scispace - formally typeset
Open AccessJournal ArticleDOI

The Structure of the First Coordination Shell in Liquid Water

TLDR
X-ray absorption spectroscopy and x-ray Raman scattering were used to probe the molecular arrangement in the first coordination shell of liquid water and set a strong limit for possible local structure distributions in liquid water.
Abstract
X-ray absorption spectroscopy and x-ray Raman scattering were used to probe the molecular arrangement in the first coordination shell of liquid water. The local structure is characterized by comparison with bulk and surface of ordinary hexagonal ice Ih and with calculated spectra. Most molecules in liquid water are in two hydrogen– bonded configurations with one strong donor and one strong acceptor hydrogen bond in contrast to the four hydrogen– bonded tetrahedral structure in ice. Upon heating from 25°C to 90°C, 5 to 10% of the molecules change from tetrahedral environments to two hydrogen– bonded configurations. Our findings are consistent with neutron and x-ray diffraction data, and combining the results sets a strong limit for possible local structure distributions in liquid water. Serious discrepancies with structures based on current molecular dynamics simulations are observed.

read more

Content maybe subject to copyright    Report

15. J. Higgins, A. Conjusteau, G. Scoles, S. L. Bernasek,
J. Chem. Phys. 114, 5277 (2001).
16. M. J. Murphy, J. F. Skelly, A. Hodgson, B. Hammer,
J. Chem. Phys. 110, 6954 (1999).
17. A. C. Luntz, J. Chem. Phys. 113, 6901 (2000).
18. L. Diekhoner et al., J. Chem. Phys. 115, 9028 (2001).
19. D. C. Jacobs, J. Phys. Condens. Matter 7, 1023 (1995).
20. G. O. Sitz, Rep. Progr. Phys. 65, 1165 (2002).
21. P. R. McCabe, L. B. F. Juurlink, A. L. Utz, Rev. Sci.
Instrum. 71, 42 (2000).
22. H. Hou et al., Science 284, 1647 (1999).
23. M. Gostein, G. O. Sitz, J. Chem. Phys. 106, 7378
(1997).
24. Details are available on Science Online.
25. K. Christmann, O. Schober, G. Ertl, M. Neumann,
J. Chem. Phys. 60, 4528 (1974).
26. Q. Y. Yang, K. J. Maynard, A. D. Johnson, S. T. Ceyer,
J. Chem. Phys. 102, 7734 (1995).
27. J. D. Beckerle, M. P. Casassa, R. R. Cavanagh, E. J.
Heilweil, J. C. Stephenson, Phys. Rev. Lett. 64, 2090
(1990).
28. R. Milot, A. P. J. Jansen, Phys. Rev. B 61, 15657 (2000).
29. L. Halonen, S. L. Bernasek, D. J. Nesbitt, J. Chem. Phys.
115, 5611 (2001).
30. A. Bukoski, I. Harrison, J. Chem. Phys. 118, 9762
(2003).
31. R. D. Beck et al., Science 302, 98 (2003).
32. A. C. Luntz, Science 302, 70 (2003).
33. H. Mortensen, L. Diekhoner, A. Baurichter, A. C. Luntz,
J. Chem. Phys. 116, 5781 (2002).
34. Supported by NSF (grant CHE-0111446) and Tufts
University.
Supporting Online Material
www.sciencemag.org/cgi/content/full/304/5673/992/
DC1
Materials and Methods
2 February 2004; accepted 5 April 2004
The Structure of the First
Coordination Shell in
Liquid Water
Ph. Wernet,
1,2
D. Nordlund,
3
U. Bergmann,
1
M. Cavalleri,
3
M. Odelius,
3
H. Ogasawara,
1,3
L.Na¨slund,
1,3
T. K. Hirsch,
4
L. Ojama¨e,
5
P. Glatzel,
6
L. G. M. Pettersson,
3
A. Nilsson
1,3
*
X-ray absorption spectroscopy and x-ray Raman scattering were used to probe the
molecular arrangement in the first coordination shell of liquid water. The local
structure is characterized by comparison with bulk and surface of ordinary hex-
agonal ice Ih and with calculated spectra. Most molecules in liquid water are in two
hydrogen– bonded configurations with one strong donor and one strong acceptor
hydrogen bond in contrast to the four hydrogen– bonded tetrahedral structure in
ice. Upon heating from 25°C to 90°C, 5 to 10% of the molecules change from
tetrahedral environments to two hydrogen– bonded configurations. Our findings
are consistent with neutron and x-ray diffraction data, and combining the results sets
a strong limit for possible local structure distributions in liquid water. Serious discrep-
ancies with structures based on current molecular dynamics simulations are observed.
Experimental studies of the hydrogen-bonded
network structure in water have mainly relied
on neutron and x-ray diffraction and infrared
(IR) spectroscopies (1). Diffraction data from
noncrystalline materials provide radial distribu-
tion functions (RDFs) (24) that do not provide
angular correlations needed to uniquely assign
local geometries in water (57). A more de-
tailed atomistic picture has been derived theo-
retically by molecular dynamics (MD) simula-
tions (4, 8) that are consistent with diffraction
data. Structural information from IR spectros-
copies generally relies on the correlation be-
tween the O-H stretching frequency and
hydrogen-bond (H-bond) length, which has been
