Tải bản đầy đủ (.pdf) (333 trang)

PROCEEDINGS OF THE INTERNATIONAL SCHOOL OF PHYSICS "ENRICO FERMI"_2 potx

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (23.94 MB, 333 trang )

This page intentionally left blank
Hydrogen
bonding
at
high pressure
J. S.
LOVEDAY
Department
of
Physics
and
Astronomy
and
Centre
for
Science
at
Extreme
Conditions
The
University
of
Edinburgh
-
Mayfield
Rd,
Edinburgh
EH9
3JZ,
Scotland,
UK


1.
-
Introduction
The
properties
of the
hydrogen bond
are
applicable
to a
wide range
of fields.
They play
a
crucial role
in
many areas
of
biology:
the
base pairings
in DNA are the
result
of
H-bonds,
the
behaviour
of
water
and

other H-bonded solvents
are
crucial
in
chemistry, H-bonds
and
their directional
nature
are
responsible
for the
structural
versatility
of ice
giving
rise
to at
least
eleven
phases
below 2GPa, hydrogen bonding plays
an
important
role
in
determining
the
dehydration properties
of
hydrous minerals, implicated

as a
possible
cause
of
deep-focus earthquakes [1],
and
since
the
outer planets
and
their
satellites
contain
large
quantities
of
ice, ammonia
and
methane,
the
properties
of
these
systems
are
crucial
to
planetary
modelling.
This

ubiquity provides
a
very
powerful
motivation
to
understand
the
microscopic behaviour
of
hydrogen bonding, including,
the
relationships between
bonding
strength, atomic species
and
bond geometry [2].
2.

Definitions
Figure
1
shows
a
schematic
of a
hydrogen bond. Atom
A is
covalently bonded
to

a
hydrogen which hydrogen bonds
to
atom
B.
Atom
A is
referred
to as the
donor
and
B the
acceptor.
The
criteria
which determine
if a
particular
contact
is a
hydrogen
bond
are
somewhat subjective
but
consist
of a
combination
of
geometric

and
vibrational
properties.
The
principal criterion
is
that
the H • • • B
distance
is
less than
the sum of
the van der
Waals
radii
of H and B
—taking
the
value
for H to be 1A
[3].
In
addition,
there
is an
expectation
that
the A-H
stretch vibrational mode should
soften

and
that
©
Societa
Italiana
di
Fisica
357
358 J. S.
LOVEDAY
Atom
A
Atom
B
Donor
,.
Acceptor
Atom
A
Atom
B
Donor
Acceptor
H
6-
Fig.
1. - A
schematic diagram
of
long

(upper)
and
short
(lower)
H-bonds.
the A-H • • • B
libration mode should
stiffen.
For
long hydrogen bonds
the
interaction
is
considered
to be
largely ionic between
a
somewhat positive hydrogen atom —indicated
by
a 8+ in fig. 1— and a
somewhat negative atom
B
—indicated
by a 6—. As
hydrogen
bonds
shorten,
they develop
a
more covalent character with transfer

of
bonding electron
density
from A-H to H • • • B as
shown.
The
example shown
is a
simple linear H-bond,
but
it is
possible
to
have poly-furcated hydrogen bonds where
H
forms
bonds
to
more than
one
B
atom,
or B
forms
bonds
to
multiple
H
atoms. Finally,
B

need
not be an
atom;
it
may
be an
accumulation
of
electron density
as in
ethyne
where
C-H
forms
H-bonds
to
the
carbon-carbon triple bonds [4].
3.

Techniques
The
principal microscopic
properties
needed
to
characterise
a
hydrogen bond
are its

geometry
and the
strength
of the
bonds;
in
addition,
it is
clearly important
to
understand
the
nature
of the
bonding.
As a
result,
for
high pressure studies,
the
techniques generally
used
are
optical
and
infra-red
measurements
of
vibrational
frequencies,

diffraction
studies
to
characterise
the
geometry,
and ab
initio modelling studies
that
explore
the
nature
of
the
bonding. Other techniques like nuclear magnetic resonance
and
neutron inelastic
scattering
have proved very
powerful
for
studies
of
H-bonds
at
ambient
pressure
but
have
not yet

been seriously applied
at
high pressure.
3'1.
Vibrational
spectroscopy.

Spectroscopy using photons
was
amongst
the
earliest
techniques
to be
applied
to
H-bonding
at
high
pressure.
Here
the frequencies of
modes
of
vibration
are
measured
by
their
coupling

to the
incident light
via a
change
in
dipole
moment
(infra-red)
or
polarisability (Raman).
The
attraction
of
such measurements
is
that
the
softening
of the A-H
stretch mode
(referred
to
here
as the
vibron)
is one of the
primary indications
of
strengthening hydrogen bonds,
and

this mode
is
easily identified
for
long hydrogen bonds. Although spectroscopic
data
are
relatively easy
to
measure,
interpretation
and
mode assignment
are
often
difficult.
In
addition,
one of the
primary
HYDROGEN
BONDING
AT
HIGH
PRESSURE
359
aims
of
spectroscopic studies
has

been
to
explore short H-bonds close
to
molecular dis-
sociation. Under these conditions
the
vibron moves into regions where diamonds have
absorption bands
and
interaction
between
the
vibron
and
other
vibrational
modes
be-
comes
significant. However, innovations
in
cell
design, improvements
in the
quality
of IR
data
made possible
by the use of

synchrotron light sources,
and the use of
modelling
in
combination with experiments have
led to
considerable improvements
in the
quality
of
information
available
[5, 6].
Other spectroscopic techniques have been used
for
measurements
of
vibrational fre-
quencies
including neutron
[7] and
X-ray
[8]
triple-axis studies
of
phonon dispersion,
incoherent
neutron spectroscopic measurements
of
density

of
states
[9] and
X-ray
nu-
clear
spectroscopy measurements
of
partial density
of
states
[10].
For
H-bonded systems
however,
the
vast bulk
of
spectroscopic
data
are
obtained using photons.
For
this rea-
son
the
term spectroscopic used
in
this lecture
refers

to
measurements
of
vibrational
frequencies
using Raman
or IR
methods.
3'2. Structural studies.
-
Diffraction
studies
are the
only means
to
measure
the ge-
ometry
of
H-bonds
and are
thus
a
crucial component
of any
attempt
to
characterise
an
H-bonded

system. Although X-ray studies
are
able
to
locate hydrogen atoms
and can
identify
the
H-bond
contacts
in a
structure, neutron diffraction
is the
only technique
able
to
measure
the
geometry
sufficiently
precisely. Studies
of
H-bonded systems were
a
primary
motivation
of the
development
of
high pressure neutron

diffraction
[11,12]
and
form
a
significant fraction
of the
studies performed.
The
Paris-Edinburgh cell
is now
able
to
achieve
a
pressure
of 30 GPa for
such studies [11, 12]. Although this represents
a
significant
pressure range,
it is not
sufficient
to
explore dissociation
of
H-bonds
in
simple
molecular

systems. Studies
of
dissociation
of
H-bonds
in
simple molecular solids remain
an
important motivation
for
further
extensions
of the
pressure range.
3'3.
Ab
initio modelling.
- The
capabilities
and
accuracy
of ab
initio modelling studies
have
seen remarkable recent improvement.
Two
basic methods exist
to
carry
out

such
modelling.
In the first
(static total-energy calculations)
the
total
energy
is
computed
for
a fixed
configuration
of
atoms
and the
best configuration
is
found
by
exploring
the
variation
in
total
energy with change
in
configuration. Static techniques have
had
success
in

studies
of
H-bonding [13,
14] but are
limited
by the
difficulty
of
handling disorder.
The
development
of ab
initio molecular dynamics (the Car-Parrinello method) [15] overcomes
this limitation
and has
revolutionised modelling
of
H-bond systems.
In
this
method,
the
time evolution
of the
system
is
followed
with
the
motion

of the
particles being determined
from
a
self-consistent solution
of the
electronic Hamilitonian calculated
at
each time
step.
Considerable
effort
has
been
put
into development
of
techniques
to
handle
the
hydrogen
atom
as a
quantum object [16].
As a
result, remarkable agreement between observation
and
modelling
can be

obtained. Theoretical studies
are
generally
not
able
to
identify
the
structure
ab
initio, however,
and
require structural information
as a
start
point.
360
J. S.
LOVEDAY
5 10 15 20
P(GPa)
25
Fig.
2. - The
measured
pressure
variation
of the
intramolecular
O-D

bond length
in ice
VIII
[18]
shown
as
open circles
and the
crosses
are the
results
of
Hartree-Fock calculations.
The
dotted
line
shows
the
variation
estimated
from
previous spectroscopic
studies
[17].
4.

