Anda di halaman 1dari 17

Journal of Non-Crystalline Solids 73 (1985) 1 17

North-Holland, Amsterdam

Section L Glass structure

S P E C T R O S C O P Y S I M U L A T I O N AND SCA'ITERING, AND T H E


M E D I U M RANGE O R D E R P R O B L E M IN GLASS
C.A. A N G E L L
Department of Chemistry, Purdue University, West Lafayette, Indiana 47907, USA

Consideration is given to phenomena which we believe to be controlled by the fluctuations in


the poorly understood intermediate range order in liquids and glasses. These include structural
relaxation, crystal nucleation and incipient liquid-liquid phase separation. Techniques which show
promise for investigating the intermediate range order are considered, and predictable or conceivable developments in these techniques which may, by the year 2004 (N.J. Kreidl's 100th), greatly
increase or even revolutionize our knowledge of the intermediate range order, are discussed. These
include difference spectroscopy, difference scattering, and computer simulation techniques. Finally, we consider possible developments in system preparation or system manipulation techniques
which may lead to new insights into relations between physical properties and intermediate range
order. An example of special interest is the preparation of noncommunicating microsystems of the
same size as the clusters which many believe are the building blocks of the glassy state.

1. Introduction

This manuscript is devoted to a consideration of some of the problems and


prospects concerning what might be called intermediate range order in glassy
solids. The microscopic distance range 8-50 A or - 3-20 atomic diameters is
the range within which are determined many important aspects of liquid and
glass phenomenology (e.g. relaxation of stress, nucleation of crystals, photochromism, incipient liquid-liquid phase separation, and possibly ultra-fast
ion conduction in glasses) and yet it is this same range which is beyond the
power of most structure-determining techniques to elucidate. We believe, with
others, that it is here that the primary challenge in glass structure understanding currently resides, and that the problems posed are challenging enough to
project, incompletely resolved, into the 21st century.
In this paper we will first give some lines to describing certain properties of
glass-forming liquids which seem to depend on the structural arrangements
well outside the first and second coordination shell of the most strongly
coordinating cations. These are considered in subsections 1 5 below. With the
importance of these regions (about which relatively little is known at the
present time) established, we will then (a) review some related phenomenology
in more detail for the case with which we are most familiar and then (b) discuss
the possible techniques which might be used in elucidating more clearly the
importance of medium range structure on material properties.
0022-3093/85/$03.30 Elsevier Science Publishers B.V.
(North-Holland Physics Publishing Division)

C.A. A ngell / Spectroscopy simulation and scattering

1.1. Structural relaxation


Structural relaxation is of course the fundamental phenomenon underlying
glass formation since it is the lengthening time scale for relaxation with
decreasing temperature which leads to the glass transition and the generation
of the "solid" state. The temperature dependence of the relaxation time near Tg
determines the ease of glass-working as well as the danger of devitrification by
growth of pre-existing nuclei. The temperature dependence of the average
relaxation time as well as the detailed relaxation function seem to be closely
connected with the nature of the intermediate range order. Thus, structural
relaxation is the first of our "glass science problems" identified as involving
the intermediate range order. We treat this matter in some detail below.

1.2. Crystal nucleation


The second phenomenon which is of vital importance and which involves
distance scales of the intermediate range is the nucleation phenomenon. We
believe that, except in extreme instability cases, nucleation in amorphous
phases requires fluctuations extending over distances somewhat greater than
those involved in structural relaxation, though, in most cases of practical
significance, it is still a phenomenon which occurs within subsystem sizes of
only tens of A in diameter. Techniques for probing the nature of structural
fluctuations involved in nucleation, and the relation between critical nucleus
structure and the structure of the thermodynamically stable phase into which it
grows, are practically nonexistent at this time, though the significance of the
question being addressed is obvious to all.

1.3. Incipient liquid-liquid phase separation


Another important characteristic of many glass-forming systems, which
involve distance scales of the intermediate class is encountered in demixing
systems. Two cases involving phase boundaries, or quasi phase-boundaries are
important. One is the phase boundary existing between microscopic liquid
droplets in pre-opalescent phase-separated glassy systems. This distance becomes macroscopic at the temperature and composition of critical demixing,
but in the majority of cases where a two phase structure has either developed
or only has a tendency to develop, the phase boundary or the preseparation
fluctuation dimension is of microscopic dimensions, and their structures are of
both academic and technological interest. The other, and the least understood
though not the most important technologically, is the mean dimension of
composition fluctuations in glasses which exhibit a tendency to phase separation, without actually achieving it. A current case in question is that of the
AgI-, CuI-, and LiI-based fast ion conductors. The best conductors seem
always to be those at the iodide-rich end of the glass-forming range where
there are indications that the iodide is segregated in some sort of ramified

C,A. Angell / Spectroscopy simulation and scattering

prephase separation clusters within which the properties approach those of the
pure iodide.

