You are on page 1of 17

Pure Appl. Geophys.

168 (2011), 2259–2275


Ó 2011 Springer Basel AG
DOI 10.1007/s00024-011-0364-5 Pure and Applied Geophysics

Stick–Slip and the Transition to Steady Sliding in a 2D Granular Medium


and a Fixed Particle Lattice
J. KRIM,2 PEIDONG YU,1,3 and R. P. BEHRINGER1

Abstract—We report an experimental study of the stick–slip to imposed vibrations (SOCOLIUC et al. 2006; JOHNSON
steady sliding behavior of a solid object pulled, via a spring, across
2D granular substrates of photoelastic disks that are either fixed in a
and JIA 2005; JOHNSON et al. 2008; CAPOZZA et al.
solid lattice (granular solid) or unconstrained, forming a granular 2009) It is unclear, however, how the fundamental
bed. We observe a progression of friction regimes with increasing details of friction in these wide-ranging systems are
sliding speed, including single-slip, double-slip, and mixed stick–
linked. Studies that provide insight into the similari-
slip regimes, steady sliding, and inertial oscillations. For the case of
the granular bed, we report a detailed analysis of frictional behavior ties and differences between atomic-scale friction and
for the low speed stick–slip regime, including spring and elastic the stick–slip phenomenon manifested at the macro-
energy dependencies during the stick and slip portions of the scale are therefore of great interest. In this work, we
motion. For the case of the granular solid, we explore friction in the
presence and absence of externally applied vibrations, and compare focus on exploring how ideas that have been devel-
it with sliding on a granular bed, which is intrinsically disordered. oped for understanding atomic-scale friction can be
We observe that external vibration reduces transition values for extended to the intermediate length scales associated
both the single-slip to double-slip transition and the stick–slip to
steady sliding transition. Moreover, we observe that the effect of
with granular materials. In particular, we study the
packing disorder on granular friction seems similar to the effect of motion of an object, a slider, that rests on and is
vibration-induced disorder, a result that, to our knowledge, has not pulled across a two-dimensional (2D) granular sur-
been reported previously in the experimental literature.
face consisting of photoelastic disks. The photoelastic
disks are either fixed in a lattice, which we refer to as
a granular solid, or unconstrained, which we refer to
as a granular bed. Employing this apparatus, we
1. Introduction
observe a progression of friction regimes with
increasing sliding speed. These include:
Friction between two mutually sheared surfaces
appears in a multitude of systems from atoms to 1. single-slip stick–slip;
geophysical fault zones (KRIM 1996, 2002; BURRIDGE 2. double-slip stick–slip;
and KNOPOFF 1967) and many characteristics are 3. steady sliding; and
shared by these systems, despite dramatic differences 4. inertial oscillations.
in length scale. These include stick–slip to steady For the case of the granular solid, we also study
sliding transitions with increased strain rate or sliding friction in the presence and absence of externally
speed, (LUAN and ROBBINS 2004) and demonstrated applied vibrations, enabling comparison of how
sensitivities to disorder (BRAIMAN et al. 1999; MAK vibration-induced disorder compares and contrasts
and KRIM,1997; GENG et al. 2003) and externally with the disorder intrinsically present in a granular
bed. Our systems enable us to probe whether trends
predicted by atomic-scale models are present in a
1
Department of Physics, Duke University, Box 90305, system characterized by macroscopic granular parti-
Durham, NC 27708, USA. E-mail: bob@phy.duke.edu cles, and whether external vibration can have an
2
NC State University, Raleigh, NC, USA. effect similar to that of (thermodynamic) tempera-
3
Present Address: German Aerospace Center, Institute of
Materials Physics in Space, Linder Hoehe, 51147 Cologne,
ture, which does not play an explicit role in granular
Germany. systems.
2260 J. Krim et al. Pure Appl. Geophys.

