Você está na página 1de 5

GEOPHYSICAL RESEARCH LETTERS, VOL. 38, L13316, doi:10.

1029/2011GL047167, 2011

The effect of particle shape on suspension viscosity


and implications for magmatic flows
S. Mueller,1 E. W. Llewellin,2 and H. M. Mader1
Received 16 February 2011; revised 13 April 2011; accepted 14 April 2011; published 6 July 2011.

[1] The rheology of crystalbearing magma and lava depends


on both the shape and volume fraction of the suspended
crystals. We present the results of analogue rheometric
experiments on monodisperse suspensions of solid particles
in a Newtonian liquid, in which particle volume fraction 
and aspect ratio rp are varied systematically. We find that
the effect of  on viscosity is well captured by the Maron
Pierce model, and that this model is valid across the range
of particle aspect ratios investigated (0.04 rp 22, i.e.,
stronglyoblate to stronglyprolate) when the maximum
packing fraction m is treated as a fitting parameter. The
value of m derived from fitting to our experimental data
depends strongly and systematically on particle aspect ratio;
hence, m represents an effective proxy for the influence of
particle shape on suspension rheology. We present a simple
relationship for m (r p ) which allows the viscosity of a
suspension to be calculated as a function of  and rp only.
We investigate the impact of accounting for crystal shape
when modelling volcanic flows by simulating the eruption
of magma carrying crystals of different aspect ratio, and
conclude that the effect of crystal shape should not be
neglected. Citation: Mueller, S., E. W. Llewellin, and H. M. Mader
(2011), The effect of particle shape on suspension viscosity and implications for magmatic flows, Geophys. Res. Lett., 38, L13316,
doi:10.1029/2011GL047167.

1. Introduction and Theoretical Background


[2] The flow of particle suspensions is central to volcanic
and magmatic processes throughout the volcanic system;
examples include the mingling of phenocrystbearing magmas, the eruption of microlitebearing magma, and the flow
of crystalbearing lava down a volcanos flank. Accurate
modelling of such flows depends on detailed knowledge
of the suspensions rheology. It is wellknown from laboratory rheometry of magmatic suspensions that the volume
fraction of suspended crystals exerts a firstorder control on the
effective viscosity of a magma [e.g., Lejeune and Richet, 1995;
Caricchi et al., 2007; Ishibashi and Sato, 2007; Lavallee et al.,
2007], hence on eruption dynamics [Melnik and Sparks, 2005].
In this article, we demonstrate that the influence of crystal
shape is similarly strong, and should not be neglected. We
perform rheometric experiments using presheared analogue
particle suspensions; whilst analogue experiments do not
capture processes such as crystal breakup and pressure solution, they have the advantage over experiments with magmatic
suspensions that they are better characterized and allow more
1
2

School of Earth Sciences, University of Bristol, Bristol, UK.


Department of Earth Sciences, Durham University, Durham, UK.

Copyright 2011 by the American Geophysical Union.


00948276/11/2011GL047167

thorough exploration of parameter space. Based on our data,


we define a relationship between the shape of suspended
particles, expressed as their mean aspect ratio, rp, and the
microstructural properties of the suspension, encapsulated by
the maximum packing fraction of particles, m. Since the
average aspect ratio of the dominant crystal phase in a magma
is often known, or can be estimated on the basis of petrological
models, the relation proposed here will facilitate the use of
rheological models in which m is a critical parameter and will
improve their accuracy by accounting for crystal shape.
[3] The rheology of a material is usually described by a
constitutive equation which relates the driving stress t to the
_ For a Newtonian liquid, this equation
resulting strain rate .
_ where m is the Newtonian viscosity, which is
is t = m,
independent of strain rate. Particlebearing liquids typically
develop a more complex, nonNewtonian rheology as particle concentration increases, becoming first shearthinning
at moderate volume fractions, then developing a yield stress
at high volume fractions [Mueller et al., 2010], though we
note that yield stress is not universally observed in experiments on natural samples [e.g., Caricchi et al., 2007]. This
nonNewtonian behaviour is welldescribed by the Herschel
Bulkley model [Herschel and Bulkley, 1926]:
 0 K _ n ;

where the consistency K is cognate with viscosity and has the


units Pa sn, the flow index n defines the degree of non
Newtonian behaviour (n < 1 for shear thinning behaviour),
and t 0 is the yield strength. For nonNewtonian liquids, the
viscosity is commonly expressed as an apparent viscosity
h = t/_ at a specified strain rate; expressed in these terms, the
HerschelBulkley model becomes:


