Vous êtes sur la page 1sur 10

Eur. J. Mineral.

This paper has been presented at the ECROFI XIII


1996, 8, 987-996 Symposium in Sitges, Spain (June 1995)

Formation of primary fluid inclusions under influence


of the hydrodynamic environment

MANUEL PRIETO*, ANDRES PANIAGUA and CELIA MARCOS

Departamento de Geologia, Universidad de Oviedo, C/ Arias de Velasco, s/n,


33005-Oviedo, Spain

Abstract: When crystals grow under convection conditions, the hydrodynamic environment can promote the
formation of inclusions in the rear side of the crystals, downstream in relation to the fluid flow. An experimental
study of this mechanism of fluid trapping is presented, and the conditions under which hydrodynamics influences
growth behaviour are discussed. Crystal growth experiments have been carried out with aqueous solutions in a
convection system under controlled conditions of temperature, solution velocity V* and supersaturation Goo. Growth
rates of {011} faces of ADP crystals have been measured for different orientations, in relation to the solution flow
and for different values of o<×, and V*. It is concluded that hydrodynamics can produce an anisotropy in the
contribution of material to crystallographically equivalent faces. At low solution velocities and moderate supersat-
uration, the growth rate depends on the flow "impact" angle. In contrast, at high solution velocities (Vs > 3 cm.s 1
for Goo = 2.45 %) the hydrodynamic configuration has no influence on the growth kinetics.
Hydrodynamics can also play an important role in the formation of fluid inclusions during crystal growth.
However, this effect only occurs at low solution velocities, when the growth kinetics is controlled by the hydrody-
namic configuration. Crystals grown at low solution velocities (Vs < 3 cm.s"1 for Goo = 2.45 %) show large fluid
inclusions in the rear zone. At high solution velocities the mass-transfer is high enough to avoid the hydrodynamic
control of growth, and no inclusions are generated by this mechanism.

Key-words: fluid inclusions, crystal growth, hydrodynamics, supersaturation.

Introduction dividuals. The trapping of fluid may also occur


under certain hydrodynamic conditions which
The various ways of fluid trapping during the bring about an uneven distribution of inclusions.
crystal growth process have been widely dis- The formation of inclusions in the rear side of
cussed in the literature (Deicha, 1955; Chernov the crystal, downstream in relation to the fluid
& Temkin, 1977; Roedder, 1984). The presence flow, has been noticed in crystals grown from
of bubbles or foreign matter, different "capping" solution under poor stirring conditions (Carlson,
processes, subsequent dissolution and growth, 1958). Janssen-van Rosmalen et al. (1978)
and loss of stability by the crystal-growth form observed the occurrence of liquid inclusions in
resulting in dendritic growth are the most com- the rear side of KDP (KH2PO4) crystals growing
mon causes promoting the formation of primary from solution at atmospheric pressure. These
intracrystalline inclusions. In general, all these authors placed crystalline seeds on a tree rotating
mechanisms lead to random and sparse distribu- in a KDP mother solution, and carried out growth
tions of inclusions within the crystalline in- experiments at different rotational speeds and at

e-mail: mprieto@ asturias.geol.uniovi.es

0935-1221/96/0008-0987 $ 2.50
© 1996 E. Schweizerbart'sche Verlagsbuchhandlung. D-70176 Stuttgart
988 M. Prieto, A. Paniagua, C. Marcos

a constant supersaturation. As a result, they only which is thermostated at a lower temperature. As