At the atomic scale, inter-particle interactions are


well defined, and a variety of models have been
suggested to describe how static friction and stick–
slip phenomena at a sliding interface arise from the
molecular nature of the materials at the interface, and
how energy is dissipated through the excitation of
molecular and electronic excitations (ROBBINS and
KRIM 1998; FILIPPOV et al. 2004). Length scales here
are of the order of a few Angstroms, and the relevant
systems are highly sensitive to both temperature
(BAREL and URBAKH 2010; JANSEN et al. 2010) and
sliding velocity levels (LUAN and ROBBINS 2004).
Stick–slip events reported by atomic-force micros-
copy studies are generally associated with one-atom
slip lengths if sufficient resolution is present. Atomic-
scale friction models have, however, demonstrated
that in some cases the slip length may increase from Figure 1
Dolomite fault gouge from the Triassic Evaporites Formation in the
single-atom to double-atom slip lengths as the sliding
Northern Apennines of Italy, the source of the 1997 Colifiorito
velocity increases (NAKAMURA et al. 2005). Transi- earthquakes. From DE PAOLA et al. (2011)
tions between regimes of single and double slip are
affected by system disorder, with the double-slip
regime extending to lower sliding speeds for some earthquake triggering (JOHNSON et al. 2008), acoustic
system conditions (FAJARDO and MAZO 2010). fluidization (MELOSH 1979), characteristic times
Increases in temperature, meanwhile, have been associated with slip, and rupture modes, among oth-
predicted to reduce the slip length in a stick–slip ers. For a recent and useful review, see BEN-ZION
event (TSHIPRUT et al. 2009). One objective of this (2001).
study was to examine whether these features and Interestingly, granular materials exhibit friction
tendencies are present in macroscopic granular beds which is both a collective property of the many-grain
and solids. system, as shown by Coulomb (COULOMB 1773;
At the geophysical scale, interactions are macro- MUSER et al. 2003), and also involves ordinary fric-
scopic, classical, highly dissipative, and involve tional interaction between grains. The former is
complex structures that can extend for many kilo- because of the inherent interlocking of grains, and in
meters (BURRIDGE and KNOPOFF 1967). Granular fact occurs for frictionless grains. The latter is
materials are typically collections of millimeter to because of the frictional interactions at particle sur-
centimeter-scale solid particles, although, in princi- faces. The fact that intergrain interactions are
ple, smaller and larger particles are still granular in frictional means there is no description in terms of a
nature, and are also characterized by inter-particle conserved energy. Indeed, granular interaction forces
interactions that are purely classical in nature. do not depend simply on the inter-particle separation,
Granular materials are known to be present in fault and trace levels of films adsorbed on the particles
zones, which generally contain ground up rock, (PANELLA et al. 1996) can profoundly affect friction
gouge (MARONE et al. 1990; DE PAOLA et al. 2011), levels (MCFADDEN and GELLMAN, 1997), with minimal
that is granular in nature, as seen in Fig. 1. An changes in inter-particle separation (KRIM and
expected outcome of detailed laboratory studies, such BEHRINGER 2009).
as the work described here, is the potential to eluci- Disorder, in various forms, can have a significant
date the complex range of behavior seen in geological effect on friction. Temperature, in the thermody-
fault settings. Among these phenomena are the namic sense, does not play an explicit role in the
long-standing heat flow paradox, the possibility of granular case, and mechanical energy dissipated by
Vol. 168, (2011) Transition to Steady Sliding in a 2D Granular Medium 2261