0
K _ n1 :
_

For dilute and moderatelyconcentrated suspensions, where


yield stress is negligible [Mueller et al., 2010], it can be seen
that K is equivalent to h evaluated at _ = 1 s1. For weakly
shearthinning materials, i.e., when n is close to 1, K can
often be sufficiently well approximated by h. For suspensions, h is commonly nondimensionalized by the Newtonian
viscosity of the suspending liquid m0 to give the relative
viscosity hr = h/m0; by analogy a relative consistency can be
defined as Kr = K/m0.
[4] Particles affect the rheology of a suspension because
additional energy is dissipated due to fluidparticle and
particleparticle interactions. As particle volume fraction 
increases, these interactions become more common, leading
to a dramatic, nonlinear increase in relative viscosity until a
maximum packing fraction m is reached [e.g., Einstein, 1906,
1911; Roscoe, 1952; Maron and Pierce, 1956], at which point

L13316

1 of 5

L13316

L13316

MUELLER ET AL.: PARTICLE SHAPE AND SUSPENSION VISCOSITY

the suspension becomes jammed. If the particles are perfectly


rigid, flow ceases when m is reached; in magmatic suspensions, however, processes such as plastic deformation or creep
may take place at  > m [e.g., Lavallee et al., 2007; Caricchi
et al., 2007].
[5] Particle shape affects rheology by changing the nature
of fluidparticle and particleparticle interactions, introducing
two additional considerations: 1) anisometric particles are
orientable, so their influence depends on their orientation with
respect to the flow; 2) anisometry enhances particleparticle
interactions because the orbit of a rotating nonspherical particle encloses a greater volume for potential interactions than
that of a spherical particle. Previous theoretical and analogue
experimental work on the rheology of suspensions of anisometric particles is less extensive than for spherical particles,
and focusses primarily on dilute suspensions [e.g., Jeffery,
1922; Brenner, 1974; Powell, 1991; Pabst et al., 2006]. An
exception is our earlier study [Mueller et al., 2010] in which
we consider suspensions of anisometric particles with volume
fractions spanning the range 0  ] m; we found that the
following relationship holds across that range:

Kr

1


m

2

and that m varies with particle aspect ratio. Equation (3) is


based on the widelyapplied relationship of Maron and Pierce
[1956] (with Kr substituted for hr).
[6] For monodisperse spherical particles, the theoretical
maximum packing fraction m 0.74 for an ordered packing and m 0.64 for a random (i.e., disordered) packing
[Bernal and Mason, 1960]. The effect of particle shape on
maximum packing fraction in the absence of hydrodynamic
shearing has been investigated experimentally [e.g., Milewski,
1973; Parkhouse and Kelly, 1995; Rahli et al., 1999], geometrically [Evans and Gibson, 1986], and numerically [e.g.,
Williams and Philipse, 2003; Gan et al., 2004; Donev et al.,
2004]. In general, these studies show that m tends to
decrease with increasing particle anisometry. There have been
comparatively few studies of m for anisometric particles in
shearing flows; we are aware of just two, Kitano et al. [1981]
and Pabst et al. [2006], who both infer m from rheological measurements of sheared suspensions and find that m
decreases for increasingly prolate particles.
[7] In the present study we define the aspect ratio rp = la/lb,
where la is the particles axis of rotational symmetry and lb
is its maximum diameter perpendicular to that axis. Various
empirical and semiempirical models have been proposed
to describe the dependence of m on rp: Evans and Gibson
[1986], Rahli et al. [1999] and Mueller et al. [2010] propose
inverse relationships between the two parameters; Parkhouse
and Kelly [1995] propose a logarithmic relationship; and
Kitano et al. [1981] and Pabst et al. [2006] suggest a linear
decrease in m with rp. In this work, we use highresolution
rheometry to investigate the relationship between rp and m for
shearing flows, and compare existing models with our experimental data.