shown to be ambiguous for liquid water (9).
Here, we report an independent experimen-
tal investigation of local bonding configurations
in the first coordination shell of liquid water by
using the near-edge fine structure in x-ray ab-
sorption spectroscopy (XAS), also denoted
XANES and NEXAFS, where a core electron is
excited into empty electronic states. The char-
acter of these states and, hence, the near-edge
fine structure in XAS depends on the chemical
environment, bond lengths, and bond angles
(10). We also obtained the same spectral infor-
mation by using nonresonant x-ray Raman scat-
tering (XRS) involving core excitations (11).
XAS and XRS at the oxygen K-edge (12)
are sensitive to distortions of H-bonds on the
H-sides (donor H-bonds) of the molecules in
the condensed phases of water (11, 13, 14 ).
Because the time scale for excitation is much
faster than the molecular (vibrational) motions
in the liquid, these spectroscopies probe the
electronic structure of a distribution of instan-
taneous configurations and thus allow decom-
position in terms of specific H-bond situations
(12). We analyzed the near-edge structures in
the liquid water XA spectrum (the terminology
“XA spectrum” is used for both XAS and XRS)
with the aid of experimental model systems and
calculated spectra. The XA spectra for water
molecules in different H-bonding configura-
tions are depicted in Fig. 1, where ice Ih bulk
and surface spectra are compared with spectra
of bulk liquid water at two temperatures. Bulk
ice Ih is tetrahedrally coordinated, but the exact
H-bonding environment at the ice Ih surface
still raises questions (15, 16 ). However, there is
consensus that a large fraction (50% or more) of
the molecules in the first half bilayer of the ice Ih
surface has one free O-H group, whereas the other
is H-bonded to the second half bilayer. The liquid
water XA spectrum closely resembles that for the
ice surface, but it is very different from that of
bulk ice. We interpret this finding, and our anal-
ysis demonstrates that the molecules in the liquid
are not predominantly four-coordinated.
The spectra in Fig. 1 can be divided into
three main regions: the pre-edge (around 535
eV), the main edge (537 to 538 eV), and the
post-edge (540 to 541 eV). The bulk ice spec-
trum (Fig. 1, curve a) is dominated by intensity
in the post-edge region and shows a weak main-
edge structure. Both the surface ice (Fig. 1,
curve b) and liquid water (Fig. 1, curve d)
spectra have a peak in the pre-edge region, a
dominant main edge, and less intensity com-
pared with bulk ice in the post-edge region.
Termination of the ice surface with NH
3
(Fig. 1,
curve c) entails a coordination of the free O-H
groups and causes the pre-edge peak to vanish
and the intensity to shift to the post-edge region.
We assign intensities in the pre- and main-edge
regions to water molecules with one uncoordi-
nated O-H group, whereas the intensity in the
post-edge region is related to fully coordinated
molecules. Remarkably, most molecules in
bulk liquid water at room temperature exhibit a
local coordination comparable to that at the ice
surface, with one strong and one non-, or only
weakly, H-bonded O-H group. The contribu-
tion to the spectrum from molecules with four-
fold coordination similar to bulk ice is very
small. Performing the measurements with D
2
O
or H
2
O led to identical spectra within the ex-
perimental resolution, and thus tunneling con-
tributions are not decisive.
Comparison of the XRS spectra of room-
temperature (25°C) and hot water (90°C) (Fig.
1, curve e) shows that heating increases inten-
sities in the pre- and main-edge regions while
decreasing that in the post-edge, but the chang-
es are small compared with the changes ob-
served between ice and the liquid. Figure 1,
curve f, shows the difference spectra of 25°C
water minus ice (17 ) (solid curve) and 90°C
minus 25°C water (circles with error bars). The
latter has been multiplied by a factor of 10 to
1
Stanford Synchrotron Radiation Laboratory, Post
Office Box 20450, Stanford, CA 94309, USA.
2
BESSY, Albert-Einstein-Strasse 15, D-12489 Berlin,
Germany.
3
FYSIKUM, Stockholm University, Al-
baNova, S-10691 Stockholm, Sweden.
4
Department
of Physical Chemistry, Stockholm University,
S-10691 Stockholm, Sweden.
5
Department of
Chemistry, Linko¨ping University, S-58183 Linko¨p-
ing, Sweden.
6
Department of Inorganic Chemistry
and Catalysis, Debye Institute, Utrecht University,
Sorbonnelaan 16, 3584 CA Utrecht, Netherlands.
*To whom correspondence should be addressed. E-
mail: nilsson@slac.stanford.edu
R EPORTS
www.sciencemag.org SCIENCE VOL 304 14 MAY 2004 995

