You are on page 1of 17

72 Chapter 2.

Crystal Structures of Minerals

2.1.7 Precision Electron Density Calculations: Relation to


Chemical Bonding and to Localization of Impurities in
Crystal Structures

V.G. TSIRELsoN and O.V. FRANK-KAMENETSKAYA

There are two ways to obtain the electron density of minerals. The first is to
calculate electron density by quantum chemistry methods, the second is by
X-ray diffraction measurement. In the latter case, a special experimental
procedure is needed and the accurate single-crystal X-ray diffraction analysis is
used for this purpose. The total X-ray intensities are measured with an accuracy
of about I % and secondary diffraction effects such as absorption, extinction,
thermal diffuse scattering, and anomalous dispersion are appropriately correc-
ted. As a result, moduli of structure amplitudes are obtained. For the extraction
of crystal chemical information from X-ray data, the values of structure
amplitudes are approximated by their theoretical analogs, calculated from the
model of a given mineral. The refined parameters of the model are atomic
coordinates, harmonic and anharmonic characteristics of atomic thermal vibra-
tions, and populations of valence electron shells of atoms. In this way the atomic
charges in many minerals have been determined.
The electron density in the unit cell can be mapped by Fourier series using
structure amplitudes as coefficients (Fig. 4). The phases of structure amplitudes
are calculated by means of a crystal model. The "experimental" electron density
maps have a resolution of 0.2 A and they are the basis for studying chemical
bonding in minerals. The deformation electron density (OED) is usually
calculated, in order to obtain characteristics of a chemical bond. OED is the

Fig. 4. Electron density distribution


in fluorite CaF 2 as observed by x-
ray diffraction. Positions of maxima
are atom sites
2.1.7 Precision Electron Density Calculations 73

difference between crystal electron density restored from the X-ray diffraction
data and electron density of the ensemble of spherically symmetric atoms
calculated by quantum chemistry methods. DED describes the redistribution of
electrons during crystal formation from atoms.
There are general principles for the interpretation of DED maps. The excess
of DED appears as peaks on these maps in regions of bonding overlap of the
atomic orbitals and in the regions of the orbitals of electron lone pairs location.
Negative DED is connected with anti-bonding orbital overlap regions, and with
the regions where bonding overlap of atomic orbitals are subjected to weaker
changes than mutual penetration of atomic clouds. The DED peaks connected
with n-bonds are elongated in the internuclei space perpendicular to the line
linking the interaction atoms. In ionic bonds, DED peaks are displaced along the
bond line to more electronegative atoms and in strained bonds these peaks are
shifted from bond line.
An example of the significance of electron density studies for minerals is the
rutile-type high pressure porymorph of silica stishovite SiOz' In this mineral,
Si is coordinated by six oxygen atoms in an octahedral configuration. The
structure contains two nonequivalent Si- O bonds: the longer (equatorial) bonds
are 1.757 A. and are about 3% shorter than axial bonds; but the ionic model
predicts that ratio between Si- Oe and Si- O. bond length should be more than
unity. Furthermore, non bonding 0 - 0 edges of Si- octahedra (2.29 A.) are

a b

Fig. Sa, h. Experimental deformation electron density (the difference between measured elec-
tron density for crystal and sum of calculated atomic electron densities) for stishovite Si0 2 . a
Map of the octahedron axial/equatorial bond plane. b Map of the coordination plane of the
oxygen atom. The contour interval is 0.05 eA - 3. Excess of electron density (dark lines) on the
bond lines is responsible for chemical bond forming: electron density peaks are shifted to more
electronegative oxygen atoms and elongated perpendicular to the bond line (n-character of the
bond). Negative electron density (dashed lines) characterizes the region from which the electron
is removed when the crystal is formed from atoms
74 Chapter 2. Crystal Structures of Minerals