1.4. Density fluctuations


On a smaller distance scale, but one still classifiable as intermediate, are the
configurational density fluctuations which determine the configurational part
of the thermodynamic isothermal compressibility and which are a necessary
feature of all vitreous systems. The magnitude of these fluctuations is known if
the compressibility of the liquid at the temperature of vitrification, and the
glass compressibility, are both known. In all but exceptional cases they will be
of the order of a few angstroms. How closely this distance scale relates to that
involved in structural relaxation is not clear at this time. Free volume theories
for transport properties in which the expansion coefficient plays a vital role in
the temperature dependence of transport, would suggest that density and
entropy fluctuations are, in fact dominating the probability of molecular
rearrangements.

1.5. Clusters
Finally, on a larger scale than the latter, both in physical size and in
potential for scientific controversy, lies the cluster dimension of the paracrystalline models of glass which are currently experiencing a resurgent tide of
interest. Championed by Phillips [1] who musters a great deal of spectroscopic
and scattering information in search of support for cluster models for silica
glass ( - 30 ,~ crystobalite paracrystals with Si = 0 internal surfaces), models of
this general type are currently being supported and rejected by roughly equal
numbers of investigators.

2. Structural relaxation and the intermediate range order: "strength" and


"fragility" in viscous liquids
We choose the structural relaxation problem for detailed discussion of
intermediate range order effects because of its overriding importance in glass
formation phenomenology, and of our own preoccupation with this subject
area.
Fig. 1 shows the behavior of the structural relaxation time, as it is reflected
by the shear viscosity, for four groups of liquids, one silica-based with various
concentrations of modifying oxides, the second BeF2-based and modified by
extra fluorides, the third is zinc chloride-based and modified by chloride ion
additions, while the fourth is ZrF4-based and is a member of the heavy metal
fluoride glasses currently under development for fiber optics applications. To
permit more effective comparisons, the data in fig. 1 are plotted on a reduced
inverse temperature scale using the temperature at which the viscosity reaches

C.A. Angell / Spectroscopy simulation and scattering

I0

n
I

E
o

t~

Q.

40.2

0.6

1.0

T/Tcj

._.

1.4

I.~3

2.2

"~

a.
t-

o B

~
0
o

0
o

Tg/T
Fig. 1.Scaled viscosity data for glass-forming liquids showing range of behavior from "strong",
characteristic of open tetrahedral network liquids, to "fragile" typical of ionic and molecular
liquids. Except for the fluorides BeF 2 (strong) and ZrF 4 - B e F 2 based mixtures (fragile) a c o m m o n
point at the high T extreme is indicated. Note added in proof." New viscosity data by Pantano et al.
for ZBLA extending to 1011 P suggest Z B L A should superimpose on 21~iCl 3. KC1.

1013 P as a normalizing parameter [2,3]. The common point at 1 / T = 0 has


been discussed elsewhere [4] and is to be understood in terms of compensating
effects of bond strength on shear modulus and quasi-lattice vibration frequency
in the expression ~/r~ oo = Goo%ibFig. 1 shows, interestingly enough, that the two glass-forming systems
currently under most active development as fiber optics materials find themselves at opposite extremes of this behavior pattern [4]. The technological
consequence of this difference is that it is very much more difficult to work in

C.A. Angell / Spectroscopy simulation and scattering

a controlled fashion with the softened fluoride glass because of its much
greater temperature dependence of relaxation time in the viscosity range where
working is carried out. We believe this distinction is to be associated with the
differences in stability of intermediate range order between the two systems.
The origin of the almost Arrhenius variation of the relaxation time in liquid
SiO 2 (and its simple exponential character which is implied by the data on
liquid GeO 2 which almost superimposes with that of SiO 2 in this reduced
representation) is clearly in the extended three-dimensionally connected network structure. This structure, in which every silicon is connected to four
nearest neighbor silicons through bridging bonds is self-reinforcing: it offers
both resistance to the rupture of any single bond, and restrictions on the
number of new configurations generated by the rupture of a single bond when
it does occur. Such a structure is resistant to thermal degradation, and reflects
this in the very small change of heat capacity which accompanies passage
through the glass transformation region [5], i.e. even though the structure can
change on the time scale of a heat capacity measurement, the additional
contribution to heat absorption per degree of temperature rise is small. This
can be demonstrated simply and semi-quantitatively with rudimentary models
of glass excitations such as the " b o n d lattice model" [5] and more recent and
refined models of the same ilk due to Brawer [6].
That the foregoing features are to be associated with the mutually supporting three-dimensional network structure can be confirmed by comparing with
silica the properties of aluminum-substituted silica structures when the charge
shortfall due to aluminum is compensated by sodium ions on the one hand and
by calcium ions on the other. In the former case (liquid albite NaA1SiBOs) the
viscosity remains close to Arrhenius in character [7] and the change in heat
capacity on glass transition is small [8]. In the calcium compensated case
(anorthite CaA12Si2Os) on the other hand, the more competitive status of the
charge compensating calcium leads to severe network distortions, as if pressure
were applied [9], with the result that the viscosity temperature dependence
becomes more strongly non-Arrhenius and the heat capacity change at T~ is
distinctly increased. Characteristic of liquids in this intermediate region of the
total range of behavior seen in fig. 1, the viscosity of anorthite remains
non-Arrhenius all the way to the glass transition. In fact, it almost superimposes on the curve for zinc chloride [3,10]. Zinc chloride itself behaves rather like
the hydrogen bonded liquid n-propanol for which a wide (reduced) range of
data exist [4,11]. The high temperature data for propanol are used to guide the
extrapolation of the ZnCI z and anorthite curves to 1 / T = O.
A different and intermediate loss of self-reinforcing capability follows when
alkali oxide is added alone to the silica structure. This, as the common wisdom
has it, ruptures bridging bonds in proportion to the number of oxides added.
At the disilicate composition the conditions for two dimensional sheet-like
structure are achieved. Indeed X-ray studies by Imaoka et at. [12] show a
distorted planar structure is achieved. This loss of bonding constraints in one
of the three dimensions is accompanied by larger departures from Arrhenius

