Sei sulla pagina 1di 13

3 Research Letter

Geochemistry
G
Volume 10, Number 3
Geophysics 18 March 2009
Q03010, doi:10.1029/2008GC002138
Geosystems
AN ELECTRONIC JOURNAL OF THE EARTH SCIENCES ISSN: 1525-2027
Published by AGU and the Geochemical Society

Click
Here
for
Full
Article

A model for the rheology of particle-bearing suspensions


and partially molten rocks
Antonio Costa
Istituto Nazionale di Geofisica e Vulcanologia, I-80124 Naples, Italy
Centre for Environmental and Geophysical Flows, Department of Earth Sciences, University of Bristol, Bristol BS8 1RJ,
UK (costa@ov.ingv.it)

Luca Caricchi
Department of Earth Sciences, ETH Zurich, CH-8092 Zurich, Switzerland

Now at Institut des Sciences de la Terre, Université d’Orleans, F-45071 Orleans CEDEX, France

Nickolai Bagdassarov
Institut für Geowissenschaften, Facheinheit Geophysik, Universität Frankfurt, D-60438 Frankfurt, Germany

[1] This contribution presents a semiempirical model describing the effective relative viscosity of crystal-
bearing magmas as function of crystal fraction and strain rate. The model was applied to an extensive data
set of magmatic suspensions and partially molten rocks providing a range of values for the fitting
parameters that control the behavior of the relative viscosity curves as a function of the crystal fraction in
an intermediate range of crystallinity (30–80 vol % crystals). The analysis of the results and of the
materials used in the experiments allows for evaluating the physical meaning of the parameters of the
proposed model. We show that the model, by varying the parameters within the ranges obtained during
the fitting procedure, is able to describe satisfactory the effective relative viscosity as a function of crystal
fraction and strain rate for suspensions having different geometrical characteristics of the suspended solid
fraction.

Components: 6984 words, 4 figures, 1 table.


Keywords: melts; concentrated suspensions; viscosity; strain rate.
Index Terms: 8160 Tectonophysics: Rheology: general (1236, 8032); 8429 Volcanology: Lava rheology and morphology;
8439 Volcanology: Physics and chemistry of magma bodies.
Received 26 June 2008; Revised 16 January 2009; Accepted 28 January 2009; Published 18 March 2009.

Costa, A., L. Caricchi, and N. Bagdassarov (2009), A model for the rheology of particle-bearing suspensions and partially
molten rocks, Geochem. Geophys. Geosyst., 10, Q03010, doi:10.1029/2008GC002138.

1. Introduction many application fields, encompassing extremes


such as molding processes in industry or magma
[2] A quantitative description of the rheology of flow in geosciences. The knowledge of the rheology
solid-liquid suspensions is of a great interest in of magmas with large fraction of crystals is important

Copyright 2009 by the American Geophysical Union 1 of 13


Geochemistry 3
Geophysics
Geosystems G costa et al.: rheology of particle-bearing suspensions 10.1029/2008GC002138

to define the dynamics of magma rise inside volcanic (SLT), occurring at melt fraction around 0.5
conduits, the extrusion of volcanic domes and more corresponding to the previously described RCMP.
generally the dynamics of volcanic eruptions.
[4] We present a semiempirical model able to
[3] Several studies on the rheological behavior of describe the rheological effects of crystals from
magmatic suspensions and partially molten sys- low to medium-high solid fractions, reproducing
tems were published in the last 30 years. The the transition from the regime where the strength is
pioneering works of Goetze [1977] and Arzi controlled by melt viscosity up to the beginning of
[1978] dominated the view on the transition from the regime where strength is controlled by the solid
a solid-state to a melt-dominated creep rheology, framework, avoiding the singularity predicted by
setting the threshold at a melt fraction about 0.25– classical models at the maximum packing fraction
0.30 (rheological critical melt percentage, RCMP). [e.g., Roscoe, 1952; Krieger and Dougherty, 1959]
Successively, a number of experimental studies Although we are aware that the relationship be-
focused on the rheological behavior of melt- tween stresses and strain rates for highly crystal-
bearing materials. The rheology of partially molten lized melts and partially molten rocks is very
rocks, with melt fractions lower than 0.12, was complex [see, e.g., Renner et al., 2000; Bercovici
explored thoroughly by Kohlstedt and Zimmerman et al., 2001; Rosenberg and Handy, 2005] and can
[1996]. The rheology of magma with solid frac- be nonunique, here our aim is to describe only an
tions lower than 0.35–0.40 was investigated by effective shear rheology of homogeneous systems
Pinkerton and Stevenson [1992]. These observa- of melts with crystals from low to medium-high
tions highlighted the necessity to address the phys- solid fractions by means of a mathematical model,
ical meaning of the critical melt fraction (transition the adjustable parameters of which can be inter-
from melt-dominated to solid-state creep rheolo- preted in terms of physical conditions of deforma-
gy). Creep data on partially molten granites col- tion experiments.
lected by Rutter and Neumann [1995] raised the
question about the rapidity of the transition [5] The parameterization includes the strain rate–
approaching the RCMP in a two phase system. dependent rheology of partially crystallized materi-
Vigneresse et al. [1996] argued that the RCMP als. The analysis of the fitting results, taking into
during crystallization differs from the RCMP dur- account the characteristics of the materials employed
ing melting. The review of Ji and Xia [2002] in the experiments and the application of some
explored a few different models of two-phase theoretical concepts on the rheology of two-phase
mixture rheology concluding that the relative vis- mixtures, allow for assigning a physical meaning to
cosity of crystal supported systems and melt sup- the parameters of the proposed equations. Finally the
ported materials can be described using the Reuss effect of particle shapes and crystal size distribution
[1929] and Voigt [1928] bounds. The exponential on the variation of viscosity as function of crystal
decrease of relative viscosity at the critical melt fraction and strain rate is also discussed.
fraction was interpreted as a transition from iso-
strain rate (solid-phase dominated; Voigt bound) to 2. Parameterization for Solid Fraction
isostress (melt-phase dominated; Reuss bound) Dependence
deformation conditions [e.g., Takeda and Obata,
2003]. Petford [2003] reviewed the existing litera- [6] A quantitative description of the variations of
ture and experimental data on rheology of mag- viscosity as function of the solid fraction present
matic suspensions and partially molten rocks. He several problems. The effective relative viscosity
pointed out the crucial role of the relationship (ratio between the effective viscosity of the solid-
between tortuosity and porosity for various pack- liquid suspension and the viscosity of the liquid
ing densities in determining the position of the phase) needs to be introduced in many cases [see
RCMP. Rosenberg and Handy [2005] identified Costa, 2005]. Since viscosity controls magma
two major rheological transitions occurring in transport, modeling of volcanic processes such as
partially molten rocks: the melt connectivity tran- dome growth requires to estimate the viscosity
sition (MCT) occurring at melt fraction around dependence on crystal content, even in the limit
0.07, which induces a major decrease of the of high solid fractions where few experimental
strength of the fully crystallized rock, and the observations are available [see, e.g., Melnik and
second transition, defined as solid-liquid transition Sparks, 1999, 2005; Costa et al., 2007a, 2007b].