match the intensities. The similarity of the
curves, in particular the same isosbestic point at
538.8 eV, supports a two-component structure
in terms of O-H coordination: Configurations
with one uncoordinated or weakly H-bonded
O-H group replace tetrahedral ones in the ice-
liquid phase transition, and heating liquid water
causes similar types of changes, but of about
one-tenth the magnitude. The contributions of
the two different species at room temperature are
80% and 20% for species with only one strong
H-bond and tetrahedrally coordinated, respective-
ly; for 90°C, we find 85% and 15% with uncer-
tainties of 15 to 20% in both cases (12).
We now compare this phenomenological
finding with a detailed theoretical analysis
based on density functional theory (DFT). The
method used provides highly accurate results
for both intensities and energy positions in x-
ray spectra of H-bonded systems (12, 18). Here,
the calculations provide model spectra for var-
ious distributions of bond lengths and angles as
expected to occur in the liquid. To allow for
calculations of a large number of local config-
urations, we use a small model cluster of 11
molecules, where a central molecule is sur-
rounded by the first coordination shell (4 mol-
ecules) plus part of the second shell (6 mole-
cules saturating the dangling O-H groups of the
first shell). Spectra were calculated for the cen-
tral molecule with systematically varied nearest
neighbor H-bond environments.
First, the model cluster was tested by calcu-
lating the three known cases of bulk ice, ice
surface, and NH
3
terminated ice surface (Fig.
2). The calculated spectra based on clusters
with 11 water molecules reproduce the general
shapes and trends seen in the experimental
spectra in Fig. 1: shift of intensity from post to
pre and main edges when going from a fully
coordinated molecule to one with a free O-H
group (compare Fig. 2, curves a and b) and shift
of intensity back to the post-edge region when
the free O-H group is coordinated upon NH
3
adsorption (compare Fig. 2, curves b and c).
These findings did not depend substantially
on the cluster size: Calculated spectra for iden-
tical nearest neighbor H-bond environments but
now with a total of 44 (Fig. 2, curve a, dashed
line) and 27 (Fig. 2, curves b and c, dashed
lines) water molecules were similar to the spec-
tra for small clusters. Furthermore, thermal
fluctuations and/or vibrational motions did not
affect the calculations considerably; for exam-
ple, superposition of spectra with different in-
tramolecular bonding distances was similar to
that for the mean bonding configuration. The
same holds for intermolecular vibrations (12).
To distinguish possible local configurations
in the liquid, spectra were calculated for the
central molecule in the 11-molecule cluster
Intensity (arb. units)
545540535
Energy (eV)
a
b
c
Fig. 2. Calculated spectra for the model systems
bulk ice Ih. (a) Fully coordinated molecules. (b) Ice
surface (molecules in the topmost surface layer
with one free O-H). (c)NH
3
-terminated ice sur-
face (molecules with an NH
3
-saturated O-H).
Solid lines, small clusters; dashed lines, large clus-
ters. Cluster sizes are 11 molecules for (a) to (c),
solid lines; 44 molecules for (a), dashed line; and
27 molecules for (b) and (c), dashed lines. For the
bulk ice cases, the calculated molecules are fully
coordinated to their four nearest neighbors and
surrounded by: (a, solid line) part of the second
coordination shell (6 molecules) and (a, dashed
line) the full second shell (12 molecules) and 27
further molecules. For the ice-surface cases, an
appropriate number of molecules from the bulk
ice clusters were removed [3 and 17 molecules
for (b, solid line) and (b, dashed line), respectively]
to generate configurations where the calculated
molecule has one free O-H as found on the ice
surface (12, 15, 16). To model the effect of NH
3
termination, the same molecule was calculated
after saturating the dangling O-H by an NH
3
molecule (lone pair electrons of the N oriented
toward the free O-H).
Intensity (arb. units)
545540535
Energy (eV)
a
b
c
d
e
1.0
0.5
-0.5
-1.0
0
x10
f
Fig. 1. (a) Bulk ice Ih (XAS secondary electron
yield). (b) Ice Ih surface (topmost surface
layer, XAS Auger electron yield). (c)NH
3
-
terminated first half bilayer of the ice Ih
surface (XAS Auger electron yield). (d) Liquid
water at ambient conditions [XAS fluores-
cence yield, taken from (13) and additionally
corrected for saturation effects]. (e) Bulk liq-
uid water at 25°C (solid line) and 90°C
(dashed line) (XRS, spectra normalized to the
same area). (f) Difference spectra: 25°C wa-
ter minus bulk ice (solid curve) and 90°C
water minus 25°C water (circles with error
bars). The latter difference spectrum has
been multiplied by a factor of 10. Note that
XRS and XAS essentially give the same infor-
mation, as indicated by the similarity of the
room-temperature water spectra (e) and (d).
Their slight differences are due to the poorer
energy resolution of XRS. The temperature
effects can best be studied with XRS because
of the larger probing depth and the resulting
easier sample preparation than for XAS. For
details on sample preparation and probing
depths, see (1113, 15).
Table 1. Relative amounts of local configurations—
double-donor (DD), single-donor (SD), and non-
donor (ND) configurations—in liquid water ac-
cording to different MD simulations at 25°C
and 90°C. The values in the first column for
each temperature follow from the experimental
observations (EXP) in Fig. 1 (12) and from fitted
spectra (FIT) such as the ones shown in Fig. 5.
The errors are given by the uncertainties relat-
ed to the experimental observations (Fig. 1) and
the calculations (12).
Method
Type EXP FIT SPC MCYL CPMD
25°C
DD 15
15
25
70 50 79
SD 80 20 27 41 20
ND 5 5391
90°C
DD 10
10
25
56 39 63
SD 85
20
15
37 47 34
ND 5 5 7 14 3
R EPORTS
14 MAY 2004 VOL 304 SCIENCE www.sciencemag.org996