observed the formation of inclusions when the a result, the solution becomes supersaturated, and
crystals grew at low rotational speeds. Gits-Léon a crystalline seed, located in this cell, will grow.
et al. (1978) studied by means of X-ray topogra- Finally, the solution returns to the saturation
phy the effect of stirring on the quality of potash- reservoir. The growth cell is made of glass, and
alum crystals grown from solution, observing the the crystal seed is held on a platinum wire. The
formation of inclusions in the rear part of the cell is mounted on the plate of a photomicro-
crystals but also dislocations in the front part. scope, in such a way that the growth rate can be
The effect was demonstrated to be especially ap- measured by taking photographs at regular inter-
parent at low rotational speeds. vals of time.
Therefore, the hydrodynamic environment The bulk relative supersaturation in the growth
seems to give rise to the formation of fluid inclu- cell is given by:
sions under certain growth conditions. At present, _ Co (Γi) - Co (72) (j)
however, there is no systematic study on the cou- Co (Γ 2 )
pling between hydrodynamics and supersatura- where c0(T\) and c0(Ti) are the saturation con­
tion that produces configurational effects and the centrations at the temperature T\ of the mother
trapping of fluid. Here, we present an experimen- reservoir and at the temperature Ti of the growth
tal study on the influence of both factors, on the cell, respectively. Obviously, in order to maintain
assumption that this mechanism of inclusion a constant supersaturation level, crystals of solute
development, although requiring very specific must always be dissolving in the saturating ves­
crystallization conditions, could play a role in sel. The pumping system works at different circu­
some natural environments. lation velocities which can be measured by
With this aim, ADP crystals were grown from means of an "on-line" flow meter. The apparatus
solutions in a controlled convection system. Al- therefore allows us to control the bulk supersat­
though these compounds have no natural coun- uration Goo, the solution velocity Vs and the crys­
terpart, Kihle & Johansen (1994) enumerate a set tallization temperature (± 0.052C).
of reasons for the use of KDP-type crystals to ADP (NH4H2PO4) crystals were grown at Ti =
investigate fluid inclusions in the laboratory. The 252C and at their natural pH. The supersaturation
main reason for chosing these compounds in the was modified by using different saturation
present study lies in the facility to grow large temperatures (T\ > 25 9 C). Moreover, for each
crystals at low temperature and atmospheric pres- specific supersaturation a set of experiments was
sure, conditions that are necessary if one does not carried out at different solution velocities. Idio-
want unattainable experimental designs to control morphic single crystals (4-5 mm long) were used
the hydrodynamic environment. as crystal seeds. KDP-typecrystals belong to the
tetragonal space group I42d. The ideal mor­
phology consists of prismatic {010} and py­
Experimental procedures ramidal {011} forms. Sometimes, moreover, the
and starting materials crystals show high-index {Okl} forms that are as­
Crystal growth experiments were carried out in sociated to the so-called "tapering" effect (Dam
a controlled convection system. In such a system, et al., 1986). The crystals are slightly elongated
a saturated solution is pumped from a thermo- along [001] as a consequence of the higher
stated reservoir to a crystallization cell (Fig. 1) growth rate of the pyramid faces. At low super-

Fig, l. Scheme of the crystallization cell.


Formation of inclusions under hydrodynamic influence 989

saturations, the growth rate of the prismatic faces


is difficult to assess unless the experiments go on
for a long time. For this reason in the present
work the growth-rate study was devoted to the
pyramid faces.
Although the experiments were carried out
over a wide range of solution velocities, the flow
in the growth cell was always laminar. This fact
was checked by means of flow visualization
experiments. Small amounts of glycerine were
added to the solution to trace the bulk flow in the
growth cell as well as the "eddy" zone formed
behind the crystal, around the face placed in the
"shade position" in relation to the flow.
After the growth process, the crystals were re-
covered from the cell and characterized by opti-
cal microscopy and scanning electron micro-
scopy (SEM). The volume fraction Vv of
inclusions (volume of inclusions per crystal
volume unit) was determined from surface meas-
urements by a stereological method (Weibel,
1989). Regularly spaced sections of the crystals
were observed by SEM and for each section /,
0.25
the surface fraction (AAM of inclusions (relative
area of inclusions per unit area of the section)
was determined by image processing. Finally, the
volume fraction of inclusions was estimated from
0.20-1 0
0.15
the surface fractions corresponding to a given
crystal, according to the expression: E 0.10 V s (cm.s- 1 )
N
E
Vv = ÄA = {\IN)Σ{AA)i, (2) <P 0.05^
i=i
where N is the number of crystal sections con­ 0.00
sidered. This equation states that the average o 1
areal fraction AA, determined on sections through Fig. 2. Influence of the solution velocity on the growth
a volume, represents an estimate of the volume rate of faces {011} of ADP crystals (Squares, circles,
fraction Vv of the phase under investigation. The and triangles represent faces in position 1, 2, and 3,
statistical error in the estimation of Vv depends respectively), a) α<×> = 2.45 %. b) Goo = 1.71 %. c) Goo =
on the number of interceptions of inclusions ob­ 1.2 %.
served (DeHoff & Rhines, 1968), and will be
shown graphically for each specific experiment.
is in the "shade position" in relation to the_bulk
stream. As the pyramid faces (011) and (Oil) of
ADP crystals are nearly perpendicular (89.609),
Hydrodynamics and growth kinetics the three orientations can be studied simul-
taneously by arranging the crystal as shown in
Growth rate measurements Fig. 1. For each specific value of Goo and Vis, the
growth rates RGI, RG2 and RG3 were measured,
Linear growth rates of pyramid {011} faces of corresponding to orientations 1, 2 and 3. The
ADP were measured for different values of Goo results are shown in Fig. 2. In these plots, each
and Vs. Measurements were carried out for three point represents the mean value of 5 experimen-
different orientations in relation to the flow. tal runs. Dispersions from the mean values are
Orientations 1 and 2 are, respectively, perpendic­ shown in the figure by means of error bars .
ular and parallel to the flow, while orientation 3 Fig. 2a represents the dependence of the
990 M. Prieto, A. Paniagua, C. Marcos

