracy). For angles at H > 150 , no dependency of the O-H bond length on the hydrogen-bond angle can be detected for constant H ... 0. The bond-valence concept that is established in inorganic chemistry also proves to be a good model for O-H...O hydrogen bonds in organic compounds.
Article
A semiclassical trajectory approach for studies of the dynamics of reactions in polar solvent is developed. This approach focusses on the effect of fluctuations of the solvent dipoles on the rate of crossing between the solute resonance forms. The calculations of the surface crossing probability are based on a time-dependent quantum mechanical treatment which uses the classical time dependence of the energy gap (evaluated along the trajectory path of the solute-solvent system). The practical potential of the method is demonstrated in preliminary simulations of electron-transfer and proton-transfer reactions. The conceptual advantage of the method is demonstrated in exploring the molecular meaning of the activation free energy of electron-transfer reactions, and in a new derivation of the rate constant of electron-transfer reactions in the intermediate temperature range where the solvent motion is treated classically and the tunneling between the solute vibronic states is treated quantum mechanically. The applicability of the method for studies of bond-breaking reactions in polar solvents is demonstrated by calculating the rate of a proton-transfer reaction in the strong coupling limit. In this case the rate constant is expressed as a product of a factor that depends on the fluctuations of the solvent dipoles and a factor that depends on the energies of the solute resonance forms. This type of treatment is expected to be particularly useful in treating the dynamics of enzymatic reaction.
Article
The interdependence of hydrogen – oxygen distances in inorganic compounds can be described by a close correlation if sufficiently high coordination numbers for hydrogen are taken into account. For this purpose, the coordination is favourably defined as chemical coordination number K which is derived from the geometrical coordination or Niggli's Wirkungsbereich. Distances Ri [pm] from one hydrogen to the surrounding oxygen atoms are connected by the rule that the sum of bond valences, s(Ri) equals the valence of hydrogen: [unk]