with systematically varied nearest neighbor H-
bond distances and angles (Fig. 3A, curves a to
h) (12). The spectrum for a configuration with
tetrahedral coordination [linear H-bonds: all
␪⫽0° and 2.75 Å for all O-O distances (1, 2),
Fig. 3A, curve a] shows a maximum close to
540 eV, similar to Fig. 2, curve a. Donor H-
bond distortions were introduced by varying the
O-O distance r and the angle of the nearest
neighbor (and the two attached molecules; see
inset of Fig. 3A) on only one H-side (Fig. 3A,
curves b to e) and on both H-sides (Fig. 3A,
curves f to h). Intermediate elongations smear out
the maximum close to 540 eV (Fig. 3A, curve b).
Upon further distortion of one donor H-bond, a
pre-edge peak emerges at 535 eV and a main-
edge structure develops at 537 eV, while post-
edge intensities decrease (Fig. 3A, curves c to e).
Additional distortion of the second donor H-bond
(Fig. 3A, curves f and g) further reduces intensi-
ties above 540 eV without considerable change in
the pre- and main-edge regions. Note that varying
r at constant and varying at constant r can
yield similar spectra (e.g., Fig. 3A, curves c and e).
Spectra c to e and g to h in Fig. 3A show
that donor H-bond distortions of a certain de-
gree manifest as a distinct pre-edge peak and an
intense main edge in the oxygen K-edge XA
spectrum of water. The analogy of these pre-
and main-edge features to the ones observed for
the ice-surface molecules with one broken do-
nor H-bond (Figs. 1 and 2) suggests denoting
the corresponding distorted H-bonds as bro-
ken. More specifically, based on spectral cal-
culations of a large number of different config-
urations in the liquid, we define the cutoff for
H-bonding in terms of O-O distances as the
boundary between zones A and B (black line in
Fig. 3B) (12). For linear H-bonds, we find the
same cutoff as conventionally used in the anal-
ysis of MD simulations (19). For bent H-bonds,
we use a smaller cutoff because of the occur-
rence of pre- and main-edge intensities for large
angular distortions at standard O-O distances
(Fig. 3A, curve e). This difference is not im-
portant, as will be addressed in detail below.
Zones A and B in Fig. 3B can be used to
classify local water configurations: The environ-
ment in bulk ice, e.g., is an AA configuration, with
both nearest neighbor H-bond acceptor molecules
within zone A, whereas that at the ice surface is an
AB configuration. We define three groups of con-
figurations in the liquid: Double-donor (DD) con-
figurations with two intact donor H-bonds are
characterized by XA spectra without pre-edge
peak (Fig. 3A, curves a, b, and f); single-donor
(SD) configurations with one intact and one bro-
ken donor H-bond exhibit a distinct pre-edge peak
and an intense main edge (Fig. 3A, curves c, d, e,
and g); and the nondonor (ND) configuration
with two broken donor H-bonds gives two dis-
tinct peaks (Fig. 3A, curve h) similar to the
spectra of gas-phase water and the species at the
surface of liquid water (20).
The pre-edge intensities for the SD configura-
tions are due to excitations to unoccupied anti-
bonding orbitals that are localized along the inter-
nal O-H bond (Fig. 4). The localization is caused
by breaking one donor H-bond while keeping the
other intact, which entails s-p-rehybridization in
the orbitals close to 535 eV (14). The resulting
increase of p-character translates to an intensity
increase in the pre-edge region of the XA spectra
through the dipole selection rule (12). The orbital
that contributes to the pre-edge intensity closely
matches zone A, as shown in Fig. 4. This com-
parison gives our H-bond criterion its physical
basis and supports our interpretation; the presence
of another molecule inside zone A changes this
orbital and thereby the corresponding spectral fea-
ture, which also explains the sensitivity of XAS
(XRS) to H-bond distortions on the H-side.
The occurrence of the pre-edge feature, fur-
thermore, can be directly correlated with the
energetics in the H-bond interaction. For the
model clusters considered here, we find that
distortions leading to enhanced intensity in the
pre-edge correspond to one donor H-bond los-
ing about 40 to 70% of the individual bond
energy of bulk ice (12). The SD species that
dominate the liquid thus have one strong and
one weak or broken donating H-bond. Consis-
tent with commonly used definitions of
H-bonds in liquid water (and for simplicity), we
use the term broken for the distorted donor
H-bond of an SD configuration.
In Fig. 3A, only the calculated spectra of SD
species in curves c to e and g exhibit all features
observed in the experimental liquid water spec-
Intensity (arb. units)
545540535
Energy (eV)
a
b
c
d
e
f
g
h
2.75, 0°
2.75, 0°
3.50, 0°
2.75, 0°
3.10, 0
°
2.75, 0°
3.50, 35
°
2.75, 0°
2.75, 50
°
2.75, 0°
2.75, 20
°
3.00, 0°
3.50, 0°
3.00, 0°
3.90, 0°
3.90, 0°
r , θ
11
r , θ
2 2
r
1
1
1
2
r
2
2
θ
θ
3.8
3.6
3.4
3.2
3.0
2.8
2.6
r (Å)
-50 -25
0
25 50
θ
(deg)
A
B
AB
Fig. 3. (A) Calculated spectra
for a cluster of 11 water mol-
ecules with O-O distances r
i
(in Å) and angles
i
(in de-
grees) of the nearest neigh-
bors i 1, 2 on the H-sides
(nearest neighbor H-bond ac-
ceptor molecules) systemati-
cally varied as given for each
spectrum. Starting from (a) a
tetrahedral structure with all
four nearest neighbor O-O
distances at 2.75 Å and linear H-bonds, for (b, c) only r
1
, for
(d) r
1
and
1,
and for (e) only
1
are varied to introduce
H-bond distortions on one H-side. For (f)
1
and r
2
and for
(g, h) r
1
and r
2
are varied to introduce H-bond distortions
on both H-sides. Compared with the cluster calculation for
ice Ih in Fig. 2, curve a, solid line, relative orientations of the
molecules have been changed arbitrarily to better account
for the less ordered local configurations in the liquid. In
addition, the calculated oscillator strengths are convoluted
with a slightly larger width to account for the larger
vibrational motion of the molecules in the liquid. This
explains the differences between (a) and Fig. 2, curve a,
solid line. (B) The systematic spectral changes allow for a
definition of zones A and B to denote the positions of the
two nearest neighbor H-bond acceptor molecules. Spectra
show similar features for positions within one zone [com-
pare, e.g., (a) and (b) or (c) to (e) and as verified by many
further calculations], and the cases shown here can be
considered to be representative. The boundary between the
zones (black line) is used as a cutoff for H-bonding (12)as
based on the occurrence of strong pre- and main-edge
features (similar to the experimental ice surface spectrum)
when the nearest neighbor is outside zone A beyond the
black line (e.g., c to e).
Fig. 4. Connecting electronic and geometric
structures. The contour plot of a typical unoc-
cupied orbital giving rise to the pre-edge inten-
sity in the XA spectrum of a water molecule
with a broken donor H-bond (solid line contour
plot) is compared with our cutoff criterion for
H-bonding (black line in Fig. 3B) (12). Broken
donor H-bonds are characterized by nearest
neighbor O’s outside the shaded area (12). The
orbital contour has been calculated for the
central molecule in a cluster with 11 molecules
(the corresponding XA spectrum is shown in
Fig. 2, curve b, solid line) where the H-bond on
one H-side is broken (SD configuration). Only 4
of the 11 molecules are displayed.
R EPORTS
www.sciencemag.org SCIENCE VOL 304 14 MAY 2004 997