C.A. Angell / Spectroscopy simulation and scattering

behavior than seen before, though the liquid is still less than halfway between
the strong and fragile extremes, see fig. 1.
If two-dimensional ordering is related to relaxational character it might be
expected that As2S 3 [13] would exhibit similar characteristics to those of
N a 2 0 . 2 S i O 2. The data for this case [14] near Tg are included in fig. 1 to
support this notion. The trend seems to be continued when additional alkali
oxide is added to reach the metasilicate composition according to limited data
from Endell and Hellbrugge [14]. At this composition, one-dimensional chain
structures should be the first order description of the intermediate range order.
Some M.D. simulations of this structure are discussed in ref. [9].
There are, unfortunately, no viscosity data on orthosilicate type liquids,
although it is known that glasses can be formed by the quenching of a mixture
of calcium and magnesium orthosilicates [5].
Some correlations with structure may be made. Although the first coordination shell remains fixed for Si-O, coordination no. = 4, the second nearest
neighbor distribution g(Si-Si) obtained from MD calculations [16-19] shows
strong variations, see fig. 2. The changes show up not in the Si-Si distance nor
in the peak width at half height, but in the coordination number and in the
second nearest Si neighbor position. Such a structural characterization is,
however, crude and inadequate to give insight into either the origin of
non-Arrhenius behavior itself or of the changes in the parameter fl in the
detailed relaxation function O(t)= e ~tt/~l~) which seem to associated with it
[20-221.
Zinc chloride, illustrated in fig. 1, is a weak network liquid at the outset.
The size of the zinc ion relative to its chloride ligands is less favorable to four
coordination than in the case of silica, and the result is a network in which the
CI-Zn-C1 bond angles are rather acute, and the packing density is high
[22,23]. This liquid has characteristics intermediate between the strong and
fragile classes [4]. However, the effects of the (less demanding) medium range
order imposed by the chloride bridging are still present as can be seen when
the data for ZnC12 are compared (see fig. 1) with those of a zinc chloride-based
system in which the bridging bonds have been broken by chloride ion additions in the form of pyridinium chloride Py+C1 [10].
In both the latter cases the restrictions on the first coordination number
have remained in force, but relief of order constraints at the intermediate range
has been achieved. If we relax the constraints further by removing restrictions
on the first coordination shell or at least making interchanges between first and
second shells simpler, a further increase in "fragility" occurs. These conditions
can be seen in the behavior of the molten heavy metal fluorides where the
coordination number of the most strongly coordinating species Zr 4+, is less
well specified. Current studies [24,25] indicate that zirconium ions exist in a
mixture of 7 and 8 coordination states even in the unmodified liquid fluoride,
and that this indefiniteness and variability in the medium range order is
enhanced by the addition of modifying fluorides.
Extending our considerations to non-ionic glass-forming liquids, we might

C.A. Angell / Spectroscopy simulation and scattering


q.o00

3.000

~-" 2.000-

1.000-

.000
000

SILICON

2.000
q.o00
6.000
R [ I , J ) (ANGSTROMS)
ABOUTSILICON
[STANDARD
NORMRLIZATION)

8.000

20.00 -

,5.oo-

IA

/
/ //

ev- i0.00-

5.00 -

.00

000

/.///

~
I

/ /

2.000

4.000

6.000

8.000

R(I,J) (ANGSTROMS)
SILICON

ABOUT SILICON

Fig. 2. Si-Si pair distribution functions for SiO 2 and alkali silicate c o m p o s i t i o n s s h o w i n g the
c h a n g e in S i - S i c o o r d i n a t i o n numbers. T h e decrease implies a relaxation of the intermediate range
order which is associated with the increasing fragility of the viscous liquid.

note that alcohols tend to occupy the center of the fig. 1 pattern [4], while
molecular liquids fall at the fragile extreme. Again this is consistent with the
restrictions on intermediate range order imposed by hydrogen bonding in the
alcohol cases and the lack of anything but repulsion potential determined
packing restrictions in the case of the van der Waals liquids.
3. Techniques for investigating intermediate range order
In this section we first discuss some ways of exploiting and extending
known techniques which have the power to give structural information which is

C.A. Angell / Spectroscopy simulation and scattering

relevant to the sort of problem outlined in the previous section. We then go to


speculate about other possible avenues for exploring the physical importance
and phenomenology of intermediate range order-determined properties, and to
consider possible direct and indirect means of observing this order.