shorter than the double oxygen radii (2.54-2.80 A). Thus, the real structure
contradicts the Pauling rule of shortening the shared edges of the polyhedra in an
ionic model. The DED maps (Fig. 5) reveal the bonding peaks with a height of
0.47 eA -3 on the Si-O. bonds and of 0.29 eA -3 on the Si-O. bonds. These
peaks are shifted toward oxygen atoms (partly ionic covalent u-bond) and shifted
from the bond line (some strain in the four-membered ring). The maxima in the
equatorial plane are linked by "bridges" of electron density. These bridges have a
shielding effect on the cation interaction across the shared edge, and reduce the
repulsive forces between Si atoms. Thus, electron density study explains the
origin of the Pauling rule violation. Note that the average distance between the
Si atom and the nearest DED peak is 1.10 A, which is close to the Slater atomic Si
radii of 1.15 A..
The observation of splitting of 3d-element electron energy levels in a crystal
field via electron density maps can be demonstrated for high-pressure phase
y-M 2 Si0 4, where M = Fe, Co, and Ni. These minerals belong to normal spinels
in which transition-metal atoms are situated in the centers of oxygen octahedra
(B-position) with symmetry position D 3d • Eight positive peaks 0.9-1.0 eA -3 in
height located at the corners of a cube have been revealed on the DED map for y-
Ni 2Si0 4 near a B-position (Fig. 6). The octahedral crystal field symmetry in the
position of a Co atom is more trigonally distorted, and peaks ofDED lying along
the threefold axis of the octahedron are higher. A further trigonal distortion of
the Fe-octahedron leads to the presence of excess DED peaks only along the
threefold axis. These DED maps may be interpreted in terms of the crystal field
theory as follows. The trigonal distortion of octahedron in this spinels is its
compression along the threefold axis [111]. This compression leads to the
splitting the 3d-electron energy levels of B-cations into a singlet ag with
corresponding orbital lies along the threefold axis, and two doublets: eg and e~
(the latter are described by n-orbitals lying in the plane perpendicular of the
threefold axis). The level a g is the lower in energy than e~. In the near perfect
octahedral environment, six of the eight 3d-electrons of the NF + must be
occupied ag and e~ orbitals and the remaining two must be occupied eg-orbitals.
Thus, Ni 2 + electronic configuration (ag)2(e~)4(eg)2 may be expected. Excess and
deficit of the electrons on corresponding orbitals as compared to the spherical
atom average orbital populations, equal to 8/5 electrons, are reflected in the
formation of positive and negative regions around Ni atoms on the DED map.
The position ofDED minima and maxima correspond to the crystal field theory
picture. As the number of 3d-electrons of the B-cation in a series Ni-Co-Fe
decreases, geometrical distortions in the structure increase, and the electron
density around the B-cation is changed: the ag-orbital electron population grows,
and the electron density of the e~-level decreases. This picture agrees well with the
simple qualitative predictions of the crystal field theory.
The results of experimental electron density analysis can be summarized as
follows. In all minerals (from halite-type to silicates, spinels, garnets, etc.) atomic
charges are smaller than the formal atomic oxidation numbers. The difference is
greater when a real chemical bond is distinguished from an idealized ionic bond
2.1.7 Precision Electron Density Calculations 75

[100] [100]

- - -....-[110]
a
-b--"-- [110]

[100]

Fig. 6a-c. Experimental deformation electron


density for y-Ni 2 Si0 4 (a) (Marumo et al. 1974),
y-Co 2 Si0 4 (b), and y-Fe 2 Si0 4 (c) (Marumo et
al. 1977). The contour intervals are 0.2 eA -3 (a),
0.4 eA - 3 (b and c). Arrows show the threefold
axis in trigonal distorted oxygen octahedron.
The d.-orbitals (e~-levels) are perpendicular to
this axis, and their electron populations de-
creases in the series Ni-Co-Fe when the octa-
- - - t.... [110] hedron is compressed
c

picture. The chemical bond in Si-tetrahedrons of silicates have a predominant


covalent character with some n-components. In ring, chain, and framework
silicates, condensation of Si-tetrahedra results in the accumulation of elec-
tron density on the bridging oxygen atoms. Multicenter partially covalent
cation-anion chemical bonds are in corundum-type minerals (space group R3c).
An additional metal-metal bond can be formed in the basal plane (karelianite
V203) or perpendicular to this plane (hematite a-Fe 20 3) if the cation is at 3d-
element. The chemical bond nature in garnet oxygen polyhedrons is remarkable
for central atoms. Nontransition elements form ionic bonds with environment,
and transition and rare-earth elements form multicenter covalent bonds having
(J- and n-components.