Molecular
systems:
water-ice
The

solid phases adopted
by the
water molecule have become model systems
for
studies
of
H-bonding
at
high pressure.
At the
molecular level water
is one of the
simplest
H-bonded
systems since H-bonds
are the
principal
attractive
interaction.
As a
result
of
this
and
because
of the
fundamental interest
of the
water molecule,
ice has

been
extensively
studied.
A
further
point
of
interest
has
been
in the
"centring" transition
where
the
protons reach
the
centre
of the
hydrogen bond
and ice
becomes
a
simple oxide,
"symmetric"
ice X.
Early measurements
of the
hydrogen bond strength using spectroscopic methods
showed
a

strong reduction
in the O-H
vibron indicating
a
weakening
of the
(covalent)
molecular
bond
and a
strengthening
of the
hydrogen bond [17].
In the
absence
of
direct
measurements, estimates
were
made
of the
extension
of the
covalent
O-H
bond length
resulting
from
this weakening. This approach requires
an

assumption
to be
made about
the
changes
in the
potentials
with
pressure.
The
assumption made
was
that
the
double-
well
mean-field potential
for the
H-atom (shown
in the
right-hand plot
of fig. 3)
could
be
described
by the
addition
of two
pressure-independent two-atom potentials (fig.
3,

left-
hand plot) describing
the
interaction
of the
H-atom with
the
donor
and
acceptor oxygen
atoms, respectively. This assumption
of
pressure-independent two-atom potentials
im-
plies
that
as the
H-bond compresses
and the
acceptor atom moves closer
to the
hydrogen
the
attraction
of H by the
acceptor causes
the
covalent
O-H
bond

to
lengthen,
and
this
lengthening
weakens
the O-H
bond
to the
donor oxygen. This model
had
previously been
found
to
describe
well
the
relationships between
O-H and
vibron
frequency
and O • • • O
determined
from
studies
of a
wide
range
of
different

H-bonded materials
at
ambient pres-
sure
[3].
The first
structural study carried
out
with
the
Paris-Edinburgh cell, studies
of
ice
VIII, tested this assumption
and
showed
that
the
intramolecular bond length
was
essentially unchanged
by
pressure
up to at
least 25GPa (fig.
2)
[18, 19]. This lack
of
HJ.YDROGEN BONDING
AT

HIGH
PRESSURE
Two-atom
O-H
potential
361
Atom
A
Donor
8
+
5-
-0.5
0.0 0.5
distance
from
h–bond
centre(A)
Fig.
3. - A
schematic diagram showing
how the
full
H-bond
potential
(right-hand plot)
is
built
up
from

two-atom
O-H
potentials
(left-hand plot) describing
the
interaction between
the El-
atom
and the
donor
and
acceptor oxygen
atoms.
This
approach
and the
assumption
of a
lack
of
change
in the
two-atom
potentials
with
pressure
underlies Klug
and
Whalley's
[17]

estimates
of
the
variation
of the O-H
bond length with
pressure
shown
in fig. 2.
change
in the
bond length implies
that
the
softening
of the
vibron
can be
interpreted
as a
changes
of the
curvature
of the
underlying two-atom
O-H
potentials
—behaviour
which
is

essentially
the
opposite
of
that
which
had
been assumed.
Two
total-energy
studies reproduced
the
observed behaviour
of the O-H
bond length
and
confirmed
this
view
of the
changes
in the
potentials [13, 14].
More
recent
ab
initio molecular dynamics
studies
of ice
also produce

the
observed behaviour. This lack
of
change
in O-H
bond
length with pressure appears
to be a
general feature
in the
0-15
GPa
range:
it is
also
ob-
served
in
ammonia [20], sodium deuteroxide [21], magnesium deuteroxide [22]
and
cobalt
deuteroxide
[23].
4'1.
Ice X. ~ The
experimental observation
of
symmetric
ice X has
been

an
important
goal
since
it was first
postulated
by
Ubbelohode
in
1949 [24].
The
search
for ice X
has led to
extensive revisions
of the ice
phase diagram
in the
very
high pressure region
throughout
the
1990's. Pruzan
et al.
[25] discovered
that
the
transition temperature
of
the

H-bond
ordering transition
from
ice VII to ice
VIII (273
K
from
2-12 GPa) decreases with
increasing pressure
and
that
at ~
60GPa
(70 GPa in
D2O)
it
reaches
OK.
This removed
an
apparent anomaly since
the
behaviour
of
this transition
was
very
different from
that
observed

for
other H-bond ordering
transitions.
In
1996
IR
studies
by
Goncharov
et
al.
[5] and
Aoki
et al.
[26] reported
the first
evidence
of a
symmetrisation
transition
at
~ 75
GPa.
The
manifestation
of the
transition
appeared more complex
than
previously

thought
and
there
has
been some dispute
as to
where
the
transition
occurs (and
as to
what structurally
constitutes
ice X); it was
clear
that
a
major change
in ice
begins
at
this
pressure
and
that
the
transition
to ice X
occurs somewhere
in the

range
75–110
GPa.
Ab
initio modelling
by
Benoit
et al.
[27] also showed
a
symmetrisation
transition
starting
at
similar pressures where
the
volume explored
by the
proton increases
as the
result
of
362 J. S.
LOVEDAY
quantum
effects.
This study
found
an
intermediate

state
where
the
volume explored
by
the
proton
is
increased
by
quantum
effects
which exist
up to ~ 120 GPa
with
a
fully
formed
ice X
above this pressure. Subsequent classical modelling
by
Bernasconi
et al. was
able
to
reproduce
the
experimental
IR
data.

As a
result,
it
appears
that
symmetrisation
occurred progressively
in the
range
65–110GPa
[28].
4'2.
Disorder
in ice
VII.