trum: a pre-edge peak close to 535 eV and a
strong main edge close to 537 eV. This result
suggests, consistent with our phenomenological
findings, that asymmetric SD species with one
strong and one broken donor H-bond dominate
the liquid water structure. In a more detailed
analysis, our experimental data were fitted with
sums of calculated spectra. The results are sum-
marized in Table 1. Different combinations of
computed spectra corresponding to different
choices of bond lengths and angles within each
class give similar spectra. All yield a dominant
contribution of SD species both at 25°C and
90°C. We note the close agreement with the
previously described analysis based on model
experimental spectra (12). For 90°C we find the
best fits for a decreased (increased) amount of DD
(SD) species as compared with room temperature,
indicating the breaking of H-bonds with increas-
ing temperature along with the decrease of fully
coordinated molecules (21). In particular, it comes
as no surprise that bulk-icelike configurations are
virtually absent in water close to the boiling point.
Consistent with the discussion of Fig. 1, we find
that DD configurations are replaced by SD con-
figurations both when heating the liquid and when
melting ice into liquid water. Computed spectra
indicate that XAS is not sensitive to the breaking
of H-bonds on the O-side (acceptor H-bond) (13).
However, symmetry arguments imply that, with a
dominant amount of molecules in SD configura-
tions, most molecules must in addition have a
broken acceptor H-bond. This entails that, within
the H-bond definition given here [Fig. 3 and (12)],
each molecule has on average 2.2 0.5 (2.1
0.5) H-bonds at 25°C (90°C).
As mentioned above, XAS cannot distin-
guish between SD configurations with a broken
bond due to elongation (Fig. 3A, curve c) or to
strong bending (Fig. 3A, curve e). However,
these configurations exhibit different nearest
neighbor O-O and O-H distances. We compare
our results with O-O and O-H RDFs from
neutron diffraction data taken at ambient con-
ditions (2, 22). Two XA spectra generated with
different distributions of H-bond lengths and
angles, but both with the same fraction of DD,
SD, and ND species, are shown in Fig. 5A,
curves a and b. Spectra of this type have been
used to estimate the amounts of the different
species (Table 1) that are consistent with the
measured XA spectrum.
The O-O (Fig. 5B) and O-H (Fig. 5C) RDFs
are shown for the two model distributions cor-
responding to the XA spectra in Fig. 5A, curves
aandb(12). The 11-molecule model configu-
rations that can adequately describe the local
structure probed by XAS are obviously not
sufficient for a full description of RDFs beyond
the first coordination shell [3.4 to 3.5 Å in the
O-O RDF (2, 4) and 2.5 to 2.6 Å in the O-H
RDF (2)]. Molecules from the second shell are
not included in the RDFs of models a to c in
Fig. 5 and hence we should not expect agree-
ment for larger distances. The a model has a
larger portion of broken H-bonds as a result of
bending, whereas the b model has more elon-
gated bonds. Both models reproduce the exper-
imental XA spectrum, but model a clearly ap-
pears to provide a better agreement with the
RDFs within the first coordination shell. Model
b lacks structural contributions in the region 2.9
to 3.3 (1.9 to 2.3) Å and overemphasizes the
contributions around 3.5 and 2.5 Å. It is thus
inconsistent with water density, as illustrated by
the poor agreement with the diffraction data.
Combining XAS and diffraction results therefore
suggests that the H-bonds are predominantly bro-
ken by bending rather than by elongation.
Finally, our results can be compared with
two classical MD simulations using the flexible
SPC (23) and MCYL (24 ) pair potential energy
models and to an ab initio Car-Parrinello MD
(CPMD) simulation (25) [for details of the sim-
ulations, see (12)]. The sampling of the distri-
bution of structures is similar to that in the
XAS/XRS experiment, i.e., the instantaneous
geometries at a sequence of time steps are
analyzed. Using our criterion for H-bonding
(12), we have determined the amounts of DD,
SD, and ND configurations (Table 1). The
MCYL potential yields the largest amount of
SD, but all of the simulations deviate substan-
tially from experimental results. Note that with
our spectroscopically determined electronic
structure criterion, SPC (CPMD) at 25°C still
gives an average of 3.3 (3.6) H-bonds per mol-
ecule, similar to earlier theoretical studies re-
porting 3.5 H-bonds per molecule (19). For
example, with the numbers in Table 1 for SPC
at 25°C, we determine that (0.7 4) (0.27
2) (0.03 0) 3.34 H-bonds per molecule.
An XA spectrum generated based on the distri-
bution of species obtained from our SPC MD
simulation at 25°C is shown in Fig. 5A, curve c.
Clearly the agreement with the experimental
spectrum is very poor: no pre-edge intensity
(SD species) and too much intensity around 540
eV (fully coordinated DD species). The corre-
sponding O-O (O-H) RDFs (12) are depicted in
Fig. 5B (C), curve c, as solid lines. We note that
the SPC simulation, in comparison with exper-
imental results, exaggerates contributions at r-
values characteristic for tetrahedral/near tetra-
hedral configurations (in the maxima) in both
the O-O and O-H RDFs. However, the O-O
(O-H) RDFs from SPC and our model a are
rather similar at 2.9 to 3.3 (1.9 to 2.3) Å. The
poor agreement with the experimental XA
spectrum is hence attributed to too few species
with large H-bond angles (SD species) in the
SPC simulation. Similarly, the other two MD
simulations shown in Table 1 cannot create an
XA spectrum that fits the data. The MD simu-
lations are, however, consistent with our results
in predicting the decrease (increase) of DD
(SD) species with increasing temperature.
The large amount of SD species can also
explain femtosecond-IR studies indicating
two different O-H groups in liquid water: one
strongly and one weakly H-bonded (26 ).
545540535
Energy (eV)
a
b
c
XAS
XRS
2.52.01.5
r (Å)
a
b
c
g
OH
(r)
3.53.02.5
r (Å)
a
b
c
g
OO
(r)
ABC
Fig. 5. Calculated XA spectra and RDFs (solid lines) for three models—(a), (b), and (c)—are
compared with (A) experimental XA spectra and RDFs (B and C) at 25°C. Spectra and RDFs have
been generated simultaneously for the central molecule in clusters consisting of 11 molecules, for
a total of 14 clusters (12). A sampling of the possible DD, SD, and ND configurations for the
calculated molecule in the liquid is guaranteed by using a variety of nearest neighbor O-O distances
and H-bond angles. The RDFs were calculated for the first coordination shell only. Experimental
RDFs [open circles in (B) and (C)] were derived from neutron diffraction (2). The calculated XA
spectra were broadened witha1eVGaussian for better comparison with experimental XRS data
[open circles in (A)] taken from Fig. 1, curve e. Models (a) and (b) use a balance of DD:SD:ND
10:85:5. Model a (b) has more (less) angularly distorted H-bonds and less (more) elongated ones.
Model (c) (solid line) represents the SPC simulation (Table 1) with a balance DD:SD:ND 70:27:3,
as determined with our criterion for H-bonding (12). For comparison, the dashed lines in B and C,
curve c, show the O-O and O-H RDFs calculated from the full SPC simulation. There is good
agreement up to 3.2 Å (O-O) and 2.4 Å (O-H), validating our approach to characterize the first shell
in the liquid. Only model (a) or similar balances with a dominant fraction of SD species (Table 1)
give good fits when simultaneously considering our experimental XRS data and the RDF curves.
R EPORTS
14 MAY 2004 VOL 304 SCIENCE www.sciencemag.org998