3.1. Spectroscopic techniques


In general it has been concluded [26] that spectroscopic techniques are
primarily useful for elucidating the first nearest neighbor arrangements, but
that with adequate care information on the nature of second nearest neighbor
arrangements can be obtained. An example of the latter is the molten salt
study of Smith et al. [27] in which systematic studies of this composition
dependence of ligand field spectra (in the visible range) studies were used to
establish that Ni 2+ in liquid KCI-LiC1 solutions existed in two coordination
states and that in the octahedrally coordinated configuration it was Li cations
which occupied the second nearest neighbor shell while in the tetrahedrally
coordinated state it was potassium ions which occupied the second nearest
neighbor position. To the best of this author's knowledge such systematic
spectroscopic studies have not yet been performed in oxide glass systems, and
the field could profit from the development of relevant studies of this type.
Trivalent titanium, for instance, could conceivably be used as a probe for the
behavior of Ga 3+, or even perhaps A13+ in glasses where the composition has
been chosen to produce both tetrahedral and octahedral coordination states.
Alternatively Ni 2+, which has usefully distinct spectroscopic signatures in
octahedral and tetrahedral sites, could be used as a probe for Mg z+ coordination in glasses where, particularly at high temperatures, tetrahedral as well as
octahedral sites are anticipated. Nickel spectra obtained in high temperature
melts have indicated [28] that except in the most acid conditions, tetrahedral
coordination is the common state in the low viscosity high temperature liquid.
It is of interest to determine whether, in such systems containing more than
one type of alkali, there is a pairing of alkali type with coordination number
type for more highly charged cations. Such studies could shed new light on the
mixed alkali effect on which little specifically structural information is available.
This problem in intermediate range (at least second nearest neighbor)
ordering would require the use of high temperature visible spectrophotometers
(of which the CARY-17DHC is the outstanding commercial example) and
computer-assisted data analysis to properly execute the spectroscopic subtractions which are needed to clarify the second nearest neighbor arrangements.
Useful, and currently unavailable, information on mutual cation ordering in
glass forming liquids could be obtained in this manner. The importance of
such information can be emphasized by referring to the striking differences in
structural relaxation characteristics of NaA1Si308 (albite) and CaA1Si2010
(anorthite) illustrated in fig. 1, the difference clearly being associated with the
difference in the aluminum second nearest neighbor relationships.

C.A. Angell / Spectroscopy simulation and scattering

A powerful technique for exploring directly the latter second nearest


neighbor problem, one which is just beginning to be exploited and will surely
contribute a great deal of structural understanding in the next two decades, is
solid state NMR. Not only does this technique distinguish with great clarity
between octahedrally and tetrahedrally coordinated AI (Muller et al. [29],
Ohtani et al. [29] but tetrahedral sites with different second nearest neighbors
may be revealed in the right conditions. Furthermore, the presence of a
microscopic crystalline fraction can be detected in partly devitirfied samples by
the line shape due to the narrow band characteristic of the ordered phase.
Finally, the fact that the spin lattice relaxation time is longer for ordered than
disordered regions means that manipulation of the "dead" time experimental
variable allows the investigator to bring ordered regions "into focus" in a
sense.

In recent studies using the Si nucleus, Gerstein (Gerstein and Nicol [29]) has
been able to detect 4 distinct Si resonances in an aluminosilicate glass, each
being attributed to a particular combination (total 4) of AI + Si in the second
neighbor shell. Si with 4 A1 in the second shell are, of course, a weak minority
occurrence but nevertheless exist.
Since these network sites are well defined it must be expected that there will
be difference in the potential energies of the alkali cations which charge
compensate the AI 3 + cations in the network. Such differences may well be the
origin of the different energy sites of alkali cations which current weak
electrolyte theories of glass conductivity are obliged to postulate. By analyzing
the thermal history dependence, and composition dependence of the Si A1
neighbor relationships and comparing with electrical conductance responses to
the same two variables and with the 23Na+ solid state N M R spectrum itself, it
might be expected that much progress in understanding the fast ion conductor
problem will be made in the near future and the problem may well be solved
by 2O04.
While spectroscopic techniques, such as the d d transition and solid state
N MR spectroscopies discussed above, may elucidate important second nearest
neighbor structure problems, they appear to be inadequate to research the
nature of third and fourth nearest neighbors. Since it is in these larger distance
ranges that much important physics, including the nucleation phenomenon, is
determined, we must inquire as to suitable methods for obtaining information
in this range.

3.2. Pair distribution functions from simulation and difference scattering techniques
3.2.1. Computer simulation techniques
The application of fast computers, using sophisticated multiparticle system
simulation programs to obtain both structural and dynamic information on
ionic liquid structures, has been exploited by several workers in recent years
[16-19,29-32], and their power is generally appreciated. Although there are

C.A. Angell / Spectroscopy simulation and scattering

10

FN.~leOe 6OOOK.z.99o

1.500-

1.000-

~o

.500

c~

$
.DO0

.000

4.000

2.000

8,000

8.000

R(I,J) (ANGSTROMS)
ABOUTBLUMINUM

SODIUM
ZB

21:12

- 21ZRF4-I2BAF2

[STRNDQRD NQRMALIZQTI~N)
- ~00K.2.990.