These revisions
of the
phase diagram have
established
the
importance
of
proton-disordered
ice
VII.
In
addition
to
dominating

the
phase
diagram
at
high pressures,
it is the
phase
which
transforms into
ice X. The
nature
of the
disorder
is,
however,
not
clear.
The
simple model
of ice VII
gives
an O-D
distance
that
is
0.05
A
shorter
than
that

found in
ordered
ice
VIII [29]. Such
a
change
cannot
be
real
(it
would
liberate enough energy
to
melt
the
sample)
and so it has
been assumed
that
the
oxygen
atoms were multi-site disordered. However,
the
model proposed
by
Kuhs
et al.
[29]
—O
displacement along

the
cubic
(100) directions— overcorrected
the O-D
distance
by as
much
as
50%.
More
recent studies [30] based
on
comparison
of the
atomic displacement
(thermal)
parameters
in
ices
VII and
VIII showed
that
displacements
along (111) gave
more plausible internal molecular geometries.
Such
displacements
imply
that
ice VII has

two
different
H-bond lengths
~ 0.1 A
longer
and
shorter than those
of ice
VIII
and
that
this
significant
difference
is
pressure independent
up to at
least
20
GPa.
This
raises
the
question
as to how
such
a
mixed network
will
symmeterise

(a
question
that
remains
to
be
addressed).
The
work
also raises
the
question
as to
what
the
vibrational spectrum
is
probing.
A
simple
view
is
that
two
H-bond
lengths
would imply
a
split
O-H

stretch
peak
which
is not
observed even
in
dilute
H in D2O
experiments
which
probe uncoupled
O-H
vibrations [17]. This suggests
that
the
simple
view
of a
direct
correlation between
H-bond
length
and O-H
stretch
frequency
may be
incorrect. This unexpected disorder
model
also raises
the

question
as to
whether
the
disorder
of the
oxygen atoms
is
driven
by
repulsive
interactions
between
the two
H-bond networks [30].
4'3.
Beyond
ice X. – Two
recent studies suggest
that
ice
will
continue
to
present
challenges
beyond
ice X.
Single-crystal
X-ray

studies
of ice VII by
Loubyere
et al.
[31]
revealed
that
the
structure
has an
incommensurate superlattice
that
persists
across
its
entire range
of
existence
and
into
that
of ice X.
This
superlattice
is not
observed
in
either X-ray
or
neutron powder

diffraction
studies
and has
been postulated
as
some kind
of
partial
ordering
—a
proposal
which
awaits detailed study. Loubeyre
et al.
also
found
evidence
of a
possible
further
structural
transition
at 150 GPa
where Goncharov
et al. [5]
also postulated
a
transition
on the
basis

of a
mode crossing (Fermi resonance).
Ab
initio
molecular-dynamics studies
by
Cavazonni
et al.
[32] explored
the
behaviour
of H2O at
the
high
pressures
and
temperatures
found
within Uranus
and
Neptune. They
found
evidence
for a
dissociation
of the
molecules
and
protonic conduction
that

may be the
source
of the
magnetic
fields of
these planets.
HYDROGEN
BONDING
AT
HIGH
PRESSURE
363
Fig.
4. - The
ordered structure
of
ammonia
phase
IV
[20].
5.

Other
ices
The
hydrides
of
non-metallic elements
are
classed

as
ices; water
ice is the
most studied
of
this class. Studies
of
other systems provide
a
means
to
explore
the
effect
of
changing
hydrogen bond strength
and
H-bond geometry.
5'1. Ammonia.
-
Ammonia
forms
weaker hydrogen bonds than water
and has an
unbalanced geometry
in
that
it has
three donor

H
atoms
and
only
one
lone pair
to
accept H-bonds.
The
high pressure phase diagram
was
explored
by
Gauthier
et al.
[33].
They
found
the
face-centred cubic phase transformed into phase
IV at 3 GPa
with
a
further
transition
at 12 GPa and
then postulated symmetrisation
at 60
GPa. Otto
et

al.
[34]
in
X-ray studies
found
a
hexagonal close-packed nitrogen arrangement between
3 and at
least
30
GPa.
As a
result,
it was
assumed
that
like
the
low-pressure solid
phases
II and
III, phase
IV and
possibly phase
V had
rotationally disordered molecules.
However,
neutron
diffraction
studies showed ammonia

IV to be
orthorhombic with
the
ordered
arrangement
shown
in fig. 4
[20]. Surprisingly,
the
arrangement
has a
bifurcated
hydrogen
bond
in
which
one
hydrogen atom
forms
bonds
to two
nitrogen atoms.
Ab
initio
molecular-dynamics
studies
by
Cavazzoni
et al.
[32]

found
this
structure
to be
stable
to
above
100 GPa
and,
like
ice
VII,
to
become
a
protonic conductor
at
high temperatures
and
pressures.
5"2.
Hydrogen
sulphide.
-
Hydrogen sulphide
has the
same internal molecular geometry
as ice but
much weaker hydrogen bonding;
its

ambient pressure structures
do not
show
evidence
of
hydrogen bonds [35]. High pressure spectroscopy reveals
the
vibron softening
characteristic
of
hydrogen bonding [36]
and at the
highest
pressures
a
blackening
that
suggests
that
metallisation occurs
at 96 GPa
[37]. X-ray
diffraction
studies
at
ambient
temperature
reveal
transitions
at 7

GPa,
11 GPa and 27
GPa,
and
that
metallisation
may
364 J. S.
LOVEDAY
be
the
result
of
short
S-S
contacts
which
are not
H-bond
contacts
[38,39].
The
relationship
between
the
primitive cubic phases
II and I' is
also
of
relevance

to
H-bonding. Both have
related space groups but, while
the
ambient pressure phase
II has a
face-centred cubic
sulphur arrangement [35],
the
sulphur atoms
in
phase
I' are
displaced
by 0.1 A
from
fcc
sites [38]. Neutron
diffraction
studies [40] revealed
that
phase
I' has a
toroidal deuterium
arrangement like phase
II but
that
it is
more ordered,
so

that
the
maxima
in the D
density
point towards
six of the
twelve nearest-neighbour atoms.
The
displacement
of the
sulphur
atoms
from
fee
sites
reduces
the S • • • S
distance
for six
neighbours
and
lengthens
it for the
other
six.
This
arrangement
suggests
the

onset
of
H-bonding
in
phase
I' and the
sharp
transition
from
phase
II to I'
found
at 245 K and 4.5 GPa can be
attributed
to the
onset
of
H-bonding. Modelling studies
by
Rousseau
et al.
also
found
a
similar behaviour [41].
They were
not
able
to
reproduce phase

I' but
found
the
phase
I to IV
transition
to be a
progressive ordering driven
by
H-bonding [41]. Fujhisa
and
co-workers [42] have recently
found
new
phases
in
what
had
been assumed
to be the
stability
field of
phase
IV
below
10 GPa at low
temperatures. These phases
may
also
reflect

the
onset
of
H-bonding.
6.

Hydroxyl
H-bonds
Hydroxyl
H-bonds
are
significantly
different
from
their molecular analogues. They
are
generally weaker
and
more prone
to
bifurcation. Such bonds
are
important
to the
problem
of
water
in the
Earth's
mantle

in
addition
to
their fundamental
interest.
6"1.
Alkali
hydroxides.
-
Potassium
and
sodium hydroxides
sit on the
boundary
of
hydrogen
bonding.
KOH
exhibits hydrogen bonding
that
strengthens with increasing
pressure. NaOH
is
only H-bonded
at low
temperatures [43]
and
spectroscopic studies
show
that

the
transition
to
phase
IV at
high pressure reverses
the
softening
of
vibron [44].
Neutron
diffraction
shows
that
phase
IV has a
bifurcated H-bond
and it
appears
that
the
bifurcation accounts
for the
lack
of
softening
of the
vibron [21].
6"2.
Brucite-structured

hydroxides.
- The
brucite-structured hydroxides
are a
model
system
for
H-bonding
in
hydroxyl-containing systems. They have layered structures
where
the
dominant interactions between
the
metal-oxygen layers
are the
H-bond inter-
action
and
repulsive interactions between
the
hydrogen atoms [23]. Mg(OH)
2
, brucite,
shows
a
softening
of the
vibron with pressure indicating
a

strengthening
of the
hydrogen
bonding [45, 46].
Parise
et al. in
neutron diffraction studies
found
an
intriguing change
in
the
disorder
of the
H(D)
atoms
[22].
The
H(D)
atoms
disordered
over
three
sites
around
a
threefold axis.
As the
pressure
is