friction or restitutional losses is lost to heat without initially spatially disordered state, without the appli-
otherwise affecting the system significantly. Con- cation of vibration. Spatial disorder at the atomic scale
versely, there is no flow of thermal energy back to has also been linked to friction reduction (BRAIMAN
mechanical energy in granular systems. The concept et al. 1999; MAK and KRIM 1997). In static granular
of granular temperature attempts to capture for systems, reduction of inter-particle friction or of
granular systems, some of the character of thermo- spatial disorder (through a change in polydispersity of
dynamic temperature of conventional matter. the grain sizes) has been shown to have a significant
External vibration of a granular system can, however, effect on the way in which forces are carried through a
potentially act in a manner similar to temperature, sample (GENG et al. 2003). Although both vibration
because it increases the kinetic energy of the granular and frozen spatial disorder in granular systems play
constituents. Such vibrational energy is captured in roles that mimic thermal disorder, at this stage there is
granular gas models, which borrow from the molec- no theory, to our knowledge, that can give a clear
ular theory of gases (GOLDHIRSCH et al. 1993). These quantitative connection between the degree of disor-
models exploit a generalized or granular temperature der or the strength of vibration and the resulting effect
defined in terms of the velocity variance. For denser on order–disorder in a granular system.
granular states (NOWAK et al. 1997), granular ‘‘fluids’’ A general experimental issue for characterizing
and ‘‘solids’’, there are still mechanical fluctuations, these diverse systems involves developing probes at
which can be quite significant, particularly in sheared the relevant scale. In the atomic case, atomic force
systems (HOWELL et al. 1999). And, the application of microscopy, microbalances, and other novel tech-
external vibration can have a dramatic effect on state niques have provided numerous important insights
selection, as seen in recent experiments (DANIELS and (KRIM 2002), although the nature of the contacts in
BEHRINGER 2005). There are also several proposals for these atomic-scale systems can be difficult to char-
new types of granular statistical mechanics (EDWARDS acterize in systems where stick–slip behavior is
and OAKESHOTT 1989; HENKES et al. 2007; MAKSE and prominent (FILIPPOV et al. 2004; BAREL and URBAKH
KURCHAN 2002). These approaches typically involve 2010). At the other extreme, fault zones have tradi-
developing new statistical ensembles based on tionally been studied by searching for faults that have
quantities such as total volume, or stress, which may been brought to the earth’s surface, by seismographic
well be conserved for granular systems, unlike the approaches, and by bore hole studies. Ordinary gran-
energy. Currently, these new versions of granular ular materials do not involve the complexities
statistics are in their infancy. Connections between involved with atomic or geophysical behavior, and
fluctuations and responses of granular systems are they can be studied directly in table-top experiments.
very much open issues, and our experiments may However, most granular systems, which are 3D in
provide insight into this subject. nature, can be difficult to observe in complete detail,
Mechanical vibration and disorder are known to because they do not allow simple optical access.
alter friction in a range of systems spanning the However, recent experimental developments provide
atomic to the macroscopic (SOCOLIUC et al. 2006; new approaches for studying these systems at the
JOHNSON and JIA 2005; JOHNSON et al. 2008; CAPOZZA grain scale. Optical access for 3D materials is possible
et al. 2009), although the fundamental mechanisms if an index-matched fluid occupies the interstices
are not well understood. In atomic systems, interfacial between grains (and if the grains are also transparent)
vibrations can act to induce interfacial melting at (TSAI et al. 2003). For quasi-2D materials, such as
temperatures far below ambient melting points those described here, it is possible to observe not only
(DAWSON et al. 2009; BOROVSKY et al. 2001). In the the detailed motion of the particles, but, through
granular experiments of DANIELS and BEHRINGER 2005, photoelastic approaches, measure the force response
the application of vibrations to a sheared granular of individual grains (MAJMUDAR and BEHRINGER 2005;
layer enabled the system to find states of ordered MAJMUDAR et al. 2007; ZHANG et al. 2010).
packing. Although these states were stable once gen- Several recent studies have probed the nature of
erated, they were generally inaccessible from an granular friction (NASUNO et al. 1997a, b; LOSERT
2262 J. Krim et al. Pure Appl. Geophys.

et al. 2000; AHARANOVand SPARKS 2003; SIAVOSHI ‘‘stick’’, the block is at rest, and Fp is less than ls Mg.
et al. 2006; YU and BEHRINGER, 2005; DANNIELS and As soon as |Fp| exceeds ls |Fn| the block slips, and
HAYMAN, 2008). NASUNO et al. (1997a, b) character- experiences a constant drag force from kinetic
ized the motion of a slider driven over a granular friction, lk Mg \ ls Mg, where lk is the kinetic
surface formed from glass spheres or sand. However, friction coefficient. This retarding force acts for a
a connection between the internal granular and slider time Dt during which the block slips a distance Dx:
dynamics was not accessible. Other recent work Thereupon, the block returns to rest, and the frictional
involves the use of photoelastic particles (YU and force reverts to its static form. The dynamics of the
BEHRINGER 2005; DANIELS and HAYMAN 2008; YU slider are simple: F ¼ M€ x ¼ ks ðxo þ Vt  xÞ  Ff :
et al. 2011). Our work builds on the earlier pre- Here, xo is a reference position which is determined
liminary studies of YU et al. (2011) and YU and by the initial state of the system. For instance, if at
BEHRINGER (2005). It combines photoelastic/force t = 0, x = 0 and Fp = ls Mg, the maximum
information with detailed measurements associated allowed static frictional force, then xo = lsMg/ks.
with motion of an object, a ‘‘slider’’, that is pulled When the block is ‘‘stuck’’, Ff is given by the static
across a granular surface. friction case, with x_ ¼ x€ ¼ 0: During sliding, the
solution for x(t) is easily obtained as sinusoidal plus
linear components in time t : xðtÞ ¼ A cosðxt þ
/Þ þ Vt þ ðls  lk ÞMg=k; and x2 = ks/M. This form
2. Stick–Slip Friction Models
introduces a characteristic length scale:

2.1. Simple Block-Spring Model b ¼ ðls  lk ÞMg=k ¼ DlMg=ks ; ð1Þ

A variety of models have been proposed to and a characteristic frequency for inertial oscillations:
characterize sliding between two frictional objects.
x ¼ ðks =MÞ1=2 : ð2Þ
As a starting point, we consider a simple ‘‘block’’
model that assumes the textbook features of friction It is also useful to define an inertial time scale,
for a slider that is pulled across a frictional surface. sint = 2p/x. As above, this assumes that at
The model consists of two frictional objects, the t = 0, x = 0, and Fp = lsMg and that the block is
slider and an underlying surface, that interact at a at rest. A = (V2o ? V2)1/2/x, and tanð/Þ ¼ V=Vo :
planar interface where there are normal and tangen- Here, Vo = bx is a natural velocity scale. A different,
tial forces, Fn and Ft. Static friction supports but clearly related, time scale can be constructed
tangential forces when |Ft| \ ls |Fn|, where, ls is from b=g ¼ Dl=x2 : Thus, M, b, and x provide a
the static friction coefficient. When |Ft| exceeds complete set of dimensioned variables for this
ls |Fn|, sliding occurs with a frictional drag force system. A length scale that is not present in this
of lk Fn. When the velocity of the slider reaches zero, inherently macroscopic model is the grain size, d.
friction reverts to its static form. In all cases, in this Slip occurs until x_ once again reaches 0. The time
scenario, there is an instantaneous switch between duration of this slip is Dt ¼ ð3p  2/Þx; and the slip
static and sliding friction. Although this model is distance is Dx ¼ VDt þ 2b: When the slider once
ultimately limited, it can still furnish useful insights again comes to rest, the pulling force is Fp ¼
into important physical scales. Mgð2lk  ls Þ: Consequently, the next slip event will
In the model, a machine moving at velocity V occur after a stick time sst ¼ 2MgDl=ðVkÞ ¼ 2b=V;
pulls with force Fp on the block through a spring of and thereafter, the stick–slip will repeat with a period
force constant ks. Fp and the block motion are along of sst þ Dt: An additional calculation yields:
the horizontal direction. In the vertical direction, the
ls  lk ¼ ðx2 =2gÞðDx  DtVÞ: ð3Þ
force of gravity is balanced by a constant normal
force Fn = Mg. We initially imagine applying this Note, too, that the maximum stored potential
model to a regime of V and ks that are ‘‘small’’, so energy in the spring is US = (FP)2/(2ks) if we use the
that stick–slip motion of the block occurs. During maximum value of Fp = ls Mg. Then the maximum
Vol. 168, (2011) Transition to Steady Sliding in a 2D Granular Medium 2263

for potential energy in the spring is US = (ls Mg)2/ chains evolve, bend, fail, and reform during the
(2ks). course of a slider experiment. Force chains contacting
The regime of small V/Vo is of some interest. In the slider are analogous to the asperities that are so
this case, / ’ p; and Dt ’ sint =2; i.e. there is important in ordinary friction.
approximately a half period of sinusoidal motion Time-resolved slip events show that the motion of
for the slider velocity during the slip regime. In this the slider can be more complex than one might expect
regime, the period of the stick–slip motion is from the simple model. Instead of having a slip phase
dominated by the stick time sst. We can approximate consisting of approximately a half cycle of oscillatory
the amplitude of this sinusoidally varying velocity as motion, events, particularly large ones, can consist of
Vo = bx. The peak kinetic energy during this a sequence of approximately half-cycle events, where
motion is then Kmax ¼ MVo2 =2 ¼ ðDlMgÞ2 =k: In this one event begins before the previous event has ended.
regime, we can use Dl ¼ ls  lk ’ ðx2 =2gÞðDxÞto One possible explanation of this behavior is that the
obtain Dl from measurements of Dx: Then the ratio granular material resists the pulling motion through
of peak kinetic energy during slip to peak stored force chains, which are discussed in more detail
energy in the spring is given simply as (dl/ls)2. An below. During a slip event, some of these chains
important dimensionless property of our system is break, others can rearrange substantially, and new
the ratio of the typical slip distance to the particle chains form. This complex process involves both the
diameter: dynamics of the slider and the micro-scale and meso-
scale dynamics of the chains. From this perspective,
R ¼ 2b=d ¼ DlMg=ðdks Þ: ð4Þ
fixed ls and lk are rough approximations only.
For typical parameter values in the experiments, In fact, experiments and simulations typically
R ’ 1;and we will see below evidence of these reveal a more complex frictional force that is often
competing length scales. characterized by logarithmic rate-dependence (DIETE-
RICH 1972; RUINA 1983; BAUMBERGER et al. 1999;
BOWDEN and TABOR 1954; HESLOT et al. 1994; MUSER
2.2. More Complex Behavior and Models
et al. 2003; MARONE et al. 1990; HARTLEY and
The simple spring-block-friction model above BEHRINGER 2003). This behavior can be explicitly
gives useful scales for mass, time, and length, and encapsulated in rate-and-state models (DIETERICH
it predicts periodic stick–slip motion irrespective of 1972; RUINA 1983). It has been postulated that this
V or k. In fact, experiments, including those described rate dependence is tied to activated processes, which
here, do show stick–slip motion for small V and small are thermal in nature (EYRING 1936), or which are a
ks. However, they also exhibit transitions to other result of force fluctuations induced by the shearing
types of motion, including non-stick–slip periodic process (BEHRINGER et al. 2008). In both these latter
motion, and nearly steady sliding which require more pictures for log rate dependence, there must be
sophisticated friction models. fluctuations (either of energy or of stress), which lead
These experiments also show more complex to logarithmic dependences of stress on shear rate
behavior in the stick–slip regime. There are several because of activated processes described through
identifiable sources of this complex behavior. First, Boltzmann factors. The atomic-scale friction com-
the granular bed, as described below, is spatially munity is actively exploring the details of such
disordered (to avoid crystallization). Second, the models (JINESH et al. 2008).
forces are carried on mesoscopic structures, force Theoretical treatments of stick–slip phenomena
chains (Fig. 2 below), which are approximately linear frequently focus on how to eliminate them (LUAN and
collections of grains that carry above-average ROBBINS 2004). It is well known that one way to
amounts of force. At any one time, the number of eliminate stick–slip is to increase the sliding velocity,
grains that are in force chains terminating on the resulting in a transition to steady sliding. However,
slider correspond to approximately 10–20% of the theoretical predictions of this transition speed have
number of grains along the bottom of the slider. Force yielded values that are orders of magnitude higher
2264 J. Krim et al. Pure Appl. Geophys.