2. Experiments
[8] Rheometric measurements were performed on suspensions of particles in silicone oil (Cannon Viscosity Standard
N15000. A variety of particles were used to cover the range of

aspect ratios 0.04 rp 22 (i.e., the range most relevant to


magmatic suspensions): biotite and polyacrylic glitter (oblate);
glass beads (spherical); glass and carbon fibres (prolate). Examples are shown in Figure 1. Suspensions were prepared by
mixing a known mass of particles with a known mass of silicone oil. Samples were centrifuged to remove trapped gas
bubbles, and particle volume fractions were calculated using
the oil and particle densities, to an estimated accuracy of 3%.
Table 1 lists the main characteristics of the particles used in
this study.
[9] The rheology of the homogenized suspensions was
determined using a ThermoHaake Mars II rotational rheometer, in parallel plate arrangement with 35 mm sensor
diameter and 1.5 mm gap width. In order to avoid wallslip
effects with high viscosity samples, the surfaces of both the
upper and lower sensor were modified by attaching sandpaper with grit size of the order of the particle sizes. Flow
_ were obtained by running a 20step upramp
curves t ()
of incrementally increasing shear stress, followed by a 20
step down ramp. At each step the strain rate was recorded
when equilibrium stress conditions were reached. The maximum shear stress was usually set between 300 and 500 Pa. In
order to eliminate the transient effects observed during flow
initiation by Mueller et al. [2010], each flowcurve determination was preceded by a preshear treatment in which the
suspension was sheared to large strain at constant strain rate.

3. Results and Analysis


[10] Between four and eight flowcurve determinations
were performed for each particle aspect ratio, at varying particle concentrations . We used the HerschelBulkley model
(equation (1)) to fit the flow curve from each experiment.
A full table of results is included in the auxiliary material.1
Measurements revealing yield stress behaviour were discarded, hence t 0 = 0 for all reported experiments. Our data
show that shear thinning is weak (0.9 ] n 1) for all but the
most strongly anisometric particles used, and that the degree
of shear thinning is very sensitive to slight changes to the
suspension composition; hence, no simple correlation between
n,  and rp could be found. By contrast, Kr and  are strongly
correlated and are plotted in Figure 2 for suspensions of
prolate (Figure 2a) and oblate (Figure 2b) particles. The error
in Kr is calculated by combining the standard error obtained
from the fitting procedure with an estimate of the measurement error (3%). Following Mueller et al. [2010], we determine a value of m for each aspect ratio, by fitting equation (3)
to each Kr dataset. The fitting process was done on log10 (Kr)
to avoid biasing the fit to large values of Kr.
[11] The values of m determined from this fitting procedure
are listed in Table 1, and are plotted in Figure 3 against the
mean aspect ratio of the suspended particles for each dataset.
The data are approximately symmetricallydistributed around
rp = 1 and are welldescribed by a logGaussian function:
" 
m m1 exp 

log10 rp
2b2

2 #
;

where m1 is the maximum packing fraction for particles with


rp = 1 and b is a fitting parameter (b2 is equivalent to the
1
Auxiliary materials are available in the HTML. doi:10.1029/
2011GL047167.

2 of 5

L13316

L13316

MUELLER ET AL.: PARTICLE SHAPE AND SUSPENSION VISCOSITY

Figure 1. Examples of particles used in this study: (a) oblate polyacrylic glitter, (b) spherical glass beads, (c) prolate glass
fibres. Length of scale bar is 1 mm.

variance of the logGaussian function). For our data, the


best fit is achieved for m1 = 0.656 and b = 1.08. This value
of m1 is in excellent agreement with previously reported
values for sheared suspensions of spheres [Rutgers, 1962].
Note that alternative functional forms were tried, including
loglogistic and logCauchy functions, but these yielded
worse fits to data.
[12] Figure 3 also compares published m(rp) models
with our experimental results and with equation (4). Agreement between the data and previous models is generally very
poor, and is restricted to small intervals of rp in the prolate
range. Furthermore, most of the previous models produce
unphysical results (such as m > 1 or m < 0) for certain
aspect ratios. By contrast, the model that we propose is in
good agreement with the data across a wide range of oblate
and prolate aspect ratios, is continuous at rp = 1, and the
predicted value of m tends asymptotically towards zero for
extreme aspect ratios.

4. Practical Applications
[13] Equations (3) and (4) can be combined so that suspension consistency can be computed as a function of
particle aspect ratio and particle volume fraction; however,
it is the apparent viscosity that is more commonly sought
for practical applications. In the absence of yield stress,
the consistency K is identical to apparent viscosity when
n = 1 or _ = 1 s1 (equation (2)). In this case, the non
dimensionalization to form Kr is robust, indicating that our
results scale to arbitrary melt viscosity. If the model is
applied to situations in which both n 1 and _ 1 s1, the
analogy between Kr and hr is inexact and the true suspension viscosity can only be calculated if the value of flow
index n is known (which it usually is not). From equation (2),
we can see that the fractional difference between Kr and
hr is given by:
Error

Kr  r
_ 1n  1;
r

hence, for values of n and _ close to unity, this error is small.