growth rate on the solution velocity at a constant the increase of the solution velocity, the growth
bulk supersaturation (Goo = 2.45 %). At low so­ rate increases up to approaching a certain con­
lution velocities the growth rate depends on the stant value at high Vs values (Mullin & Garside,
face orientation, according to the ranking RGI > 1967; Garside, 1971; Rodriguez-Clemente, 1976;
RG2 > RG3. However, as Vs increases, the values Garside, 1991). This means that at solution
of RG for the three orientations tend to be equal. velocities higher than a certain level, the growth
Moreover, at high solution velocities, the curves is no longer controlled by the volume transport
RG-VS show an asymptotic trend: the growth rate but rather by the surface processes. Moreover, the
increases with the solution velocity up to a cer­ value of the solution velocity required in order to
tain level, after which further increases in Vs have reach the asymptotic level of constant growth
no important effect on RG. rate increases as the supersaturation increases.
The results are analogous at a lower supersat­ Both effects are clearly shown in Fig. 2.
uration (Goo =1.71 %), but in this case the curves Obviously, hydrodynamics will only be relevant
corresponding to RGI, RG2 and RG3 are closer for growth phenomena at low solution velocities,
together (Fig. 2b). Finally, at a very low super- when convection is the rate-controlling process.
saturation (Goo = 1.20 %), it is impossible to dis­ For this reason RGU RG2 and RG3 tend to con­
tinguish the three curves (Fig. 2c), i.e., the width verge as the solution velocity approaches the
of the range RGI-RG3 decreases as the supersatu­ asymptotic level, and the process gradually
ration decreases. Moreover, by comparing Fig. 2a, becomes surface-integration controlled. For a
2b and 2c, it is clear that the asymptotic value of specific solution velocity, the higher the super-
RG decreases as the bulk supersaturation saturation, the higher the effect of the hydrody­
decreases. namic environment on the growth kinetics is (the
higher the difference between RGI, RG2 and RG3)-
At very low supersaturation (Fig. 2c), the growth
Discussion: growth rate and flow configura­ is controlled by the surface processes, even at
tion low solution velocities, and the face orientation
has no influence on the growth rate.
We can conclude from Fig. 2 that, when a The dependence of the growth rate on the
crystal grows under convection conditions, the flow-impact angle at low solution velocities can
hydrodynamic configuration can produce an ani- be explained by boundary-layer effects. The in­
sotropy in the contribution of material on crys- corporation of growth units to the solid involves
tallographically equivalent faces. At low solution a decrease of the solute concentration near the
velocities and moderate supersaturations, the interface, and the subsequent development of a
flow "impact" angle has an important bearing on "concentration boundary layer" of thickness bc
the linear growth rate. In contrast, at high solu­ around the crystal surface. Outside the boundary
tion velocities the hydrodynamic configuration layer the fluid is homogeneous in concentration,
seems to have no influence on the growth kinet­ but within the layer there is a concentration
ics. decrease from the bulk value Coo to an interface
These results are not surprising. Crystal growth value a. Therefore, one must distinguish between
from solution is a complex process that involves bulk supersaturation Goo and interface supersatu­
several interrelated subprocesses. First of all, ration G/.
crystallization involves the transport of solute The thickness of the concentration film de­
from the solution bulk to the crystal interface. pends on the degree of agitation of the system,
However, unlike other mass-transfer processes and can be formalized for simple flow configu­
where only the volume mass-transfer determines rations by defining another boundary layer, viz.
the rate of the process, in crystal growth there are the "hydrodynamic or velocity boundary layer".
surface-diffusion phenomena that can be rate- This is a thin region near the solid-liquid inter­
determining. In a general way, the linear growth face where the flow becomes reduced owing to
rate RG of a crystalline face depends on the sur­ the influence of the surface, i.e., where the solu­
face roughness or surface-entropy factor α tion velocity undergoes a continuous change
(Bourne & Davey, 1976; Bennema & van der from its value Vs in the solution bulk to stagna­
Eerden, 1977), the stirring or solution velocity Vs, tion. The thickness 0// of this layer depends on
and the bulk supersaturation Goo. For specific both the solution viscosity and the specific flow
values of α and Goo, it has been proved that with configuration around the crystal surface. So, as
Formation of inclusions under hydrodynamic influence 991