The orientational relaxation dynamics were
shown to be directly connected with the
strength of the two H-bond groups, where the
weak H-bonds relax much faster than the
strong H-bonds. According to our results, the
number of strong H-bonds in the liquid is
substantially smaller than expected, which
may seem in contradiction with the small heat
of melting compared with the heat of subli-
mation for ice. However, quantum chemical
calculations have shown that each bond in the
proposed SD configurations is stronger than
the average bond in four-fold coordination
because of anticooperativity effects (27, 28).
Thus, the large number of weakened/broken
H-bonds in the liquid leads to only a small
change in energy. A recently developed
quantum chemical model (1, 27 ) that propos-
es the predominance in the liquid phase of
two hydrogenbonded water molecules in
ring conformations is consistent with our re-
sults. Water is a dynamic liquid where H-
bonds are continuously broken and reformed
(29). The present result that water, on the
probed subfemtosecond time scale, consists
mainly of structures with two strong
H-bonds, one donating and one accepting,
nonetheless implies that most molecules are
arranged in strongly H-bonded chains or
rings embedded in a disordered cluster net-
work connected mainly by weak H-bonds.
References and Notes
1. R. Ludwig, Angew. Chem. Int. Ed. Engl. 40, 1808 (2001).
2. A. K. Soper, Chem. Phys. 258, 121 (2000).
3. A. H. Narten, H. A. Levy, Science 165, 447 (1969).
4. T. Head-Gordon, G. Hura, Chem. Rev. 102, 2651 (2002).
5. P. G. Kusalik, I. M. Svishchev, Science 265, 1219 (1994).
6. A. K. Soper, J. Chem. Phys. 101, 6888 (1994).
7. A. A. Chialvo, P. T. Cummings, J. Phys. Chem. 100,
1309 (1996).
8. F. H. Stillinger, Science 209, 451 (1980).
9. J. Stenger, D. Madsen, P. Hamm, E. T. J. Nibbering, T.
Elsaesser, Phys. Rev. Lett. 87, 027401 (2001).
10. J. Sto¨hr, NEXAFS Spectroscopy (Springer-Verlag, Ber-
lin, 1992).
11. U. Bergmann et al., Phys. Rev. B 66, 092107 (2002).
12. Details of the materials and methods, and supporting
analysis of the experimental data, are available at
Science Online.
13. S. Myneni et al., J. Phys. Condens. Matter 14, L213
(2002).
14. M. Cavalleri, H. Ogasawara, L. G. M. Pettersson, A.
Nilsson, Chem. Phys. Lett. 364, 363 (2002).
15. D. Nordlund et al., unpublished data.
16. A. Glebov, A. P. Graham, A. Menzel, J. P. Toennies, P.
Senet, J. Chem. Phys. 112, 11011 (2000).
17. The ice spectrum (Fig. 1, curve a) has been broadened
to the same instrumental resolution as for XRS and
normalized to the same area before subtraction.
18. M. Nyberg, M. Odelius, A. Nilsson, L. G. M. Pettersson,
J. Chem. Phys. 119, 12577 (2003).
19. A. Luzar, D. Chandler, Phys. Rev. Lett. 76, 928 (1996).
20. K. R. Wilson et al., J. Phys. Condens. Matter 14, L221
(2002).
21. The sharper main edge for hot water (537 eV, Fig. 1, curve
e, dashed line) is accounted for by a different balance
inside the group of SD species, where more of the H-bond
acceptor molecules have distances closer to the AB bound-
ary, i.e., have a larger r
2
and/or
2
but still remain in zone
A (more Fig. 3A, curve g, and less curve c, e.g.). Accordingly,
the intensity decrease above the isosbestic point with
temperature can be mainly assigned to the loss of DD
configurations with tetrahedral and near tetrahedral envi-
ronments by 5 to 10% and an increase of SD configura-
tions with H-bond acceptor molecules being closer to the
AB boundary.
22. The neutron O-O RDF in (2) is very similar to the
most recently derived O-O RDF from x-ray diffrac-
tion in (4).
23. K. Toukan, A. Rahman, Phys. Rev. B 31, 2643 (1985).
24. G. C. Lie, E. Clementi, Phys. Rev. A. 33, 2679 (1986).
25. P. L. Silvestrelli, M. Parrinello, J. Chem. Phys. 111,
3572 (1999).
26. S. Woutersen, U. Emmerichs, H. J. Bakker, Science
278, 658 (1997).
27. F. Weinhold, J. Chem. Phys. 109, 373 (1998).
28. L. Ojama¨e, K. Hermansson, J. Phys. Chem. 98, 4271
(1994).
29. C. J. Fecko, J. D. Eaves, J. J. Loparo, A. Tokmakoff, P. L.
Geissler, Science 301, 1698 (2003).
30. Supported by the Swedish Foundation for Strategic Re-
search, Swedish Natural Science Research Council, and U.S.
National Science Foundation grant CHE-0089215. Gener-
ous grants of computer time at the Swedish National
Supercomputer Center and the Center for Parallel Com-