than experimental reports, particularly for atomic-


scale measurements employing atomic-force micros-
copy. In addition, the transition to steady sliding does
not usually occur at a single velocity: Instead a
transition regime from strictly stick–slip motion to
intermittent sliding, to steady or oscillatory motion is
quite common, with bifurcations and chaotic system
dynamics all possible (LUAN and ROBBINS 2004;
ELMER 1997; BRAUN et al. 2005; HESLOT et al. 1994).
At the other extreme, experiments on solid and
granular friction frequently demonstrate dependence
of the effective friction coefficients on the rate of
sliding and the age of contacts. The block–spring
model with simple static and kinetic friction coeffi-
cients, ls [ lk, demonstrates neither the bifurcations
nor the rate/aging effects described above. One class of
models frequently used in the geophysical community
to capture these effects are the rate-and-state models
(DIETERICH 1972; RUINA 1983; HESLOT et al. 1994;
CARLSON et al. 1994). For instance, the effective
friction coefficient is proposed to depend on the
sliding velocity and on a state variable, h:

l ¼ lo ¼ a lnðx=V
_ o Þ þ b ln½ðVo hÞ=Dc  ð5Þ
where h evolves as

h_ ¼ 1  xh=D
_ c ð6Þ
Figure 2
Necessarily, this model introduces a microscopic Top Sketch of apparatus. Second from Top Sketch showing the
granular region near the bottom of a typical slider. Third from Top
scale, Dc. For the block–spring model, as applied to Photo of the granular bed. The upper black region shows a portion
atomic or molecular scales, the transition to steady of the metal slider, moving to the right. The bottom of the slider is
sliding speed is frequently reported as: corrugated with semi-circular cut-outs that are approximately the
size of a particle diameter. Brighter grains are subject to larger
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
vc ¼ c Fs d=M ; ð7Þ forces, and exhibit the characteristic force chains, which tend to be
oriented along the compressive direction for the shear. Fourth from
where d = d/2 is the lateral displacement over which Top The fixed lattice of 5 mm photoelastic disks. Again, the upper
black region of this image shows a portion of the metal slider,
the substrate potential changes from a minimum to a moving to the right. Below the disks is a solid base to which the
maximum, Fs is the static friction force, and where c disks are glued. Again, bright particles experience larger forces.
is of order unity. (Once again, a microscopic scale is For the case of the fixed solid lattice, the bright regions reveal the
higher order commensurability of the slider with the substrate. The
involved). Equation 7 is, in essence, a balance
slider in this case has 56 lattice spacings for each 50 substrate
between the stored kinetic energy of the slider just particles. In the two lowest images, a fixture (not shown) atop the
before a slip, and the frictional energy dissipation slider supports a permanent magnet that allows the slider to be
occurring over that slip, assumed to be one lattice vibrated horizontally by externally applied magnetic fields. The
two different granular configurations, namely the bed and solid,
spacing in extent. For our experiments, where Fs is provide contrast between the picture of stick and slip from the point
set bypffiffiffiffiffiffiffiffiffiffiffiffiffiffi
the weight of the slider, we obtain of view of granular physics and from solid friction, in which two
vc ¼ c ls gd=2 ’ 12cm/s, assuming that c ’ 1: contacting solids have a true contact area determined by a limited
number of contacting asperities. In this experiment, the chains
In order to explain some apparent discrepancies present in the granular bed act as the counterparts to multi-asperity
between theory and experiments, Luan and Robbins contacts
Vol. 168, (2011) Transition to Steady Sliding in a 2D Granular Medium 2265