The characteristics of fine crystal structure of minerals (real composition,


impurity distribution) can be studied by the X-ray diffraction method. They are
included into the crystallographic model by placing different atoms in the same
structure positions or by removing the atoms from others. Such a complication
76 Chapter 2. Crystal Structures of Minerals

of the crystal model results in correlations between refined structural parameters.


In order to weaken the correlation influence, one may either put restrictions on
the bulk chemical composition, optimize the model by a step-by-step fixation of
the population parameters, or refine the population parameters only with low
angle reflections, which are not affected by uncertainties in the other structural
parameters. The obtained parameters are model-dependent due to the approx-
imate nature of the crystal model, and complication of the extinction effects
considerations, so additional criteria are needed. A criterion which is used to
estimate the reliability of results is a good agreement between the results of
chemical analyses and the chemical composition obtained from X-ray data
without restrictions on the chemical composition during the process of model
optimization. When the element composition of the minerals mark ably varies,
the chemical analysis has to be obtained from the sample used in the diffraction
experiment.

References

Gibbs (1982) Molecules as models for bonding in silicates. Am Mineral 67: 421-450
Kirfel A, Eichorn k (1990) Accurate structure analysis with synchrotron radiation. Acta Cryst
46: 271-284
Lewis J, Schwarzenbach D, Flack HD (1982) Electric field gradients and charge density in
corundum, IX-Al z0 3 • Acta Cryst A38: 733-739
Marumo F, Isobe M, Saito Y et al (1974) Electron-density distributions in crystals of Ni zSi0 4 •
Acta Cryst B30: 1904-1906
Marumo F, Isobe M, Akimoto S (1977) Electron-density distributions in crystals of y-Fe zSi0 4
and y-Co zSi0 4 . Acta Cryst B30: 713-716
Sasaki S, Takeuchi Y, Fujino K, Akimoto S (1982) Electron-density distribution of three
orthopyroxenes, Mg zSi z0 6 , Co zSi z0 6 and Fe zSi z0 6 . Z Krist 158: 279-297
Spackman MA, Hill RJ, Gibbs GV (1987) Exploration of structure and bonding in stishovite
with Fourier and pseudoatom refinement methods using single-crystal and powder X-ray
diffraction data. Phys Chern Mineral 142: 139-150
Tsirelson VG, Ozerov RP (1994) Electron density and bonding in crystals. Adam Hilger,
Bristol
Tsirelson VG, Evdokimova OA, Belokoneva EL, Urusov VS (1990) Electron density distribu-
tion and bonding in silicates. Phys Chern Mineral 17: 275-292

2.1.8 High-Temperature and High-Pressure Crystal Chemistry

S.K. FILATOV and R.M. HAZEN

High-temperature, high-pressure, and compositional crystal chemistry are crys-


tallographic subdisciplines devoted to studying the nature and causes of crystal
structure variations with changes in temperature (T), pressure (P), or composi-
tion (X). These studies, collectively called "comparative crystal chemistry", have
been conducted by materials scientists for more than 65 years, since the pioneer-
2.1.8 High-Temperature and High-Pressure Crystal Chemistry 77

ing high-temperature X-ray diffraction studies of Wyckoff (1925), although the


past two decades have seen especially intensive research.
This review considers structural deformations which are common to all
crystals with changes in T, P, or X. We consider both continuous changes and
discontinuities (phase transitions) in lattice parameters and atomic coordinates.
Experimental studies of comparative crystal chemistry focus both on lattice
deformation, the external manifestation of changes, and at the atomic level
through studies of atomic coordinates, coordination polyhedral distortions, and
changes in interatomic distances and angles.
Powder and single-crystal X-ray diffraction are widely used to characterize
structural deformations and phase transitions of crystalline materials. Several
authors have summarized work at high T or P (Clark 1966; Buerger 1971;
Krishnan et al. 1979; Hazen and Finger 1982, 1985; Urusov and Pushcharovsky
1984; Taylor 1984-1988; Filatov 1990).