8.000
c~

6.000

4
8
o

2.000
o

~z
cic

.000
.000
ZiRCON]UM

2.000
R(I.J)
ABOUT ZIRC@'!UM

q'.ooo
(ANGSTROMS)

6.000

8.000

(STA',;D~RDNORHALIZATIO~,J)

Fig. 3. (a) Correlation of Na + and AI 3+ position in high temperature NaAISi206 melts at low
presure. (b) Z r - Z r pair distribution function in quenched BaF2-ZrF4 melts showing presence of
two distinct Z r - Z r distances (associated with single and double fluoride bridges).

severe limitations on the information on experimental glassy states [16,33],


these techniques are rather well suited for the study of hot fluid states where
experimentation is particularly difficult. Examples of their output have already
been given in fig. 2, and fig. 3 shows relationships between aluminum and
alkali cations in some sodium aluminosilicate melts which have recently been
studied as a function of increasing pressure [9]. Dropping the temperature
quenches in the high temperature structure but reduces vibrational smearing so
that such interesting features as the existence of two distinct Z r - Z r distances in

C.A. Angell / Spectroscopy simulation and scattering

11

liquid (hence presumably also glassy) fluorozirconates can be revealed [25] (se
fig. 3b). Similar features have been seen in preliminary simulation studies of
silicate liquids in which some "oxide" has been replaced with "sulfide" (i.e. by
doubly charged anions with an appropriately increased repulsive potential
parameter) and, more prominently, in an all-sulfide melt of stoichiometry
Li6SizS 7. Again, the new Si-Si distance is a short one, 3.5 A, also due to
double sulfide bridges such as occur in SiS2 itself.
The system represented in fig. 3 is relatively small (190 atoms) but with
growing computer power and improved algorithms, the investigation of much
larger systems is becoming feasible, Such simulations will give full detail on the
intermediate range order out to distances of the order of 15 A, which should be
sufficient to contain most of the important physics of microhomogeneous melt
behavior. Algorithms for obtaining viscosities of simple liquids with good
accuracy have been developed [34], and it should not be long before the
viscosities of complex ionic liquids of different fragilities (see fig. 1) can be
determined in the same way. Given the rate of development in power, and
decrease in price, of computing hardware, it is reasonable to suppose that
direct computer-simulation studies will be in a position to provide a great deal
of insight into the relation between intermediate range order and physical
properties by the year 2000. It, of course, must be demonstrated that the pair
potentials used in these simulations are adequate for the purpose. In this
respect it is necessary and important that experimental techniques be devised
for obtaining equivalent information for at least a few important test cases. W e
consider how this will probably be done in the following paragraphs.

3.2.2. Anomalous X-ray scattering studies using small laboratory X-ray laser
sources
The major problems encountered in attempts to unscramble total distribution functions obtained from simple X-ray scattering studies are well recognized. Major strides towards the refinement of our knowledge of pair
functions has, however, been made with the development of differencing
techniques. In particular the isotope replacement technique in neutron scattering for ionic and metallic liquid structure studies has been skillfully exploited
by Enderby and colleagues [35], and Dupuy and co-workers [36] and a start
has been made on its application to ionic glasses by Wright and co-workers
[37]. The number of pair functions contributing to the final pattern is greatly
reduced in this manner, and interpretation of structural features in multicomponent systems becomes possible. A limit is imposed by the statistical noise in
the initial patterns. With one or two orders of magnitude increase in counting
statistics 3rd and even 4th order differences could be taken and enormous
structural detail revealed. We consider how such ameliorations of the basic
idea involved in difference techniques may come about in the following
paragraphs.
A more exciting and potentially "personalizable" technique for obtaining
difference patterns involves the use of the anomalous X-ray scattering phenom-

12

C.A. Angell / Spectroscopy simulation and scattering

enon, in which the differencing may be done using only a single sample, which
is investigated by X-rays of different wavelength. If one of the wavelengths is
chosen to fall on an absorption line for one of the nuclear species of interest,
the resultant patterns can be subtracted to yield information on the pair
functions involving only this species. While this technique is currently in its
infancy [37,38], and suffers, at the moment, from the need to have access to a
synchrotron, this position is possibly about to undergo a drastic change. Work
by Rhodes and co-workers at the University of Illinois [39] has demonstrated
the feasibility of generating extreme UV or soft X-ray quanta at intensities far
in excess of those available from synchrotrons, and tunable over wide ranges.
Plausible arguments have been given by Rhodes [40] to the effect that by the
extension of current developments even hard X-ray laser sources with intensities far in excess of those available in synchrotron sources could be developed
in the near future. Evidently [40] the hardware involved in the multupling
(septupling is needed ultimately) of the initial laser frequencies to produce
intense X-ray beams could be available to the individual investigator at costs
of the order of $100000, in the not far distant future. Such instrumentation,
clearly, would revolutionize the experimental investigation of intermediate
range order in liquids of all types. Such "surrogate synchrotron" systems could
conceivably be available to the individual investigator within the decade. The
corresponding increase in information on intermediate range order could
become a surfeit by the year 2004.