(LUAN and ROBBINS 2004) have considered two time vint ¼ dx; ð10Þ
scales associated with the slider:
is 4.4 mm 9 20 s-1 ^8 cm/s for the bulk granular
td ¼ d=vc ; ð8Þ material. Indeed, the ratio (vc/vint)2 = lgM/(dks) ^ R2.

and

ts ¼ 2H=cs ; ð9Þ 3. Experimental Techniques

where H is the height of the slider, which, for many


3.1. General Features
macroscopic slider geometries, is proportional to M.
The quantity cs is the speed of sound in the slider, and The experiments described here focus on the
ts is the time for elastic waves to propagate back and nature of friction between a solid surface and a
forth over the height of the slider. If td [[ ts, the quasi-2D granular system consisting of photoelastic
authors argue that stress can equilibrate throughout disks made from a commercial polymer (PSM,
the slider and there will be no change in the elastic Vishay Measurements). As shown in Fig. 2 Top,
energy of the slider as it stops in a slip event. In this these disks are confined in a channel, formed by two
case, Eq. 7 is expected to be applicable. For systems vertically oriented Plexiglas sheets that are separated
where ts [[ td, this result may not be applicable. by a spacing that is slightly larger than the thickness
For our case, there is also the possibility of of the particles. A cart is driven along a track located
acoustic propagation reflecting off the bottom bound- above the channel, and mounted on this cart are a
ary of the granular layer. Although the speed of sound camera, a light source, and a force gauge, to which
through the slider is very high ( 5000m/s), the one end of the pulling spring is attached. The other
speed of sound through the granular material is only end of the spring is attached to the ‘‘slider’’, a solid
of the order of a few m/s. In this regard, we note that metal object that has a width comparable with the
the speed of sound through the bulk material from thickness of the grains, and that rides on the upper
which the particles are cut is 56 m/s. The speed of surface of the granular sample. The force gauge
sound through a granular assembly is typically as measures the pulling force, Fp, and the camera/light
much as an order of magnitude less than the bulk source detect the photoelastic response of the grains.
material speed. Hence, we estimate the speed through Measurements of Fp are synchronized with camera
the granular assembly here as *10 m/s. frame grabs. The light from the source passes
d is the corrugation size in the atomic case. For our through a first circular polarizer, then through the
case, the corrugation is set by the mean grain spacing of disks along their axial direction, then through a
the rigid lattice, 5mm, so d = 2.5 mm. The static second crossed circular polarizer, and finally into the
friction force, while quite variable, is of the order of 0.5– camera. The resulting images of the stressed photo-
1 N, and the mass is 0.17 kg. Taking c  1; yields the elastic particles show light and dark bands within
predicted crossover to steady sliding close to 12 cm/s. each disk, and the patterns of the bands provide
The characteristic time d/vc is then 0.021 seconds. To quantitative information on inter-particle forces
estimate the speed of sound in the metal slider we take (MAJMUDAR and BEHRINGER 2005; MAJMUDAR et al.
cs = 5,000 m/s and a slider height of 5 cm, to obtain 2007; ZHANG et al. 2010). In particular, those
ts = 2 9 10 -5 s. Equation 7 should therefore be particles which experience a relative large force
applicable. However, the time for acoustic propagation appear bright in a typical image, for example Fig. 2,
from the slider, through the granular material, and back bottom two images. Bright regions below the slider
is longer than td, approximately in the range 0.05– are particles experiencing a large force. These tend
0.1 s. One objective of this work is to examine whether to be organized along force chains oriented along the
Eq. 7 is applicable to our experiments. compressive direction for the shear. The sliders are
We note another possibly relevant velocity scale brass or aluminium and are patterned on the bottom
which is set by the inertial time and the particle size. by half-round cut-outs of diameter comparable with
This velocity, (but not equal to) d.
2266 J. Krim et al. Pure Appl. Geophys.