Our experiments indicate that typically 0.9 < n 1 for /m <
0.7 so the flow index only deviates substantially from unity
at high particle volume fractions. Consequently, Kr can be
used as a proxy for hr at intermediate strain rates for all but
the most concentrated particle suspensions. For suspensions
that are only weakly shear thinning, consistency and apparent

viscosity are approximately equivalent for all situations of


practical interest.
[14] We now present a concrete example to illustrate the
importance of accounting for particleshapedependent
rheology when considering the fluid dynamics of crystal
bearing magma. Over the past few decades, numerical models
of conduit flow have been used extensively to investigate
the physical controls on eruption style [e.g., Sahagian, 2005].
Such models typically find that magma rheology exerts a
firstorder control. Following Llewellin and Manga [2005],
we use the opensource conduit flow model CONFLOW
[Mastin, 2002] as a testbench to explore the impact of using
a rheological model that accounts for crystal shape. We use
the standard Pinatubo white pumice parameters supplied
with CONFLOW, and the fragmentation criterion of Papale
[1999]. CONFLOW uses a version of the EinsteinRoscoe
viscosity model (hr = (1 /m)2.5) to account for crystal
volume fraction. We adapt the source code to use equation (3)
in its place, introducing the assumption that consistency
and viscosity are equivalent and limiting validity to /m ]
0.7 as described above (note that computed strain rate is in
the range 0.1 < _ 2 for all simulations, hence the error in
viscosity due to this assumption, given by equation (5), is
never greater than 20% and is usually much smaller).
Dependence of viscosity on aspect ratio is included through
equation (4) for m.
[15] Results, plotted in Figure 4, demonstrate the variations in predicted eruption parameters that arise as crystal
aspect ratio is varied from rp = 1 (equant) to rp = 10 (highly
prolate). The range of aspect ratios considered is represenTable 1. Particle Characteristics
Particles

Material

Mean rpa

s (rp)b

mc

Bt
Gl1
Gl2
spheres
F1
F2
F3
F4
F5
F6
F7
F8

biotite flakes
polyacrylic glitter
polyacrylic glitter
glass spheres
glass fibres
glass fibres
glass fibres
glass fibres
carbon fibres
glass fibres
glass fibres
glass fibres

0.04
0.14
0.16
1.00
2.50
3.50
5.75
5.90
9.05
10.6
12.6
22.0

0.018
0.014
0.016

1.12
1.46
2.77
3.03
6.53
5.74
3.99
7.36

0.220
0.540
0.550
0.633
0.573
0.558
0.538
0.538
0.430
0.404
0.359
0.323

3 of 5

Determined from optical microscopy.


s (rp) is the standard deviation of the particles aspect ratio.
Obtained from equation (3) (see section 3 for description of procedure).

b
c

L13316

MUELLER ET AL.: PARTICLE SHAPE AND SUSPENSION VISCOSITY

L13316

Figure 2. Relative consistency Kr versus particle volume fraction  for suspensions of (a) prolate and (b) oblate particles.
Kr () has been fitted using the MaronPierce model (equation (3)) for each dataset (grey lines) to determine m (values in
grey boxes). Error bars are shown unless they are smaller than the symbol.

tative of the range found in nature, considering both phenocryst and microlite phases: for phenocryst phases rp < 5 is
typical [e.g., Mock and Jerram, 2005]; microlites are commonly found with 5 < rp < 15 [Castro et al., 2003; Couch
et al., 2003].
[16] Figure 4 shows that the predicted fragmentation depth
(indicated by the inflection point in each curve) increases
from 1.25 km for equant crystals to 2.6 km for the most
prolate crystals. To achieve the same change without varying
aspect ratio, the volume fraction of equant crystals would
have to be increased from  = 0.3 to  = 0.47. Whilst we
recognize that magma at depth is unlikely to carry significant
volume fractions of very high aspect ratio crystals, it seems
very likely that large vertical variations in average aspect
ratio are common [Melnik and Sparks, 2005]. This modelling
example is not, therefore, intended to represent any particular

volcanic eruption, but to demonstrate the great sensitivity of


eruption models to crystal aspect ratio through its impact on
rheology. It is likely that strongly nonlinear eruption models
would show even greater sensitivity.