increased
in
brucite
the
displacement
of
H(D)
from
the
threefold axis increases. Similar behaviour
is
observed
in
Mn(OD)
2
,
Ni(OD)
2
and
Co(OD)
2
[47].
Raman
and IR
studies
of
Co(OH)2
revealed
that
the

vibron undergoes dramatic broad-
ening
at ~ 11 GPa
[48].
This
broadening
is
very similar
to
that
observed
in
Ca(OH)
2
which
undergoes
pressure
amorphisation
[45]. However, Co(OH)
2
remains
crystalline
in
X-ray
studies
[48].
As a
result, Nyugen [48]
et al.
proposed

that
in
Co(OH)
2
only
the
H-sublattice
amorphises.
However,
Parise
et al.
[23] showed
from
neutron
data
collected
HYDROGEN
BONDING
AT
HIGH
PRESSURE
365
from
Co(OD)2
that
the
occupancy
of the
D-site remained
fully

occupied
and
that
sub-
lattice
amorphisation
did not
occur
up to at
least
16
GPa.
A
detailed examination
of
the
D-site disorder
and the
packing
of the D
layer suggested
that
the
optical anomaly
could
be
explained instead
by
changes
in the

symmetry
of the
D-site.
The
need
to
main-
tain
a D • • • D
distance
of
more
the 1.8 A
forces
the
D-atoms
to
occupy general positions.
This means
that
the
D-atoms have
a
wide
range
of
different
bonding environments that
could
account

for the
broadening
of the
vibron. Recent
ab
initio modelling
of
Ca(OH)2
produces
a
similar sort
of
disorder distribution [49].
7.

Clathrate
hydrates
and
other
water-gas
mixtures
The
behaviour
of
mixtures provides
a
very valuable extension
to
studies
of

single-
component systems. Mixtures provide
a
means
to
probe phenomena like repulsive
in-
teractions
and
mixed H-bonds
that
are not so
readily accessible
and
mixtures
may
yield
analogous structures
that
provide insight into
the
parent single-component systems.
A
classical water-gas mixture
is the
clathrate-hydrate where
the
guest
gas
molecules

sit in
the
centre
of
cages
formed
of
H-bonded water molecules;
the
whole
structure
is
stabilised
by
water-guest repulsions. High pressure studies have revealed
a
number
of
other types
of
mixture.
7'1.
Filled-ice
clathrates.

Small species
like
hydrogen
and
helium

are too
small
to
form
cage clathrates
and the
discovery
that
helium
forms
a
hydrate structure based
on
that
of ice II
caused considerable surprise [50].
Vos et al.
[51] explored
the
hydrogen water
system
and
found
an ice II
related hydrate which appeared
to be
similar
to
helium
hy-

drate
and
above 2.7GPa
a
second hydrate. This second hydrate
has a 1:1
water:hydrogen
ratio
and a
water network like
that
of ice Ic
with hydrogen sitting
in
voids
in the
net-
work.
This structure
is
related
to
that
of ice
VII, which consists
of two
interpenetrating
ice Ic
networks.
H

2
• H
2
O is
approximately
twice
as
compressible
as ice VII and
spec-
troscopic studies suggest
that
the
network
of
H-bonds
may
undergo symmetrisation
at
~ 30 GPa
[52]. Although these mixtures
are
called clathrates, their structures
do not
have
cages
and
resemble
ice
structures very closely.

It is
thus more informative
to
refer
to
them
as
filled
ice
clathrates
or
hydrates.
7'2.
Cage
clathrates.
– The
high pressure behaviour
of
cage clathrates provides
im-
portant information
on
hydrophobic interactions.
In the
cases
of
simple
gas
hydrates
like

those
of
methane, nitrogen, oxygen
and
carbon dioxide
it is
also directly relevant
to
modelling
of the
Earth
and
other planets. They have been extensively studied
in the
0–1 GPa
range; phase transitions have been reported
in
argon, methane
and
nitrogen
hy-
drates [53–56].
However,
very
little
work
had
been carried
out at
pressures above this

and
the
expectation
was
that
they would decompose into guest
and ice at 1 to 2 GPa
[54].
In
the
past
two
years this
view
has
been overturned. Initial indications
of
high pressure
gas
hydrates came
from
Raman studies
of
argon hydrate which showed hydrate phases
stable
to 3 GPa
[56]. X-ray
and
neutron
diffraction

studies
of
methane hydrate revealed
two new
phases [57].
The first is a
hexagonal hydrate stable between
0.8 GPa and 1.9 GPa
with
366 J. S.
LOVEDAY
a
methane: water
ratio
of
3.5(5):1.
This
phase
was
confirmed
in
X-ray single-crystal
studies
by
Chou
et al.
[58].
The
second phase
is

stable between
1.9 GPa and at
least
10
GPa and is an
orthorhombic dihydrate.
The
structure
of
methane dihydrate (fig.
5) is
more like those
of the filled
ices discussed above [59].
It has an
H-bond network
related
to
that
of ice Ih
(the ambient pressure
form
of
ice) with
the
methane molecules contained
in
channels.
The
network

is
somewhat
distorted
compared with
that
of ice Ih in
order
to
expand
the
channels
to
accommodate
the
methane molecules,
but the
network
is
very like
those
of the filled
ices. Hydrogen
and
helium
do not
form
cage clathrates
and
methane
hydrate

is the first
system
which
can be
transformed
from
a
cage clathrate into
a filled
ice
clathrate.
The
existence
of
these hydrates
has
important consequences
for the
modelling
of
Saturn's
moon
Titan
and the
origins
of the
methane
in its
atmosphere.
Titan

accreted
from
a
mixture
of
rock, methane hydrate
and
ammonia monohydrate [60]. Current
models assume
that
all the
methane
was
expelled
from
Titan
early
in its
history
as a
result
of the
assumed pressure decomposition
of
methane hydrate [61]. This resulted
in the
need
to
postulate
some

kind
of
methane reservoir near
the
surface
]—a
methane
ocean
or
methane
in
pores near
the
surface—
since photochemical decomposition
would
have
removed
all the
methane
from
the
atmosphere
in
less than
the
life
of the
Solar
System.

The
stability
of
high pressure methane hydrates means
that
the
methane
may
have
remained within
the ice
mantle
of
Titan
as
methane hydrates
and
that
this
is the
reservoir
supplying
the
atmosphere with methane.
7'3. Ammonia
hydrates.
- The
three ammonia hydrates
are
amongst

the
simplest
systems
to
contain mixed
N • • • O
hydrogen bonds
—such
bonds along with
O • • • O hy-
drogen bonds
are
responsible
for the
base pairings
in
DNA. They
are
likely components
of
the
outer planets. Ammonia monohydrate
is
believed
to
have been
the
dominant
ammonia-bearing phase
in

Titan
and the
assumed waterrammonia
ratio
in
Neptune
and
Uranus
(~
15%) corresponds
to a 1:1
mixture
of
water
and
ammonia dihydrate. Fur-
thermore, ammonia monohydrate
is
predicted
to
ionise
to
form
ammonium hydroxide
at
~ 13 GPa
[62].
Raman studies suggested
that
there

are no
phase
transitions
in
ammonia monohydrate
(AMH)
up to 10 GPa and
that
ammonia dihydrate (ADH)
forms
ice and
ammonia mono-
hydrate
at ~ 5 GPa
[63, 64]. This
was
contradicted
by
dilatometric studies
that
found
phase
transitions
in
both
AMH and ADH at 0.5 GPa
[65]. Neutron
diffraction
studies
of