3.3. X-ray lasers and the relaxation of intermediate range structure


The advent of the ultra high intensity X-ray sources discussed in the
previous section would also permit the investigation of important questions
concerning the dynamics of intermediate range order. Experiments equivalent
to the probe ion relaxation experiments (short range order relaxation) performed in this laboratory [41-43] will become feasible for following relaxation
of various features of the structure factor. For instance, it will be possible to
study the time development, following a temperature jump, of the small q
feature of S(q) in As2S 3 which has been studied recently by Busse [44] as a
function of temperature. This peak is related to the separation of raft-like
features of the microstructure. Combination of relaxation studies of short and
intermediate range order, with thermodynamic studies (e.g. volume relaxation)
would do much to unravel such currently vexing problems as the origin of the
e -[~/~)BI relaxation function characteristic of viscous liquids and glasses.
Time-dependent differencing techniques may also become a possibility with
the high intensity sources (projected to be 1 0 6 m o r e intense than current
synchrotron sources, and fully tunable). By alternating short exposures, the
time development of different pair functions following a perturbation should
become possible. Limitations are set by the tendency of incident energies of
these magnitudes to vaporize the sample.
Time-dependent small angle X-ray studies, which in principle could show

C.A. Angell / Spectroscopy simulation and scattering

13

how the larger scale (20-200 A) structure develops during crystal nucleation or
liquid-liquid phase separation, will also become feasible.

3.4. Other techniques


There are other techniques, such as transmission electron microscopy and
lattice imaging from quasi-crystalline or highly correlated regions of the
amorphous structure, which will be of the greatest importance in enhancing
our understanding of intermediate range order. With resolution of the order of
3 A now becoming possible [45,46] it seems clear that in the next two decades
studies in which the chemical components are chosen to maximize electron
density differences between dense-packed and loose packed regions of the
amorphous structure will contribute greatly to our conception of the essential
features of amorphous packing, and in particular the relation of intermediate
range order to paracrystallinity [1]. This author is not competent to comment
on the most probable developments in this area but their potential impact on
our thinking, given their "real space" information aspects, will be enormous.
Finally we should note the possibility of gaining fundamental insights from
the study of microscopic model systems such as the latex or lucite microsphere
(0,3 ~m diam.) suspensions of Ottewili and Pusey and possibly the new 5 nm
diam. droplet microemulsions with vitrified droplets. These systems which can
be obtained transparent, and interrogated by visible or UV light, may behave
like hard, soft or attractive spheres (see sect. 4.2 and ref. [55]).

4. Short range order and physical properties


In this section we return to our starting point and speculate briefly on
possible developments in techniques for probing the role played by intermediate range order in determining physical properties.

4.1. Dependence of "'fragili(v" on intermediate range order, bv time scale mixing


4.1.1. Fully amorphous phase studies
If glasses that have two structural order parameters with very different
relaxation times can be obtained, and if the slow order parameter involves the
intermediate range order via, for instance, a ring-chain conversion then it may
be possible to monitor the relation between liquid fragility and structure rather
directly. A possible case would be that of NaPO3 or its relatives in which small
anion rings and long chains are both possible and interconversion from one
structure to another can occur as a function of time. If this equilibration could
be slowed down relative to local structure relaxation times, by, for instance,
suitable choice of temperature, then their isothermal viscosity vs. time behavior
referred to periodic Tg measurements could reveal the relationship between
extended structure and the fragility.

14

C.A. Angell / Spectroscopy simulation and scattering

Another possibility is to combine the observations of de Neufville and


Rockstad [47] who correlated Tg and band-gap measurements with connectedness in related chalcogenide glasses, with those of Calemzcuk [48] who showed
that irradiation of Se below Tg could greatly enhance the rate of enthalpy
relaxation below Tg, to investigate in a dynamic mode the relation between
bonding dimensionality and relaxation kinetics. For instance, a photon correlation determination of the relaxation function of a viscous liquid of a given
connectedness could be determined in the presence and absence of an irradiating beam at the band gap frequency to monitor simultaneously changes in the
average relaxation time and associated changes in the /3 parameter of the
relaxation function e I~'/~)~1. By variable choice of constituents, bands associated with in-plane order and between-plane order (layer crosslinks) could be
selectively irradiated to establish the connection between structure relaxation
time and nonexponentially. Some preliminary attempts along these lines using
dielectric relaxation to detect changes have been made by Calemczuk [48], and
they deserve to be extended in the next decade.

4.1.2. Fractionally crystallized glass studies


By correct choice of composition it is possible to cause glasses to nucleate
and crystallize very slowly, with ultimate "crystallite" size far below the visible
range. In fact, both our group [49] and Wright and colleagues at the ILL,
Grenoble [50], have caused complete crystallization to occur under conditions
well below Tg in which mean diffusion paths are less than 20 A. According to
small angle neutron scattering studies, no structural inhomogeneities in excess
of 10 A in diameter can be observed in these products [50] despite the fact that
the glass ~ crystal phase change in essentially complete. These irreversible
changes could be viewed as controllable time-dependent changes in the intermediate range order, and their effect on viscosity during the period of
ultra-microcrystallite formation could be monitored using time scales adequate
to ensure local liquid-like structural equilibrium.
4.2. Study of systems which are themselves microscopic
Another approach to the characterization of processes dominated by intermediate range order characteristics may be via the preparation of systems
which themselves have the dimensions of the range of interest. This is now a
possibility with the recognition and study of thermodynamically stable systems
in which there is microscopic phase separation yielding more or less monodisperse microphases of diameter 15-100 A. For instance, water can be obtained
in microdroplets of this size range dispersed in decane using certain molecular
or ionic surfactants [51]. Study of its spectroscopic characteristics have shown
it to be in an essentially identical state to ordinary bulk water [52]. Microemulsions of molecular organic liquids have likewise been obtained dispersed in an
aqueous medium [53] and in at least one case the development of glass-forming
microemulsion systems has been reported [54,55], see fig. 4.