5. Conclusions
[17] We have performed laboratory analogue experiments
to determine the rheology of monodisperse suspensions of
particles for a range of particle aspect ratios from highly
oblate to highlyprolate. We have found that the impact of
particle shape on rheology can be captured through its effect
on the maximum packing fraction of particles m. We find
that the maximum fraction of particles that a suspension can
hold before jamming is highest for equant particles, and
decreases systematically for particles that are increasingly

Figure 3. The maximum packing fraction of particles m as a function of the particle aspect ratio rp. The data have been
fitted with a logGaussian function (equation (4); R2 = 0.905). Error bars in rp correspond to 1 s of the ladistribution of the
particles, assuming constant lb. The errors in m in Figure 2 are determined by fitting equation (3) to the data in Figure 2, but
with Kr error in (Kr) substituted for Kr.
4 of 5

L13316

MUELLER ET AL.: PARTICLE SHAPE AND SUSPENSION VISCOSITY

Figure 4. Gas volume fraction, pressure and vertical velocity against depth in a volcanic conduit, during eruption, calculated using CONFLOW [Mastin, 2002] for the standard
Pinatubo white pumice parameters. In each case  = 0.3
and rp takes the values 1, 2, 5 and 10.
prolate or oblate; consequently, a suspension of equant
particles will have a lower viscosity than a suspension with
the same volume fraction of prolate or oblate particles.
[18] We have developed a method for calculating the
viscosity of a particle suspension if particle volume fraction
 and aspect ratio rp are known: first, maximum packing
fraction m is calculated using equation (4); second, relative
consistency Kr is calculated using equation (3). Relative
consistency is approximately equivalent to relative apparent
viscosity hr for a suspension which shows only mild shear
thinning; practically, this puts the following, conservative,
constraints on the validity of our approach: 0.04 rp 22;
/m ] 0.7.
[19] Finally, we have used the practical example of the
numerical modelling of a volcanic eruption to demonstrate
the importance of accounting for crystal shape; in our
example, changing crystal aspect ratio from rp = 1 to rp = 10
has a similar impact to changing particle volume fraction
from  = 0.3 to  = 0.47. It is crucial, therefore, that
modellers of volcanic and magmatic flows ask not only what
volume fraction of crystals is in the flow, but also what
shape the crystals have.
[ 20 ] Acknowledgments. S.M. is supported by NERC Research
Fellowship NE/G014426/1. We thank M. Mangan and Y. Lavallee for their
reviews, which helped to improve the manuscript, and L. Mastin for
supplying the source code for CONFLOW and facilitating our modifications.
Dataset F5 has been collected by C. Cimarelli.
[21] The Editor thanks Yan Lavalle and Margaret Mangan for their
assistance in evaluating this paper.

References
Bernal, J. D., and J. Mason (1960), Coordination of randomly packed
spheres, Nature, 188, 910911.
Brenner, H. (1974), Rheology of a dilute suspension of axisymmetric
Brownian particles, Int. J. Multiphase Flow, 1, 195341.
Caricchi, L., L. Burlini, P. Ulmer, T. Gerya, M. Vasalli, and P. Papale
(2007), NonNewtonian rheology of crystalbearing magmas and implications for magma ascent dynamics, Earth Planet. Sci. Lett., 264, 402419.