AMH
revealed
that
there
are
seven phases
up to 6 GPa
[66].
In
general
these
phases have
rather
complex
diffraction
patterns
and
presumably complex
structures.
The
exception
to
this
is
phase
VI,
which
is
formed
by

compression
of AMH to 6 GPa at 170 K and
warming
to
room temperature [67]. This phase
has a
body-centred-cubic arrangement
of
molecular
centres somewhat
like
that
of ice VII
(fig.
6).
However,
the
molecular centres
form
H-bonds
to all
eight nearest neighbours rather than
four
in ice
VII.
The
ammonia
and
water molecules
are

substitutionally disordered
so
that
each molecular centre
is 50%
occupied
by
water
and
ammonia.
AMH-VI
is
thus
a
type
of
material:
a
hydrogen-bonded
molecular alloy (see
fig. 6).
There
is
also evidence
of
repulsive
effects
like those
found
in

HYDROGEN
BONDING
AT
HIGH
PRESSURE
(a)
MH-III
367
(a)
MH-I
Fig.
5. -
Left:
the
structure
of (a)
methane dihydrate (MH-III)
and (b) ice Ih
viewed perpen-
dicular
to
their c-axes.
The
smaller spheres
are O
atoms
of the
water network
and the
lines

denote H-bonds.
The
larger spheres
in (a) are the
methane molecules. MH-III
is
viewed approx-
imately along
its
a-axis
and ice Ih
approximately along
a
[110] direction. Right:
the
structure
of
(c)
MH-III
and (d) ice Ih
viewed parallel
to
their c-axes.
The + and —
symbols show
the
sense
of
c-axis H-bonds
from

the
puckered sheet labelled
S in
(a).
Fig.
6. - The
structure
of
AMH-VI.
368 J. S.
LOVEDAY
Co(OD)
2
with about
20% of the
deuterium density being directed along (110)
directions.
This
can be
explained
by the
need
to
avoid short
D • • • D
contacts.
The
substitutional
disorder
of

AMH-VI
and its
similarity
to ice VII
raises
the
possibility
that
it
forms
a
solid
solution with
ice VII so
that
the
relevant phase
for the
interiors
of
Uranus
and
Neptune
may
be a
water-rich variant
of
AMH-VI.
8.


Summary
As
a
result
of
recent developments,
the
depth
of our
understanding
of
hydrogen bond-
ing
at
high pressure
has
been greatly enhanced.
The use of
combined modelling
and ex-
perimental techniques
is
clearly
an
exciting development,
which
is
likely
to
prove valuable

for
tackling complex H-bonded systems. Such complex systems
are one of the
current
grand challenges
for
high-pressure studies
of
hydrogen bonding.
* * *
I
would like
to
thank
R.
NELMES
and R. J.
HEMLEY
for
reading this manuscript
and for
their
helpful
suggestions.
I
also acknowlege
the
support
of the
Engineering

and
Physical Sciences Research Council,
and of the
ISIS neutron
facility
at the
Rutherford
Appleton Laboratory.
REFERENCES
[1]
LUNINE
J. I. and
STEVENSON
D. J.,
Icarus,
70
(1987)
61.
[2]
JEFFREY
G. A.
(Editor),
An
Introduction
to
Hydrogen
Bonding (OUP, Oxford) 1997.
[3]
OLOVSSON
I. and

JCWSSON
P. G.,
X-ray
and
Neutron
Diffraction
Studies
of
Hydrogen
Bonded
Systems,
in The
Hydrogen
Bond, edited
by
SCHUSTER
P. W.,
ZUNDEL
G. and
SANDORFY
C.
(North Holland, Amsterdam) 1976,
pp.
393-456.
[4]
AOKI
K.,
USUBA
S.,
YOSHIDA

M.,
KAKUDATE
Y.,
TANAKA
K. and
FUJIWARA
S., J.
Chem. Phys.,
89
(1988) 529.
[5]
GONCHAROV
A. F.,
STRUZHKIN
V. V.,
SOMAYAZULU
M. S.,
HEMLEY
R. J. and MAO
H. K.,
Science,
273
(1996) 218.
[6]
GONCHAROV
A. F.,
STRUZHKIN
V. V., MAO H. K. and
HEMLEY
R. J.,

Phys. Rev. Lett.,
83
(1999) 1998.
[7]
KLOTZ
S., Z.
Krist.,
216
(2001) 420.
[8]
OCCELLI
F.,
KRISCH
M.,
LOUBEYRE
P.,
SETTE
F., LE
TOULLEC
R.,
MASCIOVECCHIO
C. and
RUEFF
J. P.,
Phys. Rev.
B, 63
(2001)
4306.
[9]
LI J. C.,

BURNHAM
C.,
KOLESNIKOV
A. I. and
ECCLESTON
R. S.,
Phys. Rev.
B, 59
(2001)
9088.
[10]
MAO H. K., KAO C. C. and
HEMLEY
R. J., J.
Phys. Condens. Matter,
13
(2001) 7847.
[11] BESSON
J. M.,
NELMES
R. J.,
HAMEL
G.,
LOVEDAY
J. S.,
WEILL
G. and
HULL
S.,
Physica

B, 180 & 181
(1990)
90.
[12] NELMES
R. J.,
LOVEDAY
J. S.,
WILSON
R. M.,
BESSON
J. M.,
KLOTZ
S.,
HAMEL
G.
and
HULL
S.,
Trans.
Am.
Cryst. Ass.,
29
(1993)
19.
[13]
OAJMAE
L.,
HERMANSSON
K.,
DOVESI

R. and
SAUDERS
V. R., J.
Chem. Phys.,
100
(1994)
2128.
HYDROGEN
BONDING
AT
HIGH
PRESSURE
369
[14]
BESSON
J. M.,
PRUZAN
P.,
KLOTZ
S.,
HAMEL
G,,
SILVI
B.,
NELMES
R. J.,
LOVEDAY
J. S.,
WILSON
R. M. and

HULL
S.,
Phys.
Rev.
B, 49
(1994)
12540.
[15]
CAR R. and
PARRINELLO
M.,
Phys.
Rev.
Lett.,
55
(1985)
2471.
[16]
MARX
D. and
PARRINELLO
M., J.
Chem. Phys.,
104
(1996)
4077.
[17]
KLUG
D. D. and
WHALLEY

E., J.
Chem. Phys.,
81
(1984)
1220.
[18]
NELMES
R. J.,
LOVEDAY
J. S.,
WILSON
R. M.,
BESSON
J. M.,
KLOTZ
S.,
HAMEL
G.
and
HULL
S.,
Phys.
Rev.
Lett.,
71
(1993)
1192.
[19]
NELMES
R. J.,

LOVEDAY
J. S.,
MARSHALL
W. G.,
BESSON
J. M.,
KLOTZ
S. and
HAMEL
G.,
Rev.
High
Press.
Sc^.
Technol.,
7
(1998)
1138.
[20] LOVEDAY
J. S.,
NELMES
R. J.,
MARSHALL
W. G.,
BESSON
J. M.,
KLOTZ
S. and
HAMEL
G.,

Phys.
Rev.
Lett.,
76
(1995)
174.
[21] LOVEDAY
J. S.,
MARSHALL
W. G.,
NELMES
R. J.,
KLOTZ
S.,
HAMEL
G. and
BESSON
J. M., J.
Phys. Condens. Matter,
8
(1996) L597.
[22] PARISE
J. B.,
LEINENWEBER
K.,
WEIDNER
D. J TAN K. and
VON
DREELE
R. B., Am.