puting are gratefully acknowledged. Portions of this re-
search were carried out at the Stanford Synchrotron Radi-
ation Laboratory, a national user facility operated by Stan-
ford University on behalf of the U.S. Department of Energy,
Office of Basic Energy Sciences. Use of the Advanced
Photon Source (APS) was supported by the U.S. Depart-
ment of Energy, Basic Energy Sciences, Office of Science,
under contract No. W-31-109-ENG-38. Biophysics Collab-
orative Access Team (BioCAT) is a National Institutes of
Health–supported Research Center RR-08630. The Ad-
vanced Light Source (ALS) is supported by the Director,
Office of Science, Office of Basic Energy Sciences, Materi-
als Sciences Division, of the U.S. Department of Energy
under Contract No. DE-AC03-76SF00098 at Lawrence
Berkeley National Laboratory. Assistance by the APS, ALS,
and the Swedish national laboratory MAX-lab staff is
gratefully acknowledged. We thank J. B. Hastings for his
valuable comments and discussions and S. P. Cramer for
making the XRS spectrometer available.
Supporting Online Material
www.sciencemag.org/cgi/content/full/1096205/DC1
Materials and Methods
SOM Text
Figs. S1 to S11
Tables S1 and S2
References and Notes
29 January 2004; accepted 25 March 2004
Published online 1 April 2004;
10.1126/science.1096205
Include this information when citing this paper.
Asphalt Volcanism and
Chemosynthetic Life in the
Campeche Knolls, Gulf of Mexico
I. R. MacDonald,
1
* G. Bohrmann,
2
E. Escobar,
3
F. Abegg,
2
P. Blanchon,
4
V. Blinova,
2
W. Bru¨ckmann,
5
M. Drews,
5
A. Eisenhauer,
5
X. Han,
6
K. Heeschen,
2
F. Meier,
2
C. Mortera,
7
T. Naehr,
1
B. Orcutt,
8
B. Bernard,
9
J. Brooks,
9
M. de Farago´
10
In the Campeche Knolls, in the southern Gulf of Mexico, lava-like flows of solidified
asphalt cover more than 1 square kilometer of the rim of a dissected salt dome at
a depth of 3000 meters below sea level. Chemosynthetic tubeworms and bivalves
colonize the sea floor near the asphalt, which chilled and contracted after discharge.
The site also includes oil seeps, gas hydrate deposits, locally anoxic sediments, and
slabs of authigenic carbonate. Asphalt volcanism creates a habitat for chemosyn-
thetic life that may be widespread at great depth in the Gulf of Mexico.
Salt tectonism in the Gulf of Mexico hy-
drocarbon province controls the develop-
ment of reservoirs and faults that allow oil
and gas to escape at the sea floor (1). More
than 30 years ago, investigators studying
the Gulfs abyssal petroleum system (2)
photographed an asphalt deposit (3) among
salt domes in the southern Gulf of Mexico.
During exploration of the Campeche
Knolls, about 200 km south of the photo-
graphed site (Fig. 1, A and C), we have now
found numerous, deeply dissected salt
domes with extensive slumps and mass
wasting at depths of 3000 m or greater.
Massive, lava-like flow fields of solidified
asphalt, evidently discharged at tempera-
tures higher than the ambient bottom water
(4°C), have been colonized by an abundant
chemosynthetic fauna.
The Campeche Knolls are salt diapirs ris-
ing from an evaporite deposit that underlies
the entire slope region (4) and hosts the
Campeche offshore oil fields (5 ). Numerous
reservoir and seal facies have also been at-
1
Physical and Life Sciences Department, TexasA&M
University–Corpus Christi, 6300 Ocean Drive, Corpus
Christi, TX 78412, USA.
2
Fachbereich 5 Geowissen-
schaften, University Bremen, D-28334 Bremen, Germa-
ny.
3
Universidad Nacional Autonoma de Mexico, Insti-
tuto de Ciencias del Mar y Limnologı`a, Apartado Postal
70-305, Mexico 045510, D.F. Mexico.
4
Instituto de Cien-
cias del Mar y Limnologı`a, Apartado Postal 1152, Can-
cu´n, D.F. Mexico.
5
IFM-GEOMAR, Leibniz-Institut fu¨r
Meereswissenschaften, D-24148 Kiel, Germany.
6
Second
Institute of Oceanography, State Oceanic Administra-
tion, Zhejiang 310012, China.
7
Universidad Nacional
Autonoma de Mexico, Instituto de Geofisica, Mexico,
04510, D.F. Mexico.
8
University of Georgia, 220 Marine
Sciences Building, Athens, GA 30602, USA.
9
TDI-Brooks
International, Inc. 1902 Pinion, College Station, TX
77845, USA.
10
Aurensis, SA, San Francisco de Sales 38,
28003, Madrid, Spain.
*To whom correspondence should be addressed. E-
mail: imacdonald@falcon.tamucc.edu
R EPORTS
www.sciencemag.org SCIENCE VOL 304 14 MAY 2004 999
Figures
Citations
More filters
Journal ArticleDOI