CA. Angell / Spectroscopy simulation and scattering

15

S (surfactant)

/~ween

//

(PG'3H20)

80

o o.

oo,

Vol % Aqueous Phase

(OiE,O-Xylene)

Fig. 4. Composition regions in the pseudo-ternary system (propylene glycol trihydrate)+(Tween 80


surfactant)+(o-ylene) in wl~h 20-50 ,~, nficrodroplets of o-xylene can be studied in the viscous
liquid and glassy states in a matrix of glass PG. 3H20.

The glass transition temperature in the latter was found to be almost


unchanged from that of the bulk material, and to be continuous between
normal and microemulsion phases. In the finer emulsion and the microemulsion states, crystallization after glass transition does not occur [55]. Variations
in the "oil" phase volume fraction between 0 and 50% are possible, see fig. 4,
dotted line.
The possibility of using such systems for study of system size-dependent
(indeed, wave vector-dependent) properties has just been recognized, and its
exploitation in the next decade may be anticipated. We must emphasize that in
such systems we have individual entire systems which are of the dimensions of
a single cluster of clustermodeler's glasses and not much larger than the
"amorphons" described by Hoare for packing of spherical particles [56]. The
fact that Tg remains essentially unchanged under these circumstances [55]
would argue that at least the transport properties of viscous liquids are not
determined by the presence of clusters if they exist. It will likewise be of
interest to pursue the existence and properties of secondary relaxations in such
microsystems because of the tendency to think that they originate in the
"tissue" material of microscopically inhomogeneous glasses.
A number of problems in interpretation of the behavior of such truly
microscopic (better, nanoscopic) systems will have to be solved, and probably
it will the features they show in c o m m o n with bulk system rather than the
converse which will be the most informative. A twenty year time scale is
probably appropriate in this instance also.

16

C.A. A ngell / Spectroscopy simulation and scattering

This work has benefited from the support of the NSF Solid State Chemistry
Grant No. DMR 8007053.

References
[1] J.C. Phillips, J. Non-Crystalline Solids 34 (1979) 153; Phys. Stat. Sol. (b)101 (1980) 473; Sol.
State Phys., in press.
[2] W.T. Laughlin and D.R. Uhlmann, J. Phys. Chem. 76 (1972) 2317.
[3] (a) C.A. Angell and J.C. Tucker, in: Physical Chemistry of Process Metallurgy: the 1973
Richardson Conference, eds., Jeffes and Tait (Inst. Min. Met. Publ., 1974) p. 207.
(b) C.A. Angell and W. Sichina, Ann. N.Y. Acad. Sci. (Proc. Workshop on Glass Transition
and Nature of the Glassy State) 279 (1976) 53.
[4] C.A. Angell, Strong and Fragile Liquids, in: Proc. Workshop on Relaxation Processes,
Blacksburg, Va, ed., K. Ngai (July 1983).
[5] C.A. Angell and K.J. Rao, J. Chem. Phys, 57 (1972) 470.
[61 S.A. Brawer, J. Chem. Phys. 81 (1984) 954.
[7] D. Cranmer and D.R. Uhlmann, J. Geophys. Res. 86 (1981) 7951.
[8] P. Richer and J. Bottinga, Geochim. Cosmochim. Acta 48 (1984) 453.
[9] C.A. Angell, P.A. Cheeseman and S. Tamaddon, Bull. Miner. 1 / 2 (1983) 87; Science 218
(1982) 885.
[10] A.J. Easteal and C.A. Angell, J. Chem. Phys. 56 (1972) 4231.
[11] A.C. Ling and J.E. Willard, J. Phys. Chem. 72 (1968) 1918.
112] M. Imaoka, H. Hasegawa and I. Yasui, Phys. Chem. Glasses 24 (1983) 72.
[13] S.V. Nemilov, Sov. Phys. Sol. State (1964) 1075.
[14] K. Endell and H. Hellbrugge, Angew. Chem. 53 (1940) 271.
[15] (a) P.R. McMillan, J.-P. Coutures and B. Pirion, C.R. Acad. Sci. Paris, Ser. II 292 (1981) 195.
(b) P.F. McMillan and B. Piriou, Bull. Miner. 106 (1983) 57.
[16] L.V. Woodcock, C.A. Angell and P.A. Cheeseman, J. Chem. Phys. 65 (1976) 1565.
[17] T.F. Soules, J. Chem. Phys. 71 (1979) 4570; J. Non-Crystalline Solids 49 (1982) 29.
[18] S.K. Mitra & R.W. Hockney, Phil. Mag. B48 (1983) 151.
[19] S.H. Garofilini, J. Chem. Phys. 76 (1982) 3189; J. Non-Crystalline Solids 63 (1984) 337.
[20] H. Tweer, J.H. Simmons and P.B. Macedo, J. Chem. Phys. 54 (1971) 1952.
[21] K. Ngai, Solid State ionics 5 (1981) 27.
[22] L.M. Torell and C.A. Angell, J. Chem. Phys. 78 (1983) 937.
[23] J.A.E. Desa, A.C. Wright, J. Wong and R.N. Sinclair, J. Non-Crystalline Solids 51 (1982) 57.
[24] R. Coup& D. Louer, J. Lucas and A.J. Leonard, J. Am. Ceram. Soc. 86 (1983) 523.
[25] J. Lucas, C.A. Angell and S. Tamaddon, Mat. Res. Bull. 19 (1984) 945.
[26] J. Wong and C.A. Angell, Glass: Structure by Spectroscopy (Dekker, New York, 1976) Ch.
12.
[27] G.P. Smith, J. Brynestad, C.R. Boston and W.E. Smith, in: Molten Salts, ed., G. Mamantov
(Dekker, New York, 1969).
[281 T.C. Lin and C.A. Angell, J. Am. Geram. Soc. 67 (1984) C-33.
[29] D. Muller, G. Berger, I. Grunze, G. Ludwig, E. Itallus, and U. Haubenreisser, Phys. Chem.
Glasses 24 (1983) 37; E. Ohtani, F. Taulelle and C.A. Angell, Nature (1985) in print; B.
Gerstein and A. Nicol, J. Am. Chem. Soc., to be published.
[30] Y. Matsui and K. Kawamura, Nature (Lond.) 285 (1980) 648; Y. Matsui, K. Kawamura and
Y. Syono, Adv. Earth Planet. Sci. 12 (1982) 511.
[31] M. Saboungi, M. Blander and A. Rahman, J. Chem. Phys., to be published.
[33] C.A. Angell, J.H.R. Clarke and L.V. Woodcock, Adv. Chem. Phys. 48 (1981) 397.
[34] D. Fincham and D.M. Heyes, Chem. Phys. 78 (1983) 425.
[35] J.E. Enderby D.M. North and P.A. Egelstaff, Phil. Mag. 14 (1966) 961; J.E. Enderby and
G.W. Neilson, in: Water: A Comprehensive Treatise, Vol. 6, ed., F. Franks (Plenum, New
York, 1982).