2 of 13
Geochemistry 3
Geophysics
Geosystems G costa et al.: rheology of particle-bearing suspensions 10.1029/2008GC002138

[7] A parameterization describing the relative vis- and when f ! 0, it recovers exactly to the Einstein
cosities of a melt-solid mixture, valid for relatively equation:
large solid fractions, was proposed by Costa
[2005]. At low solid content the parameterization hðfÞ ¼ ð1 þ B  fÞ: ð5Þ
is approximately reduced to the well-established
formulation of Krieger and Dougherty [1959]. 3. Model Applicability, Strain Rate
However, for large particle volume fractions (8),
the parameterization suggested by Costa [2005] Effects, Influence of Particle Shape,
tends quickly to a constant value, because the and Particle Size Distribution
nonlinear term in the error function (‘‘erf’’) rapidly
saturates as 8 approaches unity. The Costa [2005] [8] The rheology of a solid-fluid mixture is very
relationship for relative viscosity of concentrated complex and several simplifying assumptions need
suspensions was later improved and generalized by to be introduced. The full description of an isotro-
Costa et al. [2007a] (see also http://arxiv.org/pdf/ pic linear medium requires the use of two viscos-
physics/0512173). The parameterization used by ities, a bulk viscosity and a shear viscosity.
Costa et al. [2007a, 2007b] is adopted also in the Moreover, localized failure of partially molten rock
current study. The relationship between the relative does not permit a unique mapping between the
viscosity and the particle volume fraction is de- space of stress and strain rate tensor [e.g., Rutter
scribed by: and Neumann, 1995].
[9] In order to simplify the problem we consider
1 þ 8d
hðfÞ ¼ ; ð1Þ only homogeneous systems of melts with crystals
½1  F ð8; e; g ÞBf * from low to medium-high solid fractions. Results
obtained above the percolation threshold of the
where solid matrix (the maximum solid fraction to obtain
 pffiffiffi  liquid interconnection) are considered merely as
p f
F ¼ ð1  xÞ  erf 8  ð1 þ 8g Þ with 8 ¼ : mathematical extrapolations. Moreover we de-
2  ð1  x Þ f*
scribe the rheology of the homogenous systems
ð2Þ simply by means of an effective shear viscosity,
defined as the ratio between the maximum differ-
Here f* represents the critical solid fraction at the ential stress and the strain rate.
onset of the exponential increase of h; x (
1), g
and d are empirical parameters and B is the [10] Besides the increase in viscosity with crystal
Einstein coefficient (i.e., the intrinsic viscosity) fraction, the rheology of highly concentrated solid-
with a nominal value of 2.5. The parameter g is a fluid mixture strongly depends on strain rate, crystal
measure of the rapidity of relative viscosity shapes and crystal size distribution. Concerning
increase with crystal fraction, as f approaches f*. strain rate, Caricchi et al. [2007] experimentally
d controls the increase of h at f > f*. As showed that at relatively low strain rates (106 –
demonstrated below, d can be expressed in terms 105 s1) particle-silicate melt suspensions behave
of g as d = A  g where A = 13 is an empirical as a Newtonian liquid. Increasing strain rate (_e), and
constant. For large d, the second term in the consequently the applied stress, beyond these val-
numerator represents a negligible correction when ues induces a transition to non-Newtonian behavior
f < f*, while it becomes important when f > f*. (shear thinning) caused by a decrease of the degree of
At f = f*h depends on x and f*, only: randomness of the particle distribution. At high
enough strain rates ( 103 s1), a maximum degree
    pffiffiffi Bf of ordering is reached and the effective viscosity is no
p *
h ¼ h f ¼ 2  1  ð1  xÞ  erf : ð3Þ longer strain rate–dependent. This high strain rate
* * 1x
rheological behavior was observed in several parti-
Expanding F in equation (2) as a Taylor series [see cle-bearing suspensions and was defined as pseudo-
Costa, 2005], it can be noticed that with decreasing Binghamian behavior [e.g., van der Werff and de
f, at f < f*, equation (1) tends exactly to the Kruif, 1989; Barnes, 1999]. However, for the case of
Krieger and Dougherty [1959] relationship: magmas, at large strain rates (_e > 103 s1), viscous
heating can lead to an additional decrease of the melt
 Bf viscosity [Hess et al., 2008]. The proposed formula-
h ¼ 1  f=f* *; ð4Þ
tion considers the existence of a pseudo-Binghamian
rheological field for e_ > 103 s1 and, as a conse-
3 of 13
Geochemistry 3
Geophysics
Geosystems G costa et al.: rheology of particle-bearing suspensions 10.1029/2008GC002138

quence, the presented model provides only an upper limited range, a linear dependence of f* on the
limit of effective relative viscosity at higher strain average aspect ratio [e.g., Pabst et al., 2006] but
rates. The rheological transitions occurring with in- further investigations are necessary. Experimental
creasing strain rate from Newtonian to non-Newtonian data for polymer melts with different volume frac-
and finally pseudo-Binghamian behavior, influence tion of fibers show a linear dependence of f* on the
the parameters of equation (2). The parameters f*, x, g average aspect ratio [e.g., Pabst et al., 2006]:
and d are nearly constant up to the transition to non- 
Newtonian behavior, they become, again, almost f ¼ fS  aR R  1 ;
*
ð8Þ
constant at a strain rate of about 103 s1 where the
mixture behaves as a pseudo-Bingham material. where fs is the critical fraction for spherical
Parameters in equation (2) were empirically described particle, R is the average particle aspect ratio
by Caricchi et al. [2007] as a function of e_ by: defined as ratio between the long and the short