Fig. 3. Hydrodynamic boundary layers. Configurations for flow (a) perpendicular and (b) parallel to the crystal
interface.

shown in Fig. 3a, when a laminar fluid flow the relative magnitudes of 8// and 8 c may be
impinges perpendicularly to a crystal interface roughly estimated from (Mullin, 1993)
the value 8// is uniform, and is given by ç / \l/3
(Rosenberger, 1979): OH
(5)
λ l/2
D
8//= 2.4 which is the ratio of the kinematic viscosity to
a (3)
the diffusivity D of the solute in the solution. For
aqueous solutions v/D » 1 , which means that 8//
where v is the kinematic viscosity and a is a con­ exceeds 8 c considerably. From equations (3) to
stant. (5), expressions for the value of the boundary-
The situation is similar when the flow im­ layer thickness can be deduced. So, for a face
pinges parallel to the crystal face, but in this case with orientation 2, parallel to the fluid flow, Sc
the thickness of the hydrodynamic boundary decreases with xl/2 from the leading edge:
layer is not uniform. It increases with the
fDWr / Λ1/2
distance x, downstream, from the leading edge 8c = 3 (6)
( F i g . 3 b ) , a c c o r d i n g to the e x p r e s s i o n Vs
(Rosenberger, 1979): Obviously, when the growth process is control­
x l/2 led by the volume mass-transfer, a thick concen­
8/f = 3 vx (4) tration boundary layer involves a slow growth
Vs
process, and vice versa, i.e.:
For the simple configurations described above, RG oc 1/8C. (7)
992 M. Prieto, A. Paniagua, C. Marcos

Fig. 4. Flow configuration in the proximity of the rear Fig. 5. Fluid inclusions in a crystal of ADP. (Ooo = 2.45 %;
side of a crystal (Vs = 1.5 cm.s 1 ). Vs= 1.5 cm.s" ). The contour of the crystalline seed can
be observed in the center of the crystal.

It is clear from equations (3) to (6) that both


3H and ô c depend on the specific flow configura- and the eddies become free. As an illustration,
tion around the crystal surface. These values can Fig. 4 shows this region as observed in a flow
be predicted theoretically if the nutrient flow is visualization experiment. Obviously, a formal
fully understood, as has been shown for faces expression of 8// for such a configuration cannot
with orientations 1 and 2. For example, for KDP be calculated. Anyway, the thickness of the
crystals in saturated solution at 36 2 C, D = concentration boundary layer seems to depend
7.8 × 10-10 m 2 s - l ; v = L 1 1 3 × 1 0 -6 m 2 s - l 5 v /£) = on the face orientation, according to the ranking
1427.17, and ô///ôc - 11.26. Therefore, for Vs = ôc/ < &c2 < ôcj, and consequently, the ranking
0.25 cm.s 1 , at 1 mm from the leading edge of a RGI > RG2 > RG3 is to be expected. This fact
face parallel to the flow, ô// ~ 2002 µm and 5C ~ could give rise to an obvious morphologic effect:
178 µm. These values decrease as the solution homologous faces can show unequal develop-
velocity increases, being 67/ ~ 409 µm and ô c ~ ments as a result of their orientation in relation
36 µm at Vs = 6 cm.s 1 , which means an impor- to the fluid flow.
tant reduction in the mass-transfer resistance.
The value of ô c for a face in the "shade" Generation of fluid inclusions
position is not so easy to calculate. Behind the
crystal, at low solution velocities, a large, stable Experimental findings
and closed eddy zone is present. As the solution The u n e q u a l d e v e l o p m e n t of crystal-
velocity increases, the eddy zone becomes open, lographically equivalent faces is not the only