High-Temperature Crystal Chemistry

Three simple rules describe the basic features of thermal deformation of crystals.
Rule 1. The effect of temperature is scalar; therefore, thermal deformation of
a crystal is controlled by a structure and is described by a second rank tensor.
Rule 2. Anharmonicity of thermal oscillations of atoms results in an increase
in interatomic distances, i.e., in thermal expansion. More anharmonic oscilla-
tions with higher amplitudes, i.e., weaker bonds, result in more pronounced
thermal expansion.
Rule 3. An increase in crystal temperature is equivalent to an increase in the
intensity of thermal motion (vibrations, rotations, and jumps) of atoms and
molecules and gives rise to a chain of transformations whose tendency is toward
rising crystal symmetry: rising vibrational symmetry of atoms and mole-
cules -+ deformation-structural rearrangement in the direction of rising sym-
metry -+ transformation into more symmetrical high-temperature modification.
The following regularities were obtained from a generalization of experi-
mental data. They follow from the principles given above.

Deformations of Coordination Polyhedra. The mean linear coefficient of thermal


expansion, ci( ci = rty /3, where rty is the coefficient of volumetric expansion), of
a coordination polyhedron formed by essentially ionic bonds has simple re-
versible relations to Pauling bond strength (Fig. 7):

-
rt1000
1 ad
= ~ aT ~ 4.0(4)
(n) _
S2 ZcZa x 10
6 0 -
C,
1
(1)

where d is the cation-anion separation, n is the coordination number of the


cation, Zc and Za are cation and anion charges, respectively and S2 is the ionic
bond coefficient which is equal to ~ 0.5 for oxygen compounds. Thus, all types
of polyhedra with a definite cation and anion have the same parameter ci, which
78 Chapter 2. Crystal Structures of Minerals

t(S
x
points)

30
\\
(a)(S2ZxZa) = 4.0(4) X 10- 6 °e l
\~\ \
c:
0
II>
c: 2S a Oxides 8. Silicates
0 \ \
x Halides

~\1+
Co
)(
w 0 Carbides 8. Nitrides
0
E
4;-
20 \\ :
\ + Sulfides, Selenides, 8.
Tellurides

l!,,~
.J:.'
~ ~
- co
~ 0 IS
"0 -
'" X

'11
.J:.-
>- 0
0 0
Cl. 0
~
la- 10
0

'"
c:
.-I
+"
..," .......
c: ..... -....;
.......
.......
0
S
'"
:::;;
+ -- ---.:::~
--
0

0 f 1i
0 0.2 0.4 0.6 0.8 1.0 1.2
S2 ze Zo
n
Pauling Bond Strength

Fig. 7. The inverse relationship between polyhedral thermal expansion and Pauling bond
strength. (Hazen and Finger 1982)

can be used for predicting the effect of temperature on the polyhedron; fi. does
not depend on the structural bonds in the polyhedron. Furthermore, all coord-
ination polyhedra with similar values of Pauling bond strength have a similar
coefficient IX. For example, in octahedra with divalent Ni,Mg,Co,Fe,Cd,Mn,
Ca,Ba, and Sr fi. 1000 = 14 ± 1 x 1O- 6 °C- 1 .
In a structure with more than one type of cation polyhedron, the effect of
temperature is proportional to a polyhedral size. However, thermal expansion of
a crystal is usually not just the simple sum of changes in the cation-anion
separation; changes in angles between polyhedra must be also considered.