Water as an Active Constituent in Cell Biology

Philip Ball
- 01 Jan 2008 - 
TL;DR: The recent confirmation that there is at least one world rich in organic molecules on which rivers and perhaps shallow seas or bogs are filled with nonaqueous fluidsthe liquid hydrocarbons of Titan now bring some focus, even urgency, to the question of whether water is indeed a matrix of life.
Proceedings Article

Attosecond Physics

TL;DR: In this paper, an attosecond "oscilloscope" was used to visualize the oscillating electric field of visible light with an oscillator and probe multi-electron dynamics in atoms, molecules and solids.
Journal ArticleDOI

Effect of ions on the structure of water: structure making and breaking.

TL;DR: Results from Diffraction Experiments 1351 and results from Spectroscopic Measurements 1354 4.3.1.
Journal ArticleDOI

Hydrophilic interaction chromatography.

TL;DR: The review attempts to summarize the ongoing discussion on the separation mechanism and gives an overview of the stationary phases used and the applications addressed with this separation mode in LC.
References
More filters
Journal ArticleDOI

Water: From Clusters to the Bulk

TL;DR: The exploration of structural and binding properties of small water complexes provides a key for understanding bulk water in its liquid and solid phase and for understanding solvation phenomena.
Journal ArticleDOI

The radial distribution functions of water and ice from 220 to 673 K and at pressures up to 400 MPa

TL;DR: In this article, it is shown that the empirical potential structure refinement procedure, which attempts to fit a three-dimensional ensemble of water molecules to all three partial structure factors simultaneously, leads to improved reliability in the extracted radial distribution functions.
Journal ArticleDOI

Effect of environment on hydrogen bond dynamics in liquid water.

TL;DR: It is demonstrated that the long time dynamics of a single hydrogen bond in ambient liquid water is indeed characterized by significant nonexponential relaxation, and this complex relaxation is essentially uncorrelated to the specific bonding patterns near the tagged hydrogen bond.
Journal ArticleDOI

Ultrafast hydrogen-bond dynamics in the infrared spectroscopy of water.

TL;DR: Using simulations of an atomistic model of water, this work relates frequency fluctuations in the OH-stretching frequency of HOD in liquid D2O with femtosecond infrared spectroscopy to intermolecular dynamics.
Journal ArticleDOI

Molecular-dynamics study of atomic motions in water

TL;DR: Using a flexible version of a rigid-molecule model of water, the velocity autocorrelation functions are analyzed to investigate the effect of the liquid milieu on the high-frequency internal modes of molecular motion.
Related Papers (5)
Frequently Asked Questions (1)
Q1. What are the contributions mentioned in the paper "The structure of the first coordination shell in liquid water" ?

In this paper, the authors present a survey of the state-of-the-art work in this area.