4. mm
Fig. 6. Fluid inclusions in the rear side of a crystal of ADP (GOO = 1.7 %; Vs = 1.5 cm.s"1).
Formation of inclusions under hydrodynamic influence 993

Fig. 7. Development of downward macrosteps on a face


(010) parallel to the flow.
50-

^ 40« %
morphologic effect of the hydrodynamic environ­ -| o„=2.45%
ment. ADP crystals grown at low solution veloci­ "5
ties show large fluid inclusions in the rear zone, ^,30-
(O
beyond a critical point on the surface in relation
to the fluid flow. The front part of the crystals is
§ 20-
'5>
X
free of this kind of inclusion. The phenomenon
is shown in Fig. 5 and 6, which also display the
arrangement of the crystals during the growth
•§ io-

oJ— ^ F — i I —-*F—^
X
i *•* • • 1
a
• 1
process. 0 1 2 3 4 5
A more detailed observation of Fig. 5 enables
us to distinguish two different types of inclu­ Vs(cm.s-1)
sions: the solitary inclusion labeled "a", and the
bunch labelled "b". The inclusion "a" is shallow, Fig. 8. Volume fractions of inclusions of ADP crystals.
unfolds parallel to a prismatic face, and has a a) Influence of the supersaturation (Vs = 1 . 2 cm.s"1).
large surface development. Generally, this kind b) Influence of the solution velocity (Goo = 2.45 %).
of inclusion is associated with downward macro-
steps, which develop downstream on the {010}
faces (Fig. 7). of inclusions decreases as the solution velocity
Type "b" inclusions are situated in the inner increases (Fig. 8b), until at high solution veloci-
part of the rear half of the crystals. The arrange­ ties (Vs > 3 cm.s"1), the volume of inclusions be-
ment of these inclusions seems to retain a certain comes negligible.
parallelism to the {011} faces. They can form It is worth noting that no correlation is made
parallel, eventually interconnected, sets (Fig. 6), in Fig. 8 between RG and the volume fraction of
or display more irregular shapes (Fig. 5). The inclusions. This is because growth rate measure-
shape of these inclusions, as well as their size, ments were carried out in separate experiments
depends on the growth conditions, i.e., on both and for a different arrangement of the crystals
solution velocity and super saturation. (see Fig. 1 and 5). However, some additional
The influence of both parameters on the rela­ conclusions can be drawn by comparing Fig. 2a
tive amount of inclusions is shown in Fig. 8, and 8b with the same supersaturation (Goo =
where the ordinate refers to the volume fraction 2.45 %). Fig. 2a indicates that for Vs > 3 cm.s-1
of inclusions Vγ in the rear half of the crystal. As the growth rates approach the asymptotic level
shown in Fig. 8a, at a constant solution velocity (RGI = RG2 = RG3), and the volume fraction of
(Vs = 1.2 cm.s-1) , the volume fraction of in­ inclusions becomes negligible. In contrast, at
clusions increases as supersaturation increases up low solution velocities RGU RG2, and RG3 differ
to a limiting value, nearly 50 vol. %. For a con­ and Vv takes high values. Moreover, from Fig. 2c
stant supersaturation (α<×, = 2.45 %), the amount and 8a, it is clearly seen that at Vs = 1.2 cm.s 1
994 M. Prieto, A. Paniagua, C. Marcos