3.2. Granular Bed


0.7
The granular bed consists of a vertically oriented Stick

layer of approximately 10,000 particles. The layer is 0.6 Slip


confined by walls at the ends and at the bottom, but it ΔF

Fp /Mg
is free on the upper surface; the channel length and 0.5
ΔF
depth are respectively 1.5 m and 15.5 cm. The
particle size distribution is bidisperse, with the 0.4 Δt

smaller and larger particle diameters given by 4.0


Δt
and 5.0 mm, respectively. The number fraction of 0.3

smaller particles is 0.63, yielding a mean diameter of 1600 1700 1800 1900 2000
d = 4.37 mm. A bidisperse distribution is chosen to Time (s)
avoid long-range crystalline order.
M = 219.0 g, V = 0.033 cm/s, ks = 61 N/m
0.9 M = 219.0 g, V = 0.033 cm/s, ks = 34 N/m
M = 73.0 g, V = 0.033 cm/s, ks = 34 N/m
3.3. Additional Features 0.8

Two additional aspects of these experiments 0.7

Fp/Mg
involved:
0.6
1. using a fixed ordered substrate, the granular solid;
0.5
and
2. probing the effect of added vibration. 0.4

0.3
For the case of the granular solid, which is seen in 1600 1700 1800 1900 2000
the bottom-most part of Fig. 2, we glued a single Time (s)
ordered layer of particles to a strip of metal. The
Figure 3
substrate particle center-to-center spacing is 5 mm. Top Typical time series for the pulling force, with a stick and
We used a slider that has 56 lattice cutouts for every slip event highlighted (Fp, with M ¼ 219:0 g; V ¼ 0:033 cm s1 ;
50 substrate particles. In order to apply external ks ¼ 34:3 N m1 ) Bottom Examples of stick–slip behavior for other
experimental conditions. All data in this figure are for a granular
vibrations to the system, we added a permanent bed
magnet atop one of the sliders and a pair of coils that
moves with the cart and that surround the permanent
details. In a typical experiment for this regime, the
magnet. The axes of all magnets are aligned along
cart advances steadily, and the slider remains nearly
the horizontal direction. By applying a current to the
at rest during the stick phase. In fact, there is small
coils, we can apply an AC horizontal force to the
but detectable deformation of the grains because the
slider with a vibration amplitude of *0.1 mm. More
effective spring constant for the material is an order
details of the experiment can also be found elsewhere
of magnitude greater than the typical pulling spring
(YU et al. 2011).
constant. Fig. 3 shows typical time series for
Fp, where we have normalized by the slider weight,
Mg. Fp increases nearly uniformly at a rate set by V
4. Experimental Results
until a slip event occurs. Slips occur over time
intervals that are typically much shorter than those
4.1. Granular Bed: Stick–slip Regime
for stick, and the slider advances at a speed in excess
In the first set of experiments discussed here of Vp. At the end of a slip event, Fp is reduced, and a
and elsewhere (YU et al. 2011), V \ 1 mm/s, new stick period ensues. For V values and ks values in
Vo [[ vc [ V, and the motion clearly is described the stick–slip regime, we observe non-periodic
by stick–slip. We briefly discuss results presented events. Because our material is spatially disordered,
elsewhere (YU et al. 2011), and add some additional and because the number of strong force chain
Vol. 168, (2011) Transition to Steady Sliding in a 2D Granular Medium 2267

contacts is relatively small (typically *15 or less, as 0.6 0.8

in Fig. 2 Third from Top), there is no compelling

x position of the slider (cm)


0.55 0.6
reason to expect periodic stick–slips.
0.5 0.4
The stick and slip processes are relatively easily