The Intensity of Thermal Expansion of Oxygen Compounds with Oxyanion


Radicals. Strengths of crystals are limited by their weakest bonds. In oxygen
compounds with triangular and tetrahedral radicals, which compose nearly all
the Earth's crust, the bonds out of radicals are the weakest. Therefore, such
compounds are characterized by an increase in the thermal expansion coeffic-
ient, ex, and by a decrease in melting point, T m' with decreasing value of
2.1.8 High-Temperature and High-Pressure Crystal Chemistry 79

a residual charge of one radical participating in bond formation in one space


dimension, Z. In general, this relationship is illustrated by Table 15; Z = Z/N,
where N is the number of space dimensions to form bonds out of polyions. For
example, N = 3 in silicates with isolated tetrahedra, 2 in chain and double-
chain silicates, 1 in layer silicates; in frame silicates Z and N toward 0 but
Z moves toward 1.

The Character of Thermal Deformations. This is controlled by crystal symmetry


and crystal structure; thus, deformations of isostructural compounds and mem-
bers of homological series are similar (for example, Fig. 8a-b, c-f). Cubic
crystals deform isotropically, while anisotropy of deformations of noncubic
crystals is determined by anisotropy of bond strength and by structural re-
arrangements, especially shearing of the structure.

Shears. The cell angles of triclinic and monoclinic crystals are not fixed by
symmetry; with changing temperature crystal structures are subjected to shears,
which are by nature highly anisotropic. Many oblique-angle crystals reveal
anisotropic thermal deformations, including a negative thermal expansion in
some directions.
Crystals with negative linear thermal expansion are uncommon in hexagonal,
trigonal, tetragonal, and orthorhombic materials, but in monoclinic and triclinic
crystals such an expansion is common. Note, however, that negative volumetric
thermal expansion is observed in few materials, except at temperatures near
absolute zero.
Oblique-angle crystals such as feldspars, amphiboles, many pyroxenes, and
micas, due to highly anisotropic deformations, are responsible for the decom-
pression of the rocks which they make up, as well as for increases in rock
permeability by fluids and in some cases for the accumulation of ores under
metamorphic conditions. Higher thermal deformation anisotropy, for instance,
the anisotropy of calcic pyroxenes (Fig. 8a-b) relative to alkaline (Fig. 8c-f) is
suggestive of a higher capacity for thermal decompression of the rocks.
Similarly, highly anisotropic thermal expansion of calcite is responsible for
the thermal decompression of marbles; the anisotropy of feldspars, specially of
high albite, for the decompression of rocks bearing them; and a sudden change
in volume for 0.6% during rx-fJ transformation in quartz at the temperature of
573°C is responsible in part for the decompression of granites, etc.

High-Pressure Crystal Chemistry

Three rules describe continuous structural variations with changing pressure for
most oxide and silicate minerals.
Rule 1. Changes in lattice parameters and atomic coordinates may be
characterized by second rank tensors. Lattice volume change with pressure is
constrained to be negative, and crystal axes almost always compress. (In
80 Chapter 2. Crystal Structures of Minerals

Table 15. Composition, structure, and thermal properties of oxygen compounds


Triangular radicals

Valency of the central atom Examples


of geometry
2+ 3+ 4+ 5+ 6+ 7+ of polyions
(borates) (carbo- (nitra-
nates) tes)

Tm
= :x _ _

~, !~ /
/
[T0 3 ]4- [T0 3 ]3- [T0 3 ]2- [T0 3 r [T0 3 ]0 [T0 3 ]+ Isolated
/
/ triangles
/ [T0 3 ]
/
/
/
[T02.5r- [TO Z •5 ]2- [T02.5r [T02.5]O [T02.5] + [TO Z •5 ]2+ Double
/
/ triangles
/ [T Z 0 5 ]
/
/
/
[T0 2 ]2- [T0 2 r [TOz]O [TO z ]+ [TO z ]2+ [TO z ]3+ Rings
/
/ [TO z ],
/ chains
I [TO z ]",
I
I
I
[ TO 1.75 ]1.5- [TO 1.75 ]0.5-
I [TO 1.75 ]0.5+ [TO 1.75 ]1.5+ [TO 1.75 ]Z.5+ [TO 1.75 ]3.5+ Double
I chains
/ [T4 0 7 ]",
/
/
/
[T01.5]O
I
/
/