Min.,
79
(1994)
193.
[23] PARISE
J. B.,
LOVEDAY
J. S.,
NELMES
R. J. and
KAGI
H.,
Phys.
Rev.
Lett.,
83
(1999)
328.
[24] UBBELOHODE
A. R., J.
Chim. Phys.,
46
(1949)
429.
[25]
PRUZAN
P.,
CHERVIN
J. C. and
CANNY

B., J.
Chem. Phys.,
99
(1993) 9842.
[26] SONG
M.,
YAMAWAKI
H.,
FUJIHISA
H.,
SAKASHITA
M. and
AOKI
K.,
Phys.
Rev.
B, 60
(1999)
12644.
[27] BENOIT
M.,
MARX
D. and
PARRINELLO
M.,
Nature,
392
(1998)
258.
[28] BERNASCONI

M.,
SILVESTRELLI
P. L. and
PARRINELLO
M.,
Phys.
Rev.
Lett.,
81
(1998)
1235.
[29]
KUHS
W. F.,
FINNEY
J. L.,
VETTIER
C. and
BLISS
D. V., J.
Chem. Phys.,
81
(1984)
3612.
[30] NELMES
R. J.,
LOVEDAY
J. S.,
MARSHALL
W. G.,

HAMEL
G.,
BESSON
J. M. and
KLOTZ
S.,
Phys.
Rev.
Lett.,
81
(1998)
2719.
[31]
LOUBEYRE
P.,
LETOULLEC
R.,
WOLANIN
E.,
HANFLAND
M. and
HAUSERMANN
D.,
Nature,
397
(1999)
503.
[32]
CAVAZZONI
C.,

CHIAROTTI
G. L.,
SCANDOLO
S.,
TOSATTI
E.,
BERNASCONI
M. and
PARRINELLO
M.,
Science,
283
(1999)
44.
[33]
GAUTHIER
M.,
PRUSAN
P.,
CHERVIN
J. C. and
BESSON
J. M.,
Phys.
Rev.
B, 37
(1988)
2102.
[34]
OTTO

J. W.,
PORTER
R. F. and
RUOFF
A. L., J.
Phys. Chem.
Solids,
50
(1989)
171.
[35]
COCKCROFT
J. K. and
FITCH
A. N., Z.
Krist.,
193
(1990)
1.
[36]
SHIMIZU
H.,
YAMAGUCHI
H.,
SASAKI
S.,
HONDA
A.,
ENDO
S. and

KoBAYAsm
M.,
Phys.
Rev.
B, 51
(1995)
9391.
[37] SAKASHITA
M.,
YAMAWAKI
M.,
FUJIHISA
H.,
AOKI
K.,
SASAKI
S. and
SHIMIZU
H.,
Phys.
Rev. Lett.,
79
(1997) 1082.
[38]
FUJIHISA
H.,
YAMAWAKI
H.,
SAKASHITA
M. and

AOKI
K.,
Phys.
Rev.
B, 57
(1998)
2651.
[39] ENDO
S.,
HONDA
A,
KOTO
K.,
SHIMOMURA
O.,
KIKEGAWA
T. and
HAMAYA
N.,
Phys.
Rev.
B, 57
(1998) 5699.
[40] LOVEDAY
J. S.,
NELMES
R. J.,
KLOTZ
S.,
BESSON

J. M. and
HAMEL
G.,
Phys.
Rev.
Lett.,
85
(2000)
1024.
[41]
ROUSSEAU
R.,
BOERO
M.,
BERNASOONI
M.,
PARRINELLO
M. and
TERAKURA
K.,
Phys.
Rev.
Lett.,
83
(1999)
2218.
[42]
FUJIHISA
H.,
unpublished.

[43] BATOW
T. J.,
ELCOMBE
M. M. and
HOWARD
C. J.,
Solid
State Commun.,
57
(1986)
339.
370 J. S.
LOVEDAY
[44]
KROBOK
M. P.,
JOHANNSEN
P. and
HOLZAPFEL
W. B., J.
Phys. Condens. Matter,
4
(1992) 8141.
[45] KRUGER
M. B.,
WILLIAMS
Q. and
JEANLOZ
R., J.
Chem. Phys.,

91
(1989) 5910.
[46]
DUFFY
T. S.,
MEADE
C., FBI Y. W., MAO H. K. and
HEMLEY
R. J., Am.
A/in.,
80
(1995) 222.
[47]
AOKI
H.,
SYONO
Y. and
HEMLEY
R. J.
(Editors),
Physics
Meets
Mineralogy
(CUP,
Cambridge) 2000,
pp.
308-322.
[48] NGUYEN
J. H.,
KRUGER

M. B. and
JEANLOZ
R.,
Phys. Rev. Lett.,
78
(1997) 1936.
[49] RAUGEI
S.,
SILVESTRELLI
P. L. and
PARRINELLO
M.,
Phys. Rev. Lett.,
83
(1999)
2222.
[50] LONDONO
D.,
FINNEY
J. L. and
KUHS
W. F., J.
Chem. Phys.,
97
(1992) 547.
[51]
Vos W. L.,
FINGER
L. W.,
HEMLEY

R. J. and
MAO
H. K.,
Phys. Rev. Lett.,
71
(1993)
3150.
[52]
Vos W. L.,
FINGER
L. W.,
HEMLEY
R. J. and
MAO
H. K.,
Chem. Phys.
Lett.,
257
(1996) 524.
[53]
DYADIN
Y. A.,
LARIONOV
E. G.,
MIRINSKI
D. S.,
MIKINA
T. V. and
STAROSTINA
L. I.,

Mendeleev
Commun.,
7
(1997)
32.
[54]
DYADIN
Y. A.,
ALADKO
E. Y. and
LARIONOV
E. G.,
Mendeleev
Commun.,
7
(1997)
34.
[55]
VAN
HINSBERG
M. G. E.,
SCHEERBOOM
M. I. M. and
SCHOUTEN
J. A., J.
Chem. Phys.,
99
(1993)
752.
[56] LOTZ

H. T. and
SCHOUTEN
J. A., J.
Chem. Phys.,
Ill
(1999) 10242.
[57]
LOVEDAY
J. S.,
NELMES
R. J.,
GUTHRIE
M.,
BELMONTE
S. A.,
ALLAN
D. R.,
KLUG
D. D., TSE J. S. and
HANDA
Y. P.,
Nature,
410
(2001) 661.
[58]
CHOU
I. M.,
SHARMA
A.,
BURRUSS

R. C., SHU J., MAO H. K.,
HEMLEY
R. J.,
GONCHAROV
A. F.,
STERN
L. A. and
KIRBY
S. H.,
Proc. Natl. Acad. Sci.,
97
(2000)
13484.
[59]
LOVEDAY
J. S.,
NELMES
R. J.,
GUTHRIE
M.,
KLUG
D. D. and TSE J. S.,
Phys. Rev.
Lett.,
87
(2000) 215501.
[60] LUNINE
J. I. and
STEVENSON
D. J.,

Icarus,
70
(1987)
61.
[61]
LUNINE
J. I. and
STEVENSON
D. J.,
Astrophys.
J.
Suppl.
Ser.,
58
(1985)
493.
[62] JOHNSON
D. A., J.
Chem. Soc. Dalton, 1988 (445) 1988.
[63]
KOUMVAKALIS
A.,
Ph.D.
Thesis,
UCLA
(1988).
[64]
CYNN
H. C.,
BOONE

S.,
KOUMVAKALIS
A.,
NICOL
N. and
STEVENSON
D. J.,
Proc. 19th
Lunar
and
Planetary Science
Conf.,
19
(1989)
433.
[65] HOGENBOOM
D. L.,
KARGEL
J. S.,
CONSOLMAGNO
G. J.,
HOLDEN
T. C., LEE L. and
BUYYOUNOUSKI
M.,
Icarus,
128
(1997) 171.
[66] LOVEDAY
J. S. and

NELMES
R. J.,
Science
and
Technology
of
High
Pressure,
edited
by
MANGHNANI
M.,
NELLIS
W. and
NICOL
M.,
Vol.
1
(Universities
Press,
Hyderabad, India)
2000,
p.
133.
[67] LOVEDAY
J. S. and
NELMES
R. J.,
Phys. Rev.
Lett.,