axes, and aR is an empirical constant. The intrinsic
f* ¼ ft þ Df tanh bf Log10 ðe_ Þ þ cf ; viscosity B, here assumed constant for sake of


x ¼ xt þ Dx tanh bx Log10 ðe_ Þ þ cx ; simplicity, can also change with the aspect ratio


g ¼ g t þ Dg tanh bg Log10 ðe_ Þ þ cg ; [e.g., Pabst et al., 2006]. However, the lack of
d ¼ dt  Dd tanh½ bd Log10 ðe_ Þ þ cd  ð6Þ experimental data at high solid fraction does not
permit to constrain the values of the empirical
(for the numerical values of the empirical constants parameters for which further experimental data and
see equations (6)–(9) in the work by Caricchi et al. investigations are needed [e.g., Mueller et al., 2008].
[2007]; in the work by Caricchi et al. [2007], a = [12] Finally, experiments for bidispersed systems
1  x). Results from Caricchi et al. [2007] show show that the smaller the particle diameter ratio the
that f*, x and g increase with increasing strain rate higher the critical solid fraction [e.g., Chong et al.,
while d decreases with increasing strain rate. Their 1971]. The dependence of f* on the crystal size
experimental data also show that cd/bd cg/bg distribution can be given as [Chong et al., 1971]:
cx/bx cf/bf  1, indicating the occurrence of two
characteristic strain rates e_ c and e_ d 10  e_ c that  0:1041
d
should be functions of the material properties. f ¼ f md  ; ð9Þ
* * D
Moreover, their data suggest an approximately
constant exponent for the dependence on e_ in where f*md denotes the critical fraction for a
equations (6). After some analytical manipulations, monodispersed system, and d and D are the diameters
equations (6) can be rewritten as: of the smaller and bigger particles respectively.
ðe_ =_ec Þn ðe_ c =_eÞn
f* ¼ fm þ Df
ðe_ =_ec Þn þðe_ c =_eÞn
4. Data Set for the Calibration
ðe_ =_ed Þn ðe_ d =_eÞn of the Model
x ¼ x m þ Dx
ðe_ =_ed Þn þðe_ d =_eÞn
ðe_ =_ed Þn ðe_ d =_eÞn [13] Bagdassarov and Dorfman [1998] reported a
g ¼ g m þ Dg noteworthy data set of effective relative viscosities
ðe_ =_ed Þn þðe_ d =_eÞn
of partially molten granite. Their study includes
ðe_ =_ed Þn ðe_ d =_eÞn
d ¼ d m  Dd ð7Þ data from Rutter and Neumann [1995], Auer et al.
ðe_ =_ed Þn þðe_ d =_eÞn
[1981], Rushmer [1995], Arzi [1978], and van der
where fm = (f1 + f0)/2, Df = (f1  f0)/2, x m = Molen and Paterson [1979]. Effective relative
(x 1 + x 0)/2, Dx = (x 1  x 0)/2, g m = (g 1 + g 0)/2, viscosity was thereby considered as the ratio be-
Dg = (g 1  g 0)/2, and d m = (d1 + d0)/2, Dd = tween effective viscosity of the partially molten
(d 0  d 1)/2 with subscript 1 indicating the values rock and the viscosity of the melt phase.
at very large e_ and subscript 0 those at very low e_ . [14] In order to calibrate the parameterization de-
The best fit of the data of Caricchi et al. [2007] scribed by equation (2) we used all data mentioned
results n ffi 0.33, Log10 (_ec) = 4.30, Log10 (_ed) = above for large f including data from Lejeune and
3.37, and the other parameter values as reported Richet [1995], which measured relative viscosity of
in the caption of Figure 1. melt-particle suspensions up to solid fractions of
[11] Experimental data for polymer melts with 0.65 (Mg3Al2Si3O12), and results from Caricchi et
different volume fraction of fibers suggest, in a al. [2007] for a water-bearing haplogranititc melt
with suspended quartz particle fraction varying

4 of 13
Geophysics
Geosystems
Geochemistry

G
3 costa et al.: rheology of particle-bearing suspensions

Figure 1. Strain rate dependence of parameters in equation (8) with n ffi 0.33, e_ c = 104.3 s1, and e_ d = 103.37 s1. (a) f*
versus e_ with fm = 0.591 and Df* = 0.069, (b) x versus e_ with x m = 4.63  104, (c) g versus e_ with g m = 5.76 and Dg =
4.46, and (d) d versus e_ with dm = 6.78 and Dd = 6.66. The values reported in Figure 1 are from Caricchi et al. [2007].
10.1029/2008GC002138

5 of 13
Geochemistry 3
Geophysics
Geosystems G costa et al.: rheology of particle-bearing suspensions 10.1029/2008GC002138

Figure 2. Variation of the logarithm of the effective relative viscosities with respect to the solid fraction for the data
set reported in the inset. Dashed lines represent the bounding limits predicted by Caricchi et al. [2007]. d = A  g
with A = 13. The gray shaded region denotes the region above the percolation threshold of the solid matrix (the
maximum solid fraction to obtain liquid interconnection) where results need to be considered merely as mathematical
extrapolations.