L13316

Castro, J. M., K. V. Cashman, and M. Manga (2003), A technique for measuring 3D crystalsize distributions of prismatic microlites in obsidian,
Am. Mineral., 88, 12301240.
Couch, S., R. S. J. Sparks, and M. R. Carroll (2003), The kinetics of degassing
induced crystallization at Soufriere Hills volcano, Montserrat, J. Pet., 44,
14771502.
Donev, A., I. Cisse, D. Sachs, E. A. Variano, F. H. Stillinger, R. Conelly,
S. Torquato, and P. M. Chaikin (2004), Improving the density of jammed
disordered packings using ellipsoids, Science, 303, 990993.
Einstein, A. (1906), Eine neue Bestimmung der Molekldimensionen, Ann.
Phys., 19, 289306.
Einstein, A. (1911), Berichtigung zu meiner Arbeit Eine neue Bestimmung
der Molekldimensionen, Ann. Phys., 34, 591592.
Evans, K. E., and A. G. Gibson (1986), Prediction of the maximum packing
fraction achievable in randomly oriented shortfibre systems, Compos.
Sci. Technol., 25, 149162.
Gan, M., N. Gopinathan, X. Jia, and R. A. Williams (2004), Predicting
packing characteristics of particles of arbitrary shapes, Kona, 22, 8293.
Herschel, W. H., and R. Bulkley (1926), Konsistenzmessungen von Gummi
Benzollsungen, Kolloid Z., 39, 291300.
Ishibashi, H., and H. Sato (2007), Viscosity measurements of subliquidus
magmas: Alkali olivine basalt from the HigashiMatsuura district, southwest Japan, J. Volcanol. Geotherm. Res., 160, 223238.
Jeffery, G. B. (1922), The motion of ellipsoidal particles immersed in a
viscous fluid, Proc. R. Soc. London, Ser. A, 102, 161179.
Kitano, T., T. Kataoka, and T. Shirota (1981), An empiricalequation of the
relative viscosity of polymer melts filled with various inorganic fillers,
Rheol. Acta, 20, 207209.
Lavallee, Y., K.U. Hess, B. Cordonnier, and D. B. Dingwell (2007), Non
Newtonian rheological law for highly crystalline dome lavas, Geology,
35, 843846.
Lejeune, A.M., and P. Richet (1995), Rheology of crystalbearing silicate
melts: An experimental study at high viscosities, J. Geophys. Res., 100,
42154229.
Llewellin, E. W., and M. Manga (2005), Bubble suspension rheology and
implications for conduit flow, J. Volcanol. Geotherm. Res., 143, 205217.
Maron, S. H. and P. E. Pierce (1956), Application of ReeEyring generalized flow theory to suspensions of spherical particles, J. Colloid Sci., 11,
8095.
Mastin, L. G. (2002), Insights into volcanic conduit flow from an open
source numerical model, Geochem. Geophys. Geosyst., 3(7), 1037,
doi:10.1029/2001GC000192.
Melnik, O., and R. S. J. Sparks (2005), Controls on conduit magma flow
dynamics during lava dome building eruptions, J. Geophys. Res., 110,
B02209, doi:10.1029/2004JB003183.
Milewski, J. V. (1973), A study of the packing of milled fibreglass and
glass beads, Composites, 4, 258265.
Mock, A., and D. A. Jerram (2005), Crystal size distributions (CSD) in
three dimensions: Insights from the 3D reconstruction of a highly porphyritic rhyolite, J. Pet., 46, 15251541.
Mueller, S., E. W. Llewellin, and H. M. Mader (2010), The rheology of
suspensions of solid particles, Proc. R. Soc. A, 466, 12011228.
Pabst, W., E. Gregorova, and C. Bertold (2006), Particle shape and suspension rheology of shortfiber systems, J. Eur. Ceram. Soc., 26, 149160.
Papale, P. (1999), Straininduced magma fragmentation in explosive eruptions, Nature, 397, 425428.
Parkhouse, J. G., and A. Kelly (1995), The random packing of fibres in
three dimensions, Proc. R. Soc. A, 451, 737746.
Powell, R. L. (1991), Rheology of suspensions of rodlike particles, J. Stat.
Phys., 62, 10731094.
Rahli, O., L. Tadrist, and R. Blanc (1999), Experimental analysis of the
porosity of randomly packed rigid fibres, C. R. Acad. Sci., 327, 725729.
Roscoe, R. (1952), The viscosity of suspensions of rigid spheres, Br.
J. Appl. Phys., 3, 267269.
Rutgers, I. R. (1962), Relative viscosity of suspensions of rigid spheres in
Newtonian liquids, Rheol. Acta, 2, 202210.
Sahagian, D. (2005), Volcanic eruption mechanismsInsights from intercomparison of models of conduit processes, J. Volcanol. Geotherm.
Res., 143, 1236.
Williams, S. R., and A. P. Philipse (2003), Random packings of spheres
and spherocylinders simulated by mechanical contraction, Phys. Rev. E,
67, 051301.
E. W. Llewellin, Department of Earth Sciences, Durham University,
South Road, Durham DH1 3LE, UK.
H. M. Mader and S. Mueller, School of Earth Sciences, University of
Bristol, Wills Memorial Building, Queens Road, Bristol BS8 1RJ, UK.
(s.mueller@bristol.ac.uk)

5 of 5

Você também pode gostar