83
(1999) 4239.
CHEMISTRY
AND
BIOLOGY
This page intentionally left blank
High
pressure organic synthesis: Overview
of
recent applications
G.
JENNER
Laboratoire
de
Piezochimie
Organique
(UMR 7123), Institut
de
Chimie
Universite
Louis Pasteur
- 1 rue
Blaise Pascal, 67008
Strasbourg,
France
1. —
Introduction
Synthesis
is a
major concern

in
organic chemistry.
The
creation
of new
molecules
by
new
chemical routes
and new
activation processes highlights
the
power
of
organic
chemistry.
Two
major objectives must constantly
be
kept
in
mind: yield
and
selectivity.
It is
evident
that
chemical synthesis
is
optimal when highest yield

and
best selectivity
are
obtained. This means
that
the
reaction must:
-
proceed
at a
reasonable
rate
-
fulfill
precise criteria with respect
to
chemo-, regio-, stereo-, enantio-selectivity.
Yields
are
conditioned
by a
number
of
parameters depending
on
activation
modes.
These
are
various

and may be
divided essentially into physical (temperature, pressure,
light)
and
chemical (catalysis) activation methods. Pressure activation
is,
basically,
not
a new
technique although
the first use of
this
parameter
in
chemistry
dates
back
from
1892 only [1]. Sporadic reports
on
high pressure synthesis were published [2]; however,
the
technique became popular mostly
in the
last
twenty years.
The
fundamental
effect
of

pressure
in a
chemical reaction considers
its
action upon
the
rate
constant
k
according
to
transition
state
theory:
dP
T
'
©
Societa
Italians
di
Fisica
373
374
G.
JENNER
200 400 600 800
1000 1200
pressure
/ MPa

Fig.
1. -
Pressure
acceleration
of
rate
constants.
Ay* is the
activation volume.
It is
stricto sensu
the
difference
in
partial
molar volumes
of
transition
state
and
reactants. This kinetic parameter
is, in
fact,
the
basic parameter
to be
considered
for
synthetic purposes.
It is

clear
that
organic synthesis under high
pressure
is
useful
to
consider
if:
- the
sign
of AV* is
negative (the reaction
is
accelerated
by
pressure),
- the
magnitude
of
|AV*|
is
highest.
Figure
1
gives
a
quantitative idea
of the
pressure acceleration

of
rate
constants
for two
values
of
AV*.
A
number
of
name
reactions
have been
investigated
under
pressure
and
their
activa-
tion volumes determined. Table
I
lists representative values.
Such
values take into account
the
volume variations resulting
from
molecular reorgani-
zation (bond cleavage
and

bond
formation)
and
electrostriction (compression
of
molecules
by
vicinal charged species inducing volume shrinkage).
In
table
I the
most negative values
of
AV*
experience major volume contractions. Michael, Menshutkin
and
Morita-Baylis-
Hillman
reactions involve
the
formation
of
only
one
bond with
a
subsequent volume
contribution
of
about –20cm

3
mol
–1
at the
maximum.
The
additional volume value
is
ascribed
to
solute-solvent interactions
which
are
overwhelming
in all
ionogenic reactions.
Other
possible volume
effects
can
result
from
steric interactions since
the
pressure
sensitivity
of
reaction
rate
was

observed
to
increase with higher steric congestion [3].
In
the
last
few
years numerous investigations have
confirmed
these results; this, obviously,
should
stimulate
use of
pressure
to
force
reluctant sterically congested substrates
to
react [4].
In
conclusion, high pressure organic synthesis
is
particularly
useful
for
reactions
in-
volving:
HIGH
PRESSURE

ORGANIC
SYNTHESIS:
OVERVIEW
OF
RECENT
APPLICATIONS
375
-
formation
of one or
more bonds,
-
generation
of
charged species,
-
steric hindrance.
Considering
high pressure
as
activation mode,
the
best yields
are
obtained when
the
rate constant
is not too low at
ambient pressure (though notable exceptions
are

known;
see
tables VII, VIII reporting
no
reaction
at
ambient pressure
and
excellent yields
at
high
pressure
in
ionogenic reactions)
and the
activation volume
is as
negative
as
possible
(fig.
1)-
An
important cautionary remark should
be
made
as
pressure modifies
the
physical

properties
of the
liquid molecular system.
-
Pressure increases
the
solubility
of
solids
and
miscibility
of
liquids
in any
medium.
This
is
important
as it may
influence
the
homogeneity
of the
medium.
-
Pressure increases
the
viscosity
of all
liquids

in an
exponential way.
At
very high
pressures viscosity
can be so
high
that
diffusion
processes become rate-limiting
meaning
that
bimolecular rate constants
may
decrease.
-
Pressure increases melting points. Most solvents
are
solid
at
room temperature
under
a
pressure
as
high
as
1000 MPa.
A
cursory estimation

of the
solidification
point
can be
made
from
application
of the
equation
of
Simon
and
Glatzel [5]:
a, c :
constants
TO :
critical temperature.
TABLE
I. -
Experimental activation volume values
for
given reactions.
Reaction
AV*
(cm
3
mol
–1
)
Concerted

sigmatropic
rearrangements
—8 to —18
Polymerization
(propagation
step)
—15 to
—20
Wittig
addition
-20 to -30
Michael
addition
—20 to —50
Diels-Alder cycloaddition
— 25 to —40
Concerted
ene
reaction
—30 to —45
Menshutkin reaction
—30 to —50
Morita-Baylis-Hillman
reaction
—40 to —70
376 G.
JENNER
TABLE
II. -
Calculated

melting
points
(in
°C).
Compound
Acetonitrile
Ethanol
Diethyl
ether
Dichloromethane
Chloroform
Carbon
tetrachloride
Dioxan
Ethyl
acetate
Chlorobenzene
Nitromethane
Nitrobenzene
Cyclohexane
at 0.1 MPa
-43.9
-117.3
-116.0
-96.7
-63.5
-22.6
-0.2
-83.6
-45.5

-28.6
5.6
6.5
at 300 MPa
12.4
-91.2
-71.0
-64.7
-16.5
76.6
29.8
-52.6
3.5
6.4
69.8
131.0
at 500 MPa
44.1
-75.0
-40.0
-44.6
12.5
128.1
46.5
-34.6
30.3
45.4
107.0
at
1000

MPa
111.6
-37.7
20.6
2.2
79.1
81.4
4.5
85.7
111.4
Some
useful
calculated
Tp
values
for
common solvents
are
listed
in
table
II. The
importance
of the
liquid
state
has
recently been highlighted
in the
Henry addition

of
nitromethane
to
2-butanone [6].
The
nitroalcohol yield
at 750 MPa is 60%
when
the
ketone
serves
as
solvent,
but
only
9%
under identical conditions with nitromethane
as
reaction medium
due to the
solidification
of the
nitro compound.
It is,
therefore,
necessary
to
ensure
the
liquid

state
of the
reactional system
at the
working pressure
and
temperature.
2.