C.A. A ngell / Spectroscopy simulation and scattering

17

[36] Y. Derrein and J. Dupuy, J. Phys. Paris 36 (1975) 191.


[37] A.C. Wright, G. Etherington, J.A. Erwin Desa and R.N. Sinclair, J. Phys. Colloq. 43 (1982)
C9.
[38] P.H. Fuoss, P. Eisenberger, W.K. Warburton and A. Bienenstock, Phys. Rev. Lett. 46 (1981)
1537.
[39] T. Srinivasan, H. Egger, H. Pummer and C,K. Rhodes. IEEE J. Quantum Electronics QE-19
(1983) 1270.
[40] C.K. Rhodes, private communication.
[41] A. Barkatt and C.A. Angell, J. Phys. Chem. 82 (1978) 2622.
[42] A. Barkatt and C.A. Angell, J. Chem. Phys. 70 (1979) 901.
[43] C. Hunter and C.A. Angell, to be published.
[44] L.E. Busse, Phys. Rev. B29 (1984) 3639.
[45] J. Thomas, L.A. Bursill and K.J. Rao, Nature 89 (1982) 157; L.A. Bursill, and J.M. Thomas, J.
Phys. Chem. 85 (1981) 3007.
[46] P.H. Gaskell, contribution to NATO ASI on: Glass, Current Aspects (Plenum, New York~ to
be published).
[47] J.P. de Neufville and H.K. Rockstad, in: Proc. 5th Int. Conf. on Amorphous and Liquid
Semiconductors, ed. G. Lucovski (Taylor and Francis, New York, 1973) p. 420.
[48] R. Calemczuk and E. Bonjour, J. Non-Crystalline Solids 43 (1981) 427; R. Calemczuk, private
communication.
[49] C.A. Angell and D.R. MacFarlane, Advan. Ceram. 4 (1981) 66.
[50] A.F. Wright, contribution to NATO ASI on: Glass Current Aspects (Plenum, New York, to
be published).
[51] L.R. Angel, D.F. Evans and B.W. Ninham, J. Phys. Chem. 87 (1983) 538.
[52] D.L. Fields and C.A. Angell, to be published.
[53] J. Peyrelasse, C. Boned, J. Heil and M. Clausse, J. Phys. C15 (1982) 7097.
[54] D.R. MacFarlane and C.A. Angell, J. Phys. Chem. 86 (1982) 1927,
[55] C.A. Angell, R.K. Kadiyala and D.R. MacFarlane, J. Phys. Chem. 88 (1984) 4593: J.
Oubochet, J. Teixeira, R.K. Kadiyala, C.M. Alba, D.R. MacFarlane and C.A. Angell, J. Phys.
Chem. 88 (1985) 6727 (Debye Memorial Issue).
[56] M. Hoare, Ann. N.Y. Acad. Sci. 279 (1976) 186.

Anda mungkin juga menyukai