between 0.5 and 0.8. We also included data to encompass the variation of the effective viscosity
obtained from experiments with partially molten over the full range of solid fractions.
samples of olivine with MORB obtained by Scott
and Kohlstedt [2006] and Hirth and Kohlstedt [1995] [15] The available information on applied strain rate
in the range of the strain rate from 106 to 104 s1 and geometrical characteristics of the suspended
for diffusion creep regime. Additional data for a particles were extracted from each set of data
system containing particles with aspect ratios of considered for the fitting procedure. However, there
1:2.5 (Li2Si2O5 [Lejeune and Richet, 1995]) were are some limitations for the data reported by Arzi
also included. These data allow for investigating [1978] and Rutter and Neumann [1995] that is worth
some particle shape effects on the fitting parameters. stressing. Arzi [1978] reported values of relative
Data from Champallier et al. [2008] and Caricchi et viscosity of 108, 108 and 106 for samples having
al. [2008] were also considered in the data set for melt fractions of 0.06, 0.12, and 0.17 respectively.
comparison. Champallier et al. [2008] conducted The experiments on the different specimens were
experiments on a system similar that investigated performed at different temperatures and the relative
by Caricchi et al. [2007] in a strain rate range viscosity was calculated using melt phase viscosity
between 6  104 and 2  103 s1. Caricchi et al. values reported by Shaw [1963]. Unfortunately, the
[2008] investigated a natural material containing temperature effect on the viscosity of the silicate
elongated plagioclase particles. The data at medium melt was not carefully considered during the calcu-
to high crystal fraction were combined with the lation of the relative viscosity. Using the Hess and
values predicted by the Krieger-Dougherty relation- Dingwell [1996] model, which is more general
ship for f < 0.1 and data of Thomas [1965] for f < 0.2 and also able to reproduce Shaw’s [1963] data, to

6 of 13
Geochemistry 3
Geophysics
Geosystems G costa et al.: rheology of particle-bearing suspensions 10.1029/2008GC002138

Table 1. Parameters Obtained by the Best Fit of Experimental Data


Parametersa

1a 1b 2 3 4 5 6 7

f* 0.384 0.489 0.451 0.652 0.706 0.736 0.622 0.592 (0.52/0.66)


g 7.701 1.440 6.220 1.911 5.142 3.679 3.429 7.462 (13.76/1.16)
x  104 2.0 0.3 0.5 1.0 1.0 0.5 0.1 4.75 (0.32/9.18)
Log10(_e) s1 8/6 8/6 7/4 5 n.s. n.s. 6/4 6/3
a
Parameters 1a, Lejeune and Richet [1995] (for Li2Si2O5); parameters 1b, Lejeune and Richet [1995] (for Mg3Al2Si3O12); parameters 2, Rutter
and Neumann [1995]; parameters 3, van der Molen and Paterson [1979]; parameters 4, Rushmer [1995]; parameters 5, Arzi [1978]; parameters 6,
Scott and Kohlstedt [2006] and Hirth and Kohlstedt [1995]; parameters 7, Caricchi et al. [2007]. For all calculations, d = A  g with A = 13. n.s.
means ‘‘not specified’’ by the authors.

calculate the appropriate values of the melt viscos- controlled by the viscosity of the liquid phase to a
ity at the experimental conditions investigated by system where particle-particle interactions induce a
Arzi [1978], the values of relative viscosity result strong increase of viscosity [Caricchi et al., 2007].
109, 108.6 and 106.9 for the sample with melt The considered data set, summarized in Table 1,
fraction 0.06, 0.12 and 0.17 respectively. Addition- shows values of f* over a wide range of crystal
ally, the applied stresses or strain rates are not fractions ranging from 0.38 to 0.74 with an average
reported for each experiment creating difficulties in value around 0.57. The dependence of f* on the
evaluating the strain rate effect on the effective shear stress [Caricchi et al., 2007; Wildemuth and
viscosity. The data of Rutter and Neumann [1995] Williams, 1984] can partly explain this variation.
present some uncertainties regarding the amount of As mentioned above, particle dispersion and parti-
melt present during the experiments. They used cle aspect ratio [Chong et al., 1971; Solomon and
natural samples that are not in equilibrium at Boger, 1998; Yue et al., 1999; Pabst et al., 2006]
experimental conditions and, therefore, tend to can additionally influencing f* (see equations (8)
react producing either melt or crystallizing mineral and (9)). More specifically elongated particles tend
phases. Additionally, the authors report the diffi- to decrease f* as confirmed by the data collected
culty to estimate the viscosity of the sample with by Lejeune and Richet [1995] and Caricchi et al.
the lowest f (0.53–0.67) because the applied stress [2008] (Figure 2).
at flow was within the experimental error. Conse-
quently, the reported value of viscosity for f = 0.53 6. Discussion
is only an upper limit. As mentioned above, an
additional aspect of the experiments performed by [18] The proposed model provides good fitting
Rutter and Neumann [1995] that needs to be results of the experimental data (Figure 2) but for
considered is the tendency, reported by the authors, identifying the physical significance of the adjust-
of the samples to deform inhomogeneously. Brittle able parameters it is worth highlighting their effects
behavior and strain partitioning between the solid and interpreting the results on the basis of the
and liquid component of the deformed material is theory of two-phase suspensions (see Appendix A).
not accounted for by the model presented in this
study. [19] The bulk viscosity of a mixture of two phases
is bounded by two curves defined as isostrain rate
5. Fitting Results bound (called upper or Voigt bound) and the iso-
stress bound (called lower or Reuss bound). A
[16] In Figure 2 and Table 1 we present the results more accurate description of the variation of the
obtained by best fitting equation (2) with the data effective viscosity against the crystal fraction can
set described above. The results show that most of be obtained from Hashin-Shtrikman bounds. Upper
the available experimental data are within the range and lower bounds for the relative viscosity can be
predicted by Caricchi et al. [2007] for different calculated as (see Appendix A):
applied stress (or strain rate) conditions. m2 1f
[17] One of the most important parameters deter- hþ
HS ¼ 
m1 1 2 m
þ  Cþ  f  1
mining the rheological behavior of a dense solid- m2 =m1  1 5 m2
ð10Þ
liquid mixture is the critical solid fraction f*. This h ¼1þ
f
HS 1 2
critical fraction delimits the transition from a þ  C   ð1  fÞ
system where the viscosity of the suspension is m2 =m1  1 5