[TOx]Z is the formula of the radical on a one-central-atom basis; X is the extent of polymerization of
triangles or tetrahedra through common tops (X = O:T): Z is the radical charge available for bonds
out of the polyion; =0>, -+ are the basic tendencies of variations in strength of the compounds; ex is the
coefficient of thermal expansion; Tm is the melting point.
2.1.8 High-Temperature and High-Pressure Crystal Chemistry 81

Tetrahedral radicals

Valency of the central atom Examples


of geometry
2+ 3+ 4+ 5+ 6+ 7+ of poly ions
(beryl- (bora- (silicates, (phos- (sulp- (perman-
lates) tes) germa- phates, hates, ganates,
nates) vana- chro- perchlo-
dates) mates) rates,
perio-
ates)

Isolated
tetrahedra
[T0 4 ]
/
/
/
[T0 3 . 5 ]0 Double
/ tetrahedra
/
/ [T 207]
/
/
[TO;]O Rings
/ [T0 2 ],
/
/ 'pyroxene'
I chains
I [T0 3 1o
I
I
[ TO 2.75 ]3.5- [TO 2.75 ]2.5- [TO 2.75 ]1.5- [TO 2.75 ]0.5-
I [TO 2.75 ]0.5+ [TO 2.75 ]1.5+ 'Amphibole'
I double
chains
/'
I [T 4 0 11 J"
/'

[TO u/' ] ° 'Mica' layers


/'
/' [T 2 0 5 ]oooo
/'
/'

[T0 2]O 'Silic'


/'
/' frames
/'
[T0 2 ]oooooo
a 00
a b IV
a~" CaMg
\ ______?poC
~ ..... ctJ.3
II,
~. \_200
400
~.--=-;;J!!!!! -~, c
600-fOOO ...---',
\ ,4
':i1
""-------~
\
\

c d e J+ f
Nare ex,,, CY-" NaAL J +
LiALH 0<./1
a
n
po
p:
'1:l
<t
...,
t'"
otJ3
~ ;;;" C -k / c -Y . I c ,=-== C n
...,
~
E
~
...,
rl
'...,2"
(1)
en
o....,
Fig. 8a-f. The figures for the coefficients of thermal expansion of the ac plane in diopside (a), hedenbergite (b), spodumene (c),
ureyite (d), acmite (e), and jadeite (f). (After Filatov 1990). Each radius-vector represents the value of the coefficient of thermal a:::
~.
expansion in a given direction. Shaded areas correspond to negative thermal expansion ...,
~
2.1.8 High-Temperature and High-Pressure Crystal Chemistry 83

LaNb0 4 one axis becomes slightly longer with pressure, but this expansion is
coupled to greater compression of a pseudo-orthogonal cell axis).
Rule 2. Cation-oxygen bond distances, d, show an average compressibility,
if (Fig. 9):
P- = - d1ap
ad ~ 0.044 (d
S2ZcZa
3) -6 bar -1 ,
X 10 (2)

where Sand Z are defined as above. Thus a given type of cation coordination
polyhedra (e.g., Mg0 6 ) displays similar compression in different structures.
Rule 3. Total volume compression of Il}ost rock-forming minerals is the sum
of polyhedral compressibilities, plus interpolyhedral compression related to
changes in cation-oxygen-cation angles. Polyhedral distortion (as characterized
by changes in oxygen-cation-oxygen angles) plays a minor role in mineral
compression below 100 kbar.
The driving force behind structural changes with pressure is volume reduc-
tion. Discontinuous structural changes involving increase in cation coordina-
tion are of special importance in understanding the Earth's deep interior.
Transformations, such as those of MgSi0 3 from pyroxene to garnet to perov-
skite, result in large density increases because of increased atom packing effici-
ency. Surprisingly, however, cation-oxygen distance generally increases across
such transitions (cation-anion distance increases with coordination number), so
high-pressure phases are often more compressible than their low-pressure poly-
morphs.