and Goo = 1.2 % the volume fraction of inclusions flat inclusion parallel to the growing face, i.e., to
is negligible and RGI = RG2 = RG3. That is the a type "a" inclusion.
reverse for higher supersaturations (Fig. 2a Although both kinds of inclusions are formed
and 2b) where RGI > Ra > RG3 and Vγ takes under the same growth conditions, type "b"
values approaching 45 vol. %. inclusions seem to have a different genetic mech­
One can therefore conclude that the coupling anism. According to Janssen-van Rosmalen et al.
between low solution velocity and moderate (1978) this kind of inclusion is due to solute-
supersaturation promotes the formation of fluid concentration depletion in the proximity of
inclusions. The effect is so determining that, the pyramidal terminations of the growing crys­
under suitable conditions, the hydrodynamic con­ tal. Behind the crystal, at low solution velocities,
figuration may even lead to the formation of open a large and stable eddy is present. Since the eddy
holes. is closed, exhaustion of its solution occurs, which
will give rise to the formation of fluid inclusions
Discussion: hydrodynamic mechanisms of in­ along the rear side of the crystal. The situation
clusion formation will be similar at intermediate solution velocities,
when the eddy zone becomes open, but in this
Hydrodynamics seems to play an important case the concentration gradient in the vicinity of
role on the formation of fluid inclusions during the rear side will be smaller, and the effect will
crystal growth. However, this effect is only be less pronounced. Finally, at high solution
important under certain conditions, when the velocities the mass transfer will be high enough
growth process is controlled by volume mass- to avoid the hydrodynamic control of growth,
transfer. This is clear from Fig. 2a and 8b, which and no inclusions will be generated by this mech­
show the influence of the solution velocity on anism.
both the amount of inclusions and the growth It is worth noting that the control of growth by
rate: as Vs increases, the growth kinetics becomes the volume mass-transfer does not mean that the
controlled by the surface processes, and the surface phenomena are not operative. Crystal
influence of the hydrodynamic environment growth depends on the surface roughness, and
becomes negligible, whereas at low solution always involves the surface diffusion of growth
velocities the growth kinetics is mainly deter­ units until they can enter a growth kink on the
mined by the volume transport, and two different crystal surface. However, we can only speak of
mechanisms of inclusion formation seem to kinetic control by the surface processes when the
operate. transport of solute molecules from the bulk fluid
Inclusions of type "a" develop on a face par­ to the crystal interface is high enough to supply
allel to the flow of solution. For such an arrange­ all the incorporated growth units. The concentra­
ment, both 0// and Oc increase downstream (Eqs. 4 tion at the crystal surface is then assumed equal
and 6), and the interface supersaturation to that of the bulk fluid, and hydrodynamics
decreases with JΓ1^2 from the leading edge. This becomes irrelevant to the growth process.
implies that growth steps generated at the leading
edge decrease their advancement rate as they pro­
gress downstream on the surface. As a con­
s e q u e n c e , the steps will tend to pile up Conclusions and geological implications
downstream, beyond a critical point on the sur­
face. This could explain the presence of large In this paper we have outlined two different
downward macrosteps on the faces parallel to the mechanisms of primary fluid inclusion formation
flow (Fig. 7). At the same time, these macrosteps and discussed the factors which enable the hy­
can give rise to the development of type "a" drodynamic environment to influence the growth
inclusions on these faces. A downward macro- kinetics. The flow configuration seems to
step produces a sort of "shade" effect for the influence the formation of inclusions and other
stream lines, which promotes the establishing of morphologic effects during crystal growth from
a small "dead fluid" region in the vicinity of the fluid phases. The conditions under which these
step rise. During growth, the nutrition of the step effects appear have been established from
rise is better in the upper part than in the lower growth-rate measurements on faces with different
part, and a layer overhanging the growing face flow "impact" angles. The effect is so critical
will appear. This overhang will give rise to a that, under suitable conditions, the hydrodynamic
Formation of inclusions under hydrodynamic influence 995