7 of 13
Geochemistry 3
Geophysics
Geosystems G costa et al.: rheology of particle-bearing suspensions 10.1029/2008GC002138

where the symbol h denotes the ratio of the bulk ented particles are characterized by higher tortuosity
viscosity to the soft phase viscosity, f is the volume of the melt flow patterns and consequently by
fraction of the more viscous phase, m1 denotes the relatively high values of g. This observation is
soft phase viscosity, m2 represents the viscosity of the confirmed by the Caricchi et al. [2007] data in
more viscous phase m+HS and m HS denote the upper which the ordering of the particles, caused by higher
and the lower bound for shear viscosity of the two applied strain rates, induces a strong decrease of g
phase system, and C± is a constant (for a melt and (from 14 at 106 s1 to 1 at 103 s1).
solid mixture C+ = 1 and C = 9/7).
[23] As previously reported, experimental data
[20] The viscosity for the case of solid-phase show an empirical relationship between d and g
supported flow is computed as the upper Hashin- (d = A  g with A = 13), allowing for reducing the
Shtrikman boundary (or isostrain model); the flow, number of free parameters. Moreover, this implies
in this case, is controlled by the strain rate of the that the steeper the transition around the critical
solid matrix. Conversely at high melt fraction the fraction f* (controlled by g) the smoother the
pressure or the shear stress is equal for both solid increase of viscosity at larger solid fractions.
and melt phase and the flow is controlled by the
soft phase. In this condition the viscosity corre- [24] Concerning x, equation (3) shows that, for a
sponds to the lower Hashin-Shtrikman bound (or given f*, it determines the value of the characteristic
isostress case). Therefore, we know the two limits viscosity h* = h(f*). Like d, also x plays an impor-
to which relative viscosity tends as f ! 0 and f ! tant role in determining the relative viscosity at very
1 respectively. The variations of relative viscosity high solid fraction. In fact, from equations (1) and
in an intermediate range of solid fractions are (2) we can show that in the limit f ! 1 we have:
related to the geometrical characteristics and size !
distribution of the suspended particles, and to the 1 fd 1  
h Bf
1þ d Bf
1 þ fd
* : ð11Þ
applied strain rate conditions. More precisely the x * f* x *
applied strain rates influence the degree of ran-
domness of the suspended particles, which, in turn, [25] These indications provide some constraints on
results in different trends of relative viscosity the physical significance of the variations of the
variation as function of the particle fraction. adjustable parameters. The additional effect due to
strain rate are illustrated in Figure 2 and mathe-
[21] Figure 3 summarizes the effects of the param- matically formulated in equation (8).
eters used in equations (1) and (2) on the variations
of viscosity with respect to crystal fraction. The [26] The relationship between effective yield stress
critical solid fraction f* defines the limit of the and melt fraction (see Appendix A) is valid up to
regime where the viscosity is mainly controlled by the percolation limit of the solid matrix, fC, i.e.,
the melt phase from the regime where interactions the maximum solid fraction to obtain liquid inter-
among particles have a pivotal role. The lowest connection [Zhou et al., 1995]. For higher f,
values of f* are characteristic for suspensions particles start to deform substantially under the
containing elongated particles (Figure 2). Increase application of stress thereby inducing a rheological
of particle sphericity, strain rates and poly disper- change from two-phase suspension to solid rock
sion (wider distribution of particle sizes) results in deformation.
an increase of f*, which produces trends similar to
[27] Additional experimental data exploring the
those shown in Figure 3c.
effects of crystal size distributions, crystal shapes,
[22] As mentioned above, d controls the maximum and viscous heating at high strain rates are required
value of effective relative viscosity as f tends to 1 to define accurately the range of variation of the
(Figure 3a). Consequently, large values of d are fitting parameters as a function of the geometrical
expected for suspensions containing particles with properties of the suspended particles.
high resistance to deformation (i.e., high viscosity).
The parameter g controls the sharpness of the 7. Conclusions
viscosity increase around the critical particle frac-
tion (Figure 3b). The increase of the viscosity at the [28] The effects of crystal fraction and strain rate
critical solid fraction is related to a strong increase on the rheology of partially crystallized magmas
of the tortuosity of the melt flow patterns [Gibilaro, was presented and discussed. The proposed param-
2001] due to the achievement of a continuous crystal eterization allows for describing variations of 6–9
framework. Suspensions containing randomly ori- orders of magnitude of the relative viscosity due

8 of 13
Geophysics
Geosystems
Geochemistry

G
3 costa et al.: rheology of particle-bearing suspensions

Figure 3. Effect of the variation of single parameters on the effective relative viscosities with respect to the solid fraction
for the specific values reported in the inset.
10.1029/2008GC002138

9 of 13
Geochemistry 3
Geophysics
Geosystems G costa et al.: rheology of particle-bearing suspensions 10.1029/2008GC002138