.......'... 2 o
III
.c
~ o
o
ICC

Fig. 9. The bulk modulus-vol-


ume relationship for polyhedra in
a variety of materials. (Hazen
and Finger 1979)
10 20 30 40 50 60 70
d3
84 Chapter 2. Crystal Structures of Minerals

Comparative Crystal Chemistry

The Analogy of the Variation of Volume with Temperature, Pressure, and Com-
position. Variation of coordination polyhedral volume can be predicted on the
basis of simple bond parameters: cation- anion separation (d), coordination
number of the cation (n), cation radius (r), cation and anion formal charges
(Zc and Za , respectively) and ionization (S2) :

IXv=-y
I (av)
aT (n)
~12.0 S2ZcZa xlO
-6 0
C
-1
, (3)

(4)

(5)

where X is the atomic fraction of the larger cation.


All crystalline substances can be represented in T - P- X space by surfaces of
a constant molar volume (isochoric surfaces) (see, for example, Fig. 10). One of
the consequences of a structural analogy between T, P, and X is the fact that
isochoric surfaces for many substances are at the same time isostructural
surfaces in T - P- X space.

800

MgO 0.2 0.4 0.6 0.8 FeO


Fe/( Mg + Fe)
Fig. 10. Isostructural surfaces for (Mg,Fe)O in T- P- X space. (Hazen and Finger 1982)
2.1.8 High-Temperature and High-Pressure Crystal Chemistry 85

Equations (3) and (4) lead to an important relationship of 0( to /3 for


a polyhedron (Hazen and Finger 1982):
0(//3 ~ 90 n/d 3 barre. (6)
Deformations caused by T or P or X may reveal similar features. In
compounds with essentially ionic bonds, an increase in temperature causes
structural deformation similar to the effect of a decrease in hydrostatic pressure
(inverse relationships), or to the effect of isomorphic substitutions of smaller
structural elements for larger ones. Substitutions in highly anisometric positions
or occurring together with other disturbances in the quasi-scalar effect can result
in moderation of such a similarity or even in its absence.
Figure 11 shows examples of the similarity of thermal (b) and pressure (c) as
well as thermal (d) and chemical (e) deformations of clinopyroxene structures
caused by the transformations of the most "soft" (large and irregular) M2
polyhedron (a).

d e

Fig. lla-e. Examples of the similarity of crystal deformations of various nature. (Filatov
1990). Compound~ with the clinopyroxene structure. a Coordination polyhedron M2 control-
ling the deformation of the crystal lattice. b Thermal-induced deformations of diopside.
c Pressure-induced deformations of diopside. d Thermal-induced deformations of vanadate
Li2 V206' e Chemical deformations of Li2 V206 (substitution Li-Na in M2). Arrows indicate
an increase in bond lengths, LlI, in M2 multiplied by 10 for thermal-induced and pressure-
induced deformations and by 5 for chemical deformations. Figures of deformations are given
for the crystal lattice
86 Chapter 2. Crystal Structures of Minerals

The inverse relationship between thermal-induced and pressure-induced


deformations occurs if the ratio n/d 3 is similar for all cation polyhedra in the
crystal structure (Hazen and Prewitt 1977a) or if one of the polyhedra in the
structure is more rigid (for example, in silicates) compared to other polyhedra
with similar alP. Many cation polyhedra in rock-forming minerals have similar
alP ratios; all Mg, Fe2+, AI, and Fe3+ octahedra have alP ~ 65 barre. There-
fore, many silicate minerals reveal inverse relationships, such as low albite
(Fig. 12b), diopside, and pyroxmangite. Rutile, on the other hand, being a com-
pound with essentially covalent bonds, demonstrates the absence of inverse
relationships (Fig. 12c).