environment may even lead to the formation of boundary-layer effects in the formation of both
open holes in the rear side of the crystals. aqueous inclusions and silicate melt inclusions.
Our results are obviously limited to laboratory The "ability" of a mineralizing fluid to form con-
experiments with KDP-type crystals, but their centration boundary layers depends on both the
high reproducibility allows us to suppose that viscosity and the diffusion coefficient of the
these phenomena can be effective over a wide growth units at the crystallization temperature. In
range of substances. If a similar asymmetrical low-temperature aqueous systems, considerable
distribution of inclusions is observed on crystals concentration gradients can form near the crystal
whose growth conditions are unknown, for interface. However, such effects have not been
example on natural crystals, a flux direction studied for hydrothermal crystallization at
could be suggested. temperatures where the diffusion is fast. Diffu-
Although the experiments described in this sion rates in viscous fluids such as silicate melts
paper were carried out under specific steady-state are much lower than in aqueous solutions, so the
conditions, the conclusions can be extended to boundary-layer effects might be expected to be
natural systems with configurational similarities, more important. Furthermore, as the ability to
regardless of the specific mode of supply of min- form boundary layers increases with the viscos-
eralizing material. In a general way, it is easy to ity, boundary layers in rhyolitic melts should be
imagine flow configurations in natural environ- larger than those found in basaltic melts, because
ments that involve unidirectional mass-transfer: the viscosities of rhyolites are orders of magni-
advection and directional diffusion through tude higher.
porous media, convection in magmatic fluids, A final problem is the degree of supersatura-
flow in hydrothermal veins or in low-temperature tion required for the kinetic control of the growth
hydrothermal systems in which the flow direction process by the volume mass-transfer. There is a
is controlled by the orientation and permeability tacit assumption that natural crystallization oc-
of the strata, etc.. Examples of morphological curs near equilibrium, and there remains scepti-
phenomena induced by the flow direction are cism in accepting that super saturation is relevant
relatively common in the literature, but only to natural fluids. However, there is unequivocal
scarce attention has been paid to the formation of evidence that supersaturated fluids can exist over
inclusions by this mechanism. Grigor'ev (1965) a wide variety of geological environments. For
has reviewed different cases of non-uniform crys- instance, in low-temperature sedimentary en-
tal growth by a unidirectional supply of the min- vironments authigenic feldspars can crystallize
eralizing material. For instance, trough-like crys- with almost any state of (Al, Si) order, depending
tals of beryl in some pegmatite veins are on the supersaturation of the fluid (Kastner &
non-uniformly developed due to the direction of Siever, 1979). In hydrothermal environments, the
the solution flow: the growth was inhibited in the formation of veins of adularia with (Al, Si) order
rear side of the crystals, where the solutions were ranging from high-sanidine to maximum-micro-
i m p o v e r i s h e d in c o n c e n t r a t i o n . Quartz cline testifies to the possible range of supersatu-
phenocrysts in rhyolites are frequently oriented ration in the parent fluid phase. There have been
subparallel to the magmatic flow, and in such numerous studies demonstrating a relationship
cases show inclusions and open holes systemati- between morphology and undercooling degree in
cally situated on the rear side. igneous melts (Muncill & Lasaga, 1988). The
In recent years, an important theoretical frame- presence of non-equilibrium internal morpholo-
work has been developed to characterize the gies, like sector and oscillatory zoning (Prieto et
mass flow in natural systems (Foster & Smith, al., 1993), provides another argument supporting
1990; Tait & Jaupart, 1990; Nicholls, 1990). the existence of high super saturations in natural
Numerical modeling carried out within such a fluids, either aqueous or magmatic. Putnis et al.
framework provides a rigorous basis for estab- (1995) have outlined a range of causes and re-
lishing the likelihood for a postulated flow viewed natural environments where the existence
system to develop. However, these studies are of supersaturated fluids has been proved.
u s u a l l y d e v o t e d to m e s o - or large-scale All these arguments support the possibility of
processes. Micro-scale studies, such as the study the volume mass-transfer control of crystal-
of the formation of boundary layers near the in- lization in some natural environments, which is
terface of crystallizing minerals, are rare. imperative for the formation of fluid inclusions
Roedder (1984) has reviewed the role of the by the influence of flow configuration. However,
996 M. Prieto, A. Paniagua, C. Marcos