to the presence of crystals and up to 3 orders of the shear modulus of a two component mixture
magnitude due to shear thinning for a large exper- [Mavko et al., 1998]:
imental data set. The uncertainties in the relative
proportions of melt and crystals and the lack of f
mHS ¼ m1 þ ; ðA1Þ
information on important experimental conditions 1 2 ð1  fÞ 
ðm2  m1 Þ þ  C
(e.g., applied strain rate), result in 1–2 orders of 5 m1
magnitude uncertainty for effective relative viscos-
ity at high crystal fractions. The influence of the where m is the shear viscosity of the two phases,
parameters on the relative viscosity curve with f is the volume fraction of the more viscous phase,
respect to the crystal fraction is presented and their C± is a constant. Upper and lower boundaries are
variations associated with changes of the geome- interchanged swapping the indexes of phase 1 and
tries of the suspended particles and different ap- 2. For the mixture of melt and solid phase C+ = 1
plied stress (or strain rate). and C = 9/7, respectively. In the elastic case, the
constant C± is the ratio (K + 2m)/(K + 4/3m), where
Appendix A K and m are the bulk and shear modulus. In the
viscous case instead of K one needs to insert a
[29] In order to identify and interpret the physical volume viscosity, which, for the solid phase, is
significance of the adjustable parameters of the infinity and for liquid phase is about the value of
semiempirical model presented in this paper and the shear viscosity. The sample flow supported by
some limitations of the model itself, here we briefly solid phase is the case of the upper Hashin-
describe some concepts of the theory of two-phase Shtrickman boundary or isostrain model. The total
suspensions. sample flow in this case is controlled by the strain
rate of a solid matrix. At high melt fractions the
[30] Lets revisit the rheology of a suspension con- flow is supported by the soft phase, the pressure or
sisting of two phases. The rheology of a mixture of the shear stress is equal in two phases. Thus, this
two phases in general can be bounded by two situation corresponds to a lower Hashin-Strickman
curves which are called isostress model (the Reuss boundary or isostress case.
bound) and isostrain rate model (the Voigt bound),
for details see Ji and Xia [2002] and Takeda and [32] Basically, the behavior of a flow with or
Obata [2003]. These two viscosity bounds are without drainage can be understood as a smooth
valid for pure shear deformation. In the case of a transition from one Hashin-Shtrickman boundary to
uniaxial deformation the situation is more compli- another (see Figure A1). With respect to the Voigt
cated, instead of a shear stress, the sample is and Reuss averaging the Hashin-Strickman bounds
sheared because of the normal stress difference. exclude anisotropic microstructures in two phase
Until the total deformation remains small, i.e., up mixtures, but essentially they correspond to two
to the point when the sample becomes anisotropic limiting situations: the isostress (mHS) and isostrain
(i.e., no melt segregation, uniform compaction rate (mHS+) rheologies of a two component mixture.
regime), the measured viscosity is a combination [33] Concerning the dependence on the strain rate,
of the volume (bulk) and shear viscosity of sample, the relative viscosity of two phase mixture can be
and the interpretation of the experimental results is formally written in a form of Cross equation [e.g.,
based on the assumptions about bulk and shear Barnes, 1999]:
viscosities of two phases [McKenzie, 1984; Ricard
et al., 2001]. Splitting bulk and shear viscosities ðm0  m1 Þ
m ¼ m1 þ ; ðA2Þ
from the results of rock deformation experiments is 1 þ ðt  e_ Þm
not very common but it is still adequate if we
consider crystal concentrations in the medium to where m1 und m0 are the viscosities at very high
medium-high solid fraction range. and very slow strain rates, t is the characteristic
time of viscous stress relaxation and e_ is the strain
[31] Let us consider two more robust and general
rate. Here again, t represents an effective time
bounds of the effective relative viscosity of a two
scale, which separates two extreme rheological
phase mixture, that can be derived from Hashin-
cases: with and without building of the pressure
Shtrickman bounds (HS±). If we assume the anal-
difference between two phases. In equation (A2),
ogy between elastic and viscous cases of deforma-
for Newtonian viscosity m = 2, for non-Newtonian
tion, the Hashin-Shtrickman bounds for the shear
behavior 1 < m < 2. The relaxation time spectrum is
viscosity can be obtained from the expression for

10 of 13
Geochemistry 3
Geophysics
Geosystems G costa et al.: rheology of particle-bearing suspensions 10.1029/2008GC002138

Figure A1. Examples of relative viscosities of a two phase mixture and bounds calculated using equations (A1),
(A3), and (A4) using different parameters. The contrast of viscosities between the hard and the soft phase is assumed
to be 109; the strength of the solid phase is arbitrarily taken to be 108 Pa. In equation (A3) we set m = 2, and the stress
s was replaced by the product of a melt shear viscosity times the strain rate values indicated in the inset; in equation
(A4) we set C = 5104 Pa, 8K = 0.35, 8c = 0.95, n = 1, and k = 1.

difficult to infer for mixtures, in the work by Renner between these two limiting cases, the mixture has
et al. [2000], t has been related to the pore pressure two viscosities: at high stresses the mixture will
relaxation time. A more practical expression which flow with a smaller viscosity, whereas at low
includes shear stress s and pseudo–yield stress sy is stresses the viscosity will become higher and the
the Ellis empirical equation [Barnes, 1999]: flow is creeping. The variation of the yield stress sy
as a function of f between these two concentrations
ðm0  m1 Þ fK < f < fc is given by Barnes [1999]:
m ¼ m1 þ : ðA3Þ
1 þ ðs=sY Þm
ðf  fK Þn
In the case of a two phase mixture we may associate sY ¼ C  ; ðA4Þ
ðfc  fÞk
the low-stress rheology with the situation when the
liquid phase surrounds the solid phase m1 = mHS (at where C, n and k are empirical constants describing
f
1). The high-stress rheology corresponds to the how fast the yield stress sy varies between the two
rheology in the isostrain rate case or the upper limiting concentrations. Above the percolation
Hashin-Strickman boundary m0 = mHS+. The scale of threshold, f > fc, the yield stress sy is a smooth
stress s in equation (A3) can be taken formally as a monotonic function of melt concentration up to the
product of melt viscosity and strain rate. yield strength of the solid phase with closed
[34] A mixture of a solid and a liquid phase is porosity, f = 1. The effective shear stress appears
characterized by two concentration limits of the to be a monotonously increasing function from the
solid phase between which the yield stress sy varies solid phase fraction fK up to fc. Then, sy increases
as a smooth function of f. There are a nontouching stepwise with the decrease of porosity reaching the
limit of concentration, fK, and a percolation thresh- value of the yield stress of a solid matrix. The
old of the solid matrix, fc. According to Barnes critical porosity of the solid matrix fc is associated
[1999], this means that, if the solid fraction f is with the rupture of a critical cluster according to