Recent
applications
2'1.
Cycloadditions.
-
Cycloadditions
are
typical examples
of
pressure-accelerated
reactions, particularly those showing reluctance
to
occur
at
ambient pressure
due to
steric
B
CN
R,
DR

H
N
R,
-CN
-OR
Fig.
2. -
Synthesis
of
l-alkoxy-2,2-dicyanocyclobutanes.
HIGH
PRESSURE
ORGANIC
SYNTHESIS:
OVERVIEW
OF
RECENT
APPLICATIONS
377
R
i
OR
H
R
1
+
H-
-N=C=O
COOMe
CHR'

COOMe
Fig.
3. -
Synthesis
of
ßlactams
via
high pressure
[2 + 2]
cycloaddition
of
enol
ethers
and
isocyanates.
hindrance
or for
electronic reasons.
In
this section
we
will
give
a
nonexhaustive overview
of
recent applications
in
this
field

encompassing
different
types
of
cycloadditions.
21.1.
[2 + 2]
Cycloadditions.
The
absence
of
concertedness
for [2 + 2]
cycload-
ditions implies moderately negative activation volumes related
to the
formation
of one
bond
in the
transition
state
(about
—15
to

20cm
3
mol
-1

). However, depending
on
sub-
strates
the
transition
state
can be
more polar than
the
initial
state
in
such
a way
that
electrostrictive
effects
generate
an
additional
volume
term
making such
reactions
fairly
to
strongly pressure sensitive.
As an
example, enol ethers

add to
1,1-dicyanoalkenes
to
afford
l-alkoxy-2,2-dicyanocyclobutanesat 1200
MPa
(fig.
2)
[7].
The
reaction
is
particularly
adapted
for
sterically hindered enol ethers.
If
R
1
= R
2
= R = Me, the
yield
is
80%,
if R
I
= R
2
= Me and R = Et, the

yield
is 90%
(no
reaction
at 0.1 MPa in
both cases). Even, silyl enol ethers
can be
used
in
uncat-
alyzed
reactions
at
high pressures where only
low
yields
of
cyclobutanes
are
obtained
at
normal pressure
in the
presence
of
Lewis acid
catalysts.
An
interesting application
of

[2+2] cycloadditions concerns
the
synthesis
of
ß-lactams.
The
simplest route involves
[2 + 2]
cycloadditions
of
imines derived
from
aminoacids
and
ketenes. High pressure promotes addition reactions
of
enol ethers
and
isocyanates
(fig.
3)
[8].
The
cycloadditions
are
completely regio-selective adding
to the
utility
of the
high

pressure process.
In
the
same way,
2,3-dihydrofuran
reacts with phenyl isocyanate
at 100 °C
under high
pressure. Eighty percent yield
of the
corresponding
ß-lactam
is
obtained
at 800 MPa
[9].
A
related reaction concerns
the [2 + 2]
cycloaddition
of
2,3-dihydrofuran
to
Schiff
bases
(fig.
4)
[10].
The
reaction

is
extremely sluggish
at 0.1
MPa.
It is
promoted
by
pressure
800 MPa
N'
O
V
T/
160°C
Fig.
4. -
High
pressure synthesis
of
azetidines.
378
G.
JENNER
Fig.
5. -
Diels-Alder
reaction
of
pyridones
with

cyclooctyne.
in
virtue
of the
formation
of one
bond
and a
zwitterionic intermediate
in the
transition
state.
Yields
of
azetidines
are
modest
to
good.
2'1.2.
[4 + 2]
Cycloadditions. These reactions
are
generally concerted (simultane-
ous
formation
of two
bonds
in the
transition

state).
Consequently,
the
pressure
effect
is
considerable. Many recent examples take advantage
of
this
mechanistic property,
partic-
ularly
in
heterocyclic chemistry.
2(lH)-pyridones
show
poor reactivity
as
dienes
due to
their
partial
aromaticity.
Ac-
tivation
by
pressure permits
to
remove their reactional lethargy. Thus, alkyl substituted
2(lH)-pyridones

react with cyclooctyne
to
give stable bridged cycloadducts (20–80%
at
800MPa, 90°C,
10
days) (fig.
5)
[11].
Furans
show
strong reluctance
to
enter cycloaddition.
The low
reactivity
is
ascribed
to the
easy retro-Diels-Alder process
where
the
aromaticity
of the
diene
is
recovered.
High
pressure
is an

efficient
way to
shift
the
equilibrium toward formation
of
oxabi-
cyclo[2.2.1]heptane derivatives.
In the
last
years numerous successful syntheses were
reported
yielding intermediates
for the
synthesis
of
important drugs.
The
high pres-
sure
(1500 MPa) addition
of
some substituted
furans
to
cyclopenten-2-enones
affords
cycloadducts
which
can be

used
for
further
synthesis
of
highly
functionalized hydrinde-
nones.
These compounds
are key
intermediates
for the
preparation
of
ottelione
A, a
potent inhibitor
of
tubulin polymerization (fig.
6)
[12].
At
normal pressure Lewis acid
catalysis leads
to
Michael adducts only.
Palasonin
is an
inhibitor
of

phosphorylation
of
proteins.
Its
total
synthesis
can be
OMe
hydrindenones
OMe
Fig.
6. -
Cycloaddition
of
furans
to
cyclopenten-2-ones.
HlGH
PRESSURE
ORGANIC
SYNTHESIS:
OVERVIEW
OF
RECENT
APPLICATIONS
379
O
Me
Fig.
7. -

Two-step
synthesis
of
palasonin.
effected
very
efficiently
by
high pressure cycloaddition
of
citraconic anhydride
to
furan
followed
by
hydrogenation (fig.
7)
[13].
The
reaction
was
also used
for the
partial
synthesis
of a
complex molecule, paclitaxel.
The
CD-ring
is

known
for its
antineoplastic activity. Addition
of
citraconic anhydride
to
i)
pressure
O"
D
MeOH
O
Fig.
8. -
Synthesis
of
precursors
of
paclitaxel.
380
G.
JENNER
1000
MPa
Fig.
9. -
Intramolecular
[4 + 2]
cycloaddition
of

tethered
furans.
2-methylfuran
at
1500
MPa
affords
two
stereoisomers (1:1) (90%)
which
are
immediately
hydrogenated
in
order
to
avoid reverse
reaction.
In the
same way,
other
furans
and
maleic
anhydrides
can be
brought
to
reactivity. Figure
8

shows
the
high pressure synthesis
of
1 and 2 as
potential CD-ring precursors which
can be
used
in the
total
synthesis
of
paclitaxel analogues [14].
It is
interesting
to
note
that
not
only cycloaddition
is
favored
by
pressure,
a
further
step—acid-catalyzed ether cleavage
of
esters
and

lactones—is also
promoted
by
application
of
pressure.
Intramolecular Diels-Alder reactions
of
furans
are
also strongly accelerated
by
pres-
sure.
Complex structures have been obtained
by
reacting
furans
tethered
by
bicyclo-
propylidenes
(fig.
9)
[15].
A
key
step
in the
synthesis

of
brassinosteroids (hormons
for
vegetal growth)
is the
intramolecular
[4 + 2]
cycloaddition
of
enone
3.
Whereas cyclisation
of 3 (R = H)
occurs
almost spontaneously, only high pressure (1000 MPa)
is
able
to
bring enone
3 (R = Me)
to
reactivity (fig.
10)
[16].
The
aromaticity
of
pyrroles precludes Diels-Alder reactivity.
The 7-
azabicyclo[2.2.1]heptan

skeleton
can be
constructed
in one
step
via [4 + 2]
cycloaddition
of
activated pyrroles with dienophiles. High pressure activation
permits
to
extend
largely
the
scope
of
these reactions [17].
A
particularly interesting application
is the
synthesis
of an
analogue
of
epibatidine
which
is a
potent analgesic. This
is
achieved

by
high pressure reaction
of
l-methoxycarbonyl-3-phenylthiopyrrole
with
phenyl
vinyl
sulfone
(fig.
11)
[18].
Indoles
react
as
dienophiles only under extreme conditions. Activation
by
high
pres-
sure
and
Lewis
acid catalysis
is a
straightforward
way to
enhance
the
dienophilicity
of
O

O
Fig.
10. -
Intramolecular
Diels-Alder
reaction
of 3.
R
= H
yield: 100% (0.1 MPa)
R
= Me
yield:
53%
(1000
MPa)

×