the geological implications of the issues we have the hydrodynamic environment on the growth and
raised here have yet to be fully tested, either ex- the formation of liquid inclusions in large
perimentally or in the field. potassium dihydrogen phosphate (KDP) crystals.
Kristall und Technik, 13, 17-28.
Acknowledgements: This work was supported Kastner, M. & Siever, R. (1979): Low temperature
feldspars in sedimentary rocks. Amer. Jour. Sci.,
by DGCYT (Ministry of Science and Education
279, 435-479.
of Spain), Grant PB92-0998. Drafts of this man-
Kihle, J. & Johansen, H. (1994): Low-temperature
uscript were much improved by the careful re- isothermal trapping of hydrocarbon fluid inclusions
view by R.J. Bakker, and an anonymous referee. in synthetic crystals of KH2PO4. Geochim. Cosmo-
chim.Acta, 58, 1193-1202.
Muncill, G.E. & Lasaga, A.C. (1988): Crystal-growth
References kinetics in igneous systems: isothermal H2O-satu-
rated experiments and extension of a growth model
Bennema, P. & van der Eerden, J.P. (1977): Crystal to complex silicate melts. Am. Mineral., 73, 982-
growth from solution: development in computer 992.
simulation. / . Cryst. Growth, 42, 201-213. Mullin, J.W. (1993): Crystallization. Butterworth-
Bourne, J.R. & Davey, R.J. (1976): The role of solvent- Heinemann, Oxford, 527 p.
solute interactions in determining crystal growth Mullin, J.W. & Garside, J. (1967): Crystallization of
mechanisms from solution. I. The surface entropy aluminium potassium sulphate: a study in the
factor. J. Cryst. Growth, 36, 278-286. assesment of crystallizer design data: I: Single
Carlson, A. (1958): The fluid mechanics of crystal crystal growth rates, II: Growth in a fluidized bed.
growth from solution, in "Growth and perfection of Trans. Inst. Chem. Engrs., 45, 285-295.
crystals", R.H. Doremus, ed. J. Wiley and Sons, Nicholls, J. (1990): The mathematics of fluid flow and
New York, 368-373. a simple application to problems of magma trans-
Chernov, A.A. & Temkin, D.E. (1977): Capture of in- port, in "Modern methods of igneous petrology:
clusions in crystal growth, in "Crystal growth and understanding magmatic processes", J. Nicholls &
materials", E. Kaldis & H J . Schoel, eds. North J. K. Russel, eds. Mineral. Soc. Am., Reviews in
Holland, Amsterdam, 3-77. Mineralogy, 24, 107-124.
Dam, B., Bennema, P., van Enckevort, WJ.P. (1986): Prieto, M., Putnis, A., Fernández-Diaz, L. (1993):
The mechanism of tapering on KDP-type crystals. Crystallization of solid solutions from aqueous so-
J. Cryst. Growth, 74, 118-128. lutions in a porous medium: zoning in (Ba, Sr)SO4.
DeHoff, R. T. & Rhines, F.N. (1968): Quantitative mi- Geol. Mag., 130, 289-299.
croscopy. McGraw-Hill, New York, 404 p. Putnis, A., Prieto, M., Fernández-Diaz, L. (1995): Fluid
Deicha, A. (1955): Les lacunes des cristaux et leur in- supersaturation and crystallization in porous media.
clusions fluides. Masson and Cie, Paris, 285 p. Geol. Mag., 132, 1-13.
Donaghey, L.F. (1980): Hydrodynamics of crystal Rodriguez-Clemente, R. (1976): Crystal growth kinet-
growth processes, in "Crystal growth", B. Pamplin, ics of sodium chloride from solution, in "Industrial
ed. Pergamon Press, New York, 65-101. crystallization", J.W. Mullin, ed. Plenum Press,
Foster, C. & Smith, L. (1990): Fluid flow in tectonic New York, 187-199.
regimes, in "Short Course on fluids in tectonically- Roedder, E. (1984): Fluid inclusions. Mineral. Soc.
active regimes of the continental crust", B.E. Am., Reviews in Mineralogy, 12, 664 p.
Nesbitt, ed. Mineralogical Association of Canada, Rosenberger, F. (1979): Fundamentals of crystal
Vancouver, 1-47. growth I. Macroscopic equilibrium and transport
Garside, J. (1971): The concept of effectiveness factors concepts. Springer-Verlag, Berlin, 530 p.
in crystal growth. Chem. Eng. Sci., 26, 1425-1431. Tait, S. & Jaupart, C. (1990): Physical processes in the
— (1991): The role of transport processes in crystal- evolution of magmas, in "Modern methods of ig-
lization, in "Advances in industrial crystallization", neous petrology: u n d e r s t a n d i n g magmatic
J. Garside, R. J. Davey, A. G. Jones, eds. Butter- processes", J. Nicholls & J.K. Russel, eds. Mineral.
worth-Heinemann, Oxford, 92-104. Soc. Am., Reviews in Mineralogy, 24, 125-190.
Gits-Léon, S., Lefaucheux, F., Robert, M.C. (1978): Ef- Weibel, E.R. (1989): Measuring through the micro-
fect of stirring on crystalline quality of solution scope: development and evolution of stereological
grown crystals-case of potash alum. / . Cryst. methods. / . Microsc., 155, 393-403.
Growth, 44, 345-355.
Grigor'ev, D.P. (1965): Ontogeny of minerals. Israel
Program for Scientific Translations, Jerusalem, 249 p. Received 22 September 1995
Janssen-van Rosmalen, R., van der Linden, W.H., Modified version received 1 April 1996
Dobbinga, E., Visser, D. (1978): The influence of Accepted 23 May 1996

Vous aimerez peut-être aussi