11 of 13
Geochemistry 3
Geophysics
Geosystems G costa et al.: rheology of particle-bearing suspensions 10.1029/2008GC002138

the percolation theory [Guéguen and Palcialuskas, Auer, F., H. Berkhemer, and G. Oehlschlegel (1981), Steady
1994]. state creep of fine grain granite at partial melting, J. Geo-
phys., 49, 89 – 92.
[35] The introduction of an ‘‘effective’’ or a pseu- Bagdassarov, N., and A. Dorfman (1998), Granite rheology:
Magma flow and melt migration, J. Geol. Soc., 155, 863 –
do-yield stress sY can implicitly incorporate a 872, doi:10.1144/gsjgs.155.5.0863.
drainage effect of melt in the matrix and interaction Barnes, H. (1999), The yield stress – a review or ‘panta rei’ –
effects between melt and solid phase. For the everything flows?, J. Non-Newtonian Fluid Mech., 81, 133 –
relaxation of the melt overpressure a diffusion type 178, doi:10.1016/S0377-0257(98)00094-9.
of equation can been used, with a relaxation time Bercovici, D., Y. Ricard, and G. Schubert (2001), A two-phase
model of compaction and damage: 1. General theory,
inversely proportional to the permeability of the J. Geophys. Res., 106(B5), 8887 – 8906, doi:10.1029/
solid matrix. Approaching the percolation thresh- 2000JB900430.
old, the relaxation time tends to a power law on Caricchi, L., L. Burlini, P. Ulmer, T. Gerya, M. Vassalli, and
porosity [Renner et al., 2000]. The problem of the P. Papale (2007), Non-Newtonian rheology of crystal-bear-
description of this low melt concentration regime is ing magmas and implications for magma ascent dynamics,
Earth Planet. Sci. Lett., 264, 402 – 419, doi:10.1016/
out of the scope of this paper. j.epsl.2007.09.032.
[36] The transition from high-viscosity to lower- Caricchi, L., D. Giordano, L. Burlini, P. Ulmer, and C. Romano
(2008), Rheological properties of magma from the 1538 erup-
viscosity regime in creep experiments with partial- tion of Monte Nuovo (Phlegrean Fields, Italy): An experimen-
ly molten samples of olivine with MORB was tal study, Chem. Geol., 256, 158 – 171, doi:10.1016/j.chemgeo.
described by Scott and Kohlstedt [2006] in the 2008.06.035.
range of the strain rate from 106 to 104 s1. The Champallier, R., M. Bystricky, and L. Arbaret (2008), Experi-
underlined mechanism has been attributed to mental investigation of magma rheology at 300 MPa: From
pure hydrous melt to 76 vol. % of crystals, Earth Planet. Sci.
the transitional nature of the creep process, from Lett., 267, 571 – 583, doi:10.1016/j.epsl.2007.11.065.
diffusion creep at low strain rates to grain boundary Chong, J., E. Christiansen, and A. Baer (1971), Rheology of
sliding at high strain rates. The variation of the concentrated suspension, J. Appl. Polym. Sci., 15, 2007 –
grain size in samples influences the proportionality 2021, doi:10.1002/app.1971.070150818.
constant in equation (A4), C d2 [Barnes, 1999]. Costa, A. (2005), Viscosity of high crystal content melts: De-
pendence on solid fraction, Geophys. Res. Lett., 32, L22308,
In smaller grain size mixture the relative viscosity doi:10.1029/2005GL024303.
tends to the upper Hashin-Shtrickman boundary Costa, A., O. Melnik, and R. Sparks (2007a), Controls of
mHS+ and the rheologically critical melt threshold conduit geometry and wall rock elasticity on lava dome
shifts to larger melt fractions. eruptions, Earth Planet. Sci. Lett., 260, 137 – 151,
doi:10.1016/j.epsl.2007.05.024.
[37] As an example, in Figure A1 we show few Costa, A., O. Melnik, R. Sparks, and B. Voight (2007b), Con-
case of relative viscosities of a two phase mixture trol of magma flow in dykes on cyclic lava dome extrusion,
Geophys. Res. Lett., 34, L02303, doi:10.1029/
and corresponding bounds calculated using equa- 2006GL027466.
tions (A1), (A3), and (A4) using the parameter Gibilaro, L. G. (2001), Fluidization-Dynamics, 256 pp.,
values reported in the caption. Butterworth-Heinemann, Oxford, U. K.
Goetze, C. (1977), A brief summary of our present day under-
[38] Obviously this kind of description of a two phase standing of the effect of volatiles and partial melt on the
rheology is a simplified idealization. However, the mechanical properties of the upper mantle, in High-Pressure
shape of the curves plotted in Figure A1 obtained Research: Applications in Geophysics, edited by M. H.
from equations (A1), (A3) and (A4) mimics the trend Manghnani and S. Akimoto, pp. 3 – 23, Academic, New York.
Guéguen, Y., and V. Palcialuskas (1994), Introduction to the
of the experimental results and it is qualitatively similar Physics of Rocks, 294 pp., Princeton Univ. Press, Princeton,
to the behavior of the model discussed in this paper. N. J.
Hess, K. U., and D. B. Dingwell (1996), Viscosities of hydrous
Acknowledgment leucogranite melts: A non-Arrhenian model, Am. Mineral.,
81, 1297 – 1300.
[39] This work was supported by a NERC grant (NE/ Hess, K., B. Cordonnier, Y. Lavallée, and D. B. Dingwell
(2008), Viscous heating in rhyolite: An in situ experimental,
C509958/1). Discussions with O. Melnik and S. Muller greatly
Earth Planet. Sci. Lett., 275, 121 – 126, doi:10.1016/
improved the manuscript. The work benefited from critical and j.epsl.2008.08.014.
helpful reviews by J. Mecklenburgh, G. Hirth, and J. Renner. Hirth, G., and D. L. Kohlstedt (1995), Experimental con-
straints on the dynamics of the partially molten upper man-
References tle: Deformation in the diffusion creep regime, J. Geophys.
Res., 100, 1981 – 2001, doi:10.1029/94JB02128.
Arzi, A. (1978), Critical phenomena in the rheology of par- Ji, S., and B. Xia (2002), Rheology of Polyphase Earth Mate-
tially melted rocks, Tectonophysics, 44, 173 – 184, rials, 300 pp., Polytech. Int. Press, Québec, Que., Canada.
doi:10.1016/0040-1951(78)90069-0.