The Relationship Between Thermal Expansion and Lithostatic Compression of the


Earth's Shells. During subsidence into the depths from the earth's surface, the
rocks, being subjected to the contrary influences of pressure and temperature,
expand and compress simultaneously. The relationship between these two
components is controlled by the equivalents alP for minerals and by temper-
ature and pressure gradients which are different at different depths.
The minerals of a single zone of the Earth have similar deformation equiva-
lents alP. The average value of the equivalent alP for atmospheric conditions
through a zone (Fig. 13), increases with increase in (1) depth of the zone, h,
(b) rock density, p, (3) p-wave velocities within the zone, Vp , and (4) the mean
symmetry of the minerals of the zone, (T. It is apparent that the higher coordina-
tion numbers (n) of the atoms in the deep-seated phases [or shorter bonds
(d) with constant n numbers] determine their greater values of density and
deformation equivalent alP [see Eq. (6)].
There is experimental and theoretical evidence that the deformation equiva-
lent alP increases with rise in pressure and increase in the depth of rocks in the
Earth. However, experimental data on the coefficients a and Punder conditions
of high temperature and pressure are insufficient at present to calculate the
values of alP for the minerals in abyssal zones.

IDEAL CASE

Structural Parameter

~creaSing
Pressure
a

Fig. 12a-c. To the problem of the "inverse" relationship between thermal- and pressure-
induced deformations of crystals. a Ideal case, after Hazen and Finger 1982. b Example of the
inverse relationship in low albite (Hazen and Prewitt 1977b). c Example of the absence of the
inverse relationship in rutile, after Hazen and Finger 1981
2.1.8 High-Temperature and High-Pressure Crystal Chemistry 87

LOW ALBITE NaAISi 30 S

~ c
111.0 ::J
Ul
Ql
Cl (')
"C ~
Ql
116.6 III
Qi ::J

-,
u co
CD
Ul
c 116.2
::J 0

7.18

7.16

7.14

b 7.12

\jlYo

D
1.02

/lncreaSing Temperature
1.01

t
Y...
cIa ,/
0.646
./"
0.647
Vo 1.00

/ . ~ 'ncreasing Pressure
, Cryogenic "'-
0.99
Temperatures '~

l(
0.98
c

Fig. 12. Contd.


88 Chapter 2. Crystal Structures of Minerals

50 '0
;(3 ~
:l
'>0 c:
c:0
Q)

0 0 a
40 .t:: 0 U
0 .!!!
~ "0

(J
.....
(/)
30
«;
.0

I~ 20

10
4km 40km 2900 km

10 100 1000 h. km

60 b c d

/(
40

20

2 3 4 6 8 10 12 1 12 24 36 48

p. g/cm 3 (f

Fig. 13a-d. Pressure equivalent of thermal deformation (rxIP), averaged on a mineral basis for
different zones of the Earth, as a function of depth to zone h (a), mean rock density under
atmospheric conditions p (b), P-wave velocities Vp within zone (c) and mean symmetry of
minerals in zone (J (d). (Filatov 1988). AI granite-metamorphic layer of continental crust;
A2 granulite-basite layer of crust; B upper mantle; C I transitional zone of mantle; C2 mantle at
depths of 600-800 km

In the deep-seated zones such a comparison can be made with the use of
approximate values a./p. With depth, expansion is slowed down as a conse-
quence of a decrease in the temperature gradient with depth, the proportion of
compression increases.
Beneath the stable continental regions, the top of the upper mantle separates
a near-surface shell of intense cooling of the planet (we may assume that the
mean value of the temperature gradient with depth is approximately 20 DCjkm
near the surface and 10 DC/km at a depth of 40 km) from the quite thermostable
upper mantle, in which the temperature gradient may be taken as constant and
equal to 1.3 DC/km. The pressure gradient is also constant to a first approxima-
tion and may be estimated at 1/3 kbar/km. As a result, during the subsidence of
rocks in the upper mantle, a temperature increase of 1 DC appears to correspond

You might also like