12 of 13
Geochemistry 3
Geophysics
Geosystems G costa et al.: rheology of particle-bearing suspensions 10.1029/2008GC002138

Kohlstedt, D. L., and M. E. Zimmerman (1996), Rheology of tion segregation, J. Geophys. Res., 100, 15,681 – 15,695,
partially molten mantle rocks, Annu. Rev. Earth Planet. Sci., doi:10.1029/95JB00077.
24, 41 – 62, doi:10.1146/annurev.earth.24.1.41. Rutter, E., and D. Neumann (1995), Experimental deformation
Krieger, I., and T. Dougherty (1959), A mechanism for non- of partially molten Westerly granite under fluid-absent con-
Newtonian flow in suspension of rigid spheres, Trans. Soc. ditions, with implications for the extraction of granitic mag-
Rheol., 3, 137 – 152, doi:10.1122/1.548848. mas, J. Geophys. Res., 100, 15,697 – 15,715, doi:10.1029/
Lejeune, A., and P. Richet (1995), Rheology of crystal-bearing 94JB03388.
silicate melts: An experimental study at high viscosity, Scott, T., and D. L. Kohlstedt (2006), The effect of large melt
J. Geophys. Res., 100, 4215 – 4229, doi:10.1029/94JB02985. fraction on the deformation behaviour of peridotite, Earth
Mavko, G., T. Mukerji, and J. Dvokin (1998), The Rock Phy- Planet. Sci. Lett., 246, 177 – 187, doi:10.1016/j.epsl.
sics Handbook, Tools for Seismic Analysis in Porous Media, 2006.04.027.
329 pp., Cambridge Univ. Press, Cambridge, U. K. Shaw, H. (1963), Obsidian-HO viscosities at 1000 and 2000
McKenzie, D. (1984), The generation and compaction of par- bars in the temperature range 700°C to 900°C, J. Geophys.
tially molten rocks, J. Petrol., 25, 713 – 765. Res., 68, 6337 – 6343.
Melnik, O., and R. S. J. Sparks (1999), Nonlinear dynamics of Solomon, M., and D. Boger (1998), The rheology of aqueous
lava dome extrusion, Nature, 402, 37– 41, doi:10.1038/46950. dispersions of spindle-type colloidal hematite rods, J. Rheol.,
Melnik, O., and R. S. J. Sparks (2005), Controls on conduit 42, 929 – 949, doi:10.1122/1.550961.
magma flow dynamics during lava dome building eruptions, Takeda, Y.-T., and M. Obata (2003), Some comments on the
J. Geophys. Res., 110, B02209, doi:10.1029/2004JB003183. rheologically critical melt percentage, J. Struct. Geol., 25,
Mueller, S., H. Mader, and E. Llewellin (2008), The influence 813 – 818, doi:10.1016/S0191-8141(02)00080-9.
of crystal shape on magma rheology, paper presented at Thomas, D. (1965), Transport characteristics of suspensions:
IAVCEI General Assembly, Int. Assoc. of Volcanol. and VIII. A note on the viscosity of Newtonian suspensions of
Chem. of the Earth’s Inter., Reykjavı́k, Iceland, 18 – 25 Aug. uniform spherical particles, J. Colloid Sci., 20, 267 – 277,
Pabst, W., E. Gregorova, and C. Berthold (2006), Particle doi:10.1016/0095-8522(65)90016-4.
shape and suspension rheology of short-fiber systems, J. Eur. van der Molen, I., and M. Paterson (1979), Experimental
Ceram. Soc., 26, 149 – 160, doi:10.1016/j.jeurceramsoc. deformation of partially melted granite, Contrib. Mineral.
2004.10.016. Petrol., 70, 299 – 318, doi:10.1007/BF00375359.
Petford, N. (2003), Rheology of granitic magmas during ascent van der Werff, J. C., and C. G. de Kruif (1989), Hard-sphere
and emplacement, Annu. Rev. Earth Planet. Sci., 31, 399 – colloidal dispersions: The scaling of rheological properties
427, doi:10.1146/annurev.earth.31.100901.141352. with particle size, volume fraction, and shear Rate, J. Rheol.,
Pinkerton, H., and R. Stevenson (1992), Methods of determin- 33, 421 – 454, doi:10.1122/1.550062.
ing the rheological properties of magmas at sub-solidus tem- Vigneresse, J. C., P. Barbey, and M. Cuney (1996), Rheologi-
peratures, J. Volcanol. Geotherm. Res., 53, 47 – 66, cal transitions during partial melting and crystallization with
doi:10.1016/0377-0273(92)90073-M. application to felsic magma, J. Petrol., 37(6), 1579 – 1600,
Renner, J., B. Evans, and G. Hirth (2000), On the rheologically doi:10.1093/petrology/37.6.1579.
critical melt fraction, Earth Planet. Sci. Lett., 181, 585 – 594, Voigt, W. (1928), Lehrbuch der Kristallphysik, Teubner,
doi:10.1016/S0012-821X(00)00222-3. Leipzig, Germany.
Reuss, A. (1929), Berechnung des Fliessgrenze von Mischkris- Wildemuth, C., and M. Williams (1984), Viscosity of suspen-
tallen auf Grund der Plastizitätsbedingung für Einkristalle, sions modeled with a shear-dependent maximum packing
Z. Angew. Math. Phys., 9, 49 – 58. fraction, Rheol. Acta, 23, 627 – 635, doi:10.1007/
Ricard, Y., D. Bercovici, and G. Schubert (2001), A two-phase BF01438803.
model for compaction and damage: 2. Application to compac- Yue, Y., C. Moisescu, G. Carl, and C. Russel (1999), The
tion, deformation, and the role of interface tension, J. Geo- influence of suspended iso-and anisometric crystals on the
phys. Res., 106, 8907 – 8924, doi:10.1029/2000JB900431. flow behavior of fluoroapatite glass melts during extrusion,
Roscoe, R. (1952), The viscosity of suspensions of rigid Phys. Chem. Glasses, 40, 243 – 247.
spheres, Br. J. Appl. Phys., 3, 267 – 269. Zhou, J. Z. Q., T. Fang, G. Luo, and P. H. T. Ulherr (1995),
Rosenberg, C., and M. Handy (2005), Experimental deforma- Yield stress and maximum packing fraction of concentrated
tion of partially melted granite revisited: Implications for the suspensions, Rheol. Acta, 34, 544 – 561, doi:10.1007/
continental crust, J. Metamorph. Geol., 23, 19 – 28, BF00712315.
doi:10.1111/j.1525-1314.2005.00555.x.
Rushmer, T. (1995), An experimental deformation study of
partially molten amphibolite: Application to low-melt frac-

13 of 13

Potrebbero piacerti anche