Fp/Mg
distinguished, because they are characterized by slow 0.45 0.2
and rapid motion. We define a stick (slip) event Creep
0.4 0
(Fig. 3 as any increase (decrease), DFp ; of Fp with
time that is larger than a threshold slightly larger than 0.35 Fp -0.2
x
the force gauge resolution. In fact, the values of Fp at 0.3 -0.4
which slips start (Fstart) or end (Fstop) are broadly 0 1 2 3 4 5
Time (s)
distributed, as shown in YU et al. (2011). These
probability distribution functions (PDFs) are insensi- Figure 4
Example of creep preceding a slip event, as seen in the pulling
tive to modest increases of the above threshold. We
force, Fp and in the slider position
also obtain PDFs for the duration over which stick
(slip) occurs, Dt; and the energy changes of the pulling
spring, DEp : As seen by YU et al. (2011), the PDFs for discussed in more detail below, we estimate the
Dt are described relatively well as nearly exponential effective spring constant for the grains as kg ’ 250N/
in Dt=tint ; and the PDFs for DEp scaled by (Mg)2/ks are m. If we think of the grains and pulling spring as two
described relatively well as a universal power-law springs in series, acted on by the force Fp, then the
with an exponent slightly greater than 1.0. effective spring constant will be ke = ks kg/(ks ? kg).
We note that when Fp is at a maximum just before The energies stored in the spring and in the grains are,
the beginning of a slip event, Fp = Fstart, the static respectively, F2p/(2kp) and F2p/(2kg), and the fraction of
friction coefficient is simply ls = Fp/Fn = Fstart/Mg the energy stored in the grains is ð1 þ kg =ks Þ1 : DEg
where the normal force is set by Fn = Mg. As shown ranges from 10 to approximately 30% of the energy
by YU et al. (2011), typical PDFs of Fstart/Mg and loss for springs used in the stick–slip regime, as in
also Fstop/Mg, are broad, nearly overlapping, and Fig. 5, for example. Note that in this picture, the total
roughly gaussian (see also ALBERT et al. 2001). The energy loss is an O(1) factor different from the energy
2 2
means of the PDFs for both Fstart/Mg and Fstop/Mg are loss in the spring, and hence, DE ¼ ðFstart  Fstop Þ=
close to 0.5 and have variances of *0.07. Necessar- ð2ks Þ provides a reasonable, though not complete
ily, however, the mean for Fstart/Mg is slightly higher measure of energy lost in a slip event.
than the mean for Fstop/Mg. The PDFs for both Data for DE are of some interest, because they can
quantities do not change significantly for different V be compared with data for earthquake magnitudes,
and ks in the stick–slip regime. m, and the well known Gutenberg–Richter (GR)
As would be expected, there is creep immediately power-law distribution, N(m). We have shown (YU
preceding slip events, as shown in Fig. 4. On the scale et al. 2011) that the PDFs for the energy loss at slips
of Fig. 3, this motion is not easily observable. In the can be described by a power law with an exponent
low-V case, the elementary block–spring model -1.2 ± 0.1. Some comments about these results are
indicates that slip occurs for roughly half a period in order. First, the fact that the data from our
of sinusoidal motion at frequency x. It is then experiments show a power law is interesting in its
straightforward to show that ls  lk ¼ ðx2 =2gÞ own right, because it indicates that slip events involve
ðDx  DtVÞ: Hence, the effective static and kinetic complex non-repetitive processes. Of course, com-
friction coefficients are determined: as above, Fp/Mg parison with the GR relation can only be made in a
at failure yields ls, and determination of Dx and Dt heuristic sense, because the details of the experiment
yields lk. and the multiple faults that are represented in a
The change in energy after a slip event is DE ¼ typical GR distribution are very different. Second,
DEp þ DEg where DEg is the potential energy stored there are several details involved in comparing our
inside the granular disks, and Ep = F2p/(2ks). As PDFs of slip energies and the GR law. The GR
2268 J. Krim et al. Pure Appl. Geophys.

t25v033f30d1009E4-15
300 50
Energy in spring (gram*cm)

Energy in grains (gram*cm)


250 41.667 2
10

P(Δ Fp/Mg)
200 33.333

0
150 25 10
Energy in the spring m=219g, k =107 N/m, V=0.033cm/s
s
Energy in the grains m=219g, k =61 N/m, V=0.033cm/s
s
100 16.667 m=219g, k =34 N/m, V=0.033cm/s
0 2000 4000 6000 8000 s
m=219g, k =34 N/m, V=0.066cm/s
data indexes (30 Hz) −2
10
s
m=219g, k =34 N/m, V=0.017cm/s
s
Figure 5 m=73g, k =34 N/m, V=0.033cm/s
s
Pulling spring energy (black) and granular elastic energy (red)
−3 −2.5 −2 −1.5 −1 −0.5 0
during stick–slip motion. Here, M = 219.0 g, V = 0.033 cm/s, and
kp = 34.3 N/m log10(Δ Fp/Mg)

Figure 6
PDFs for drops, DFp ; after a slip. Note that these distributions show
relation is a cumulative probability distribution scaling collapse when DFp is normalized by Mg
function (CDF). In YU et al. (2011), we present our
data as a probability distribution function (PDF),
P(m)dm, which is related to the CDF by NðmÞ ¼ that largest spring constant considered in the previous
R1 0 0
m Pðm Þdm : The GR CDF is log10N(m) = a - bm, section and we explore velocities that are up to 150
where b ^ 1, from the global range of earthquake times greater.
data. m can be related to the seismic energy Es by In all cases, we observe transitions from stick–slip
log10 ðDEs Þ ¼ 1:5m þ 4:8; and the PDF of DEs is dynamics to oscillatory dynamics, sometimes with an
related to the PDF of m by PðDEs Þ ¼ PðmÞdm= intervening steady sliding state. Before turning to a
dðDEs Þ. The PDF for DEs is PðEs Þ / DEs where discussion of oscillatory sliding, we consider possible
 ¼ 1 þ 2b=3: If b = 1, then  ’ 5=3: (See BEN-ZION oscillatory frequencies associated with our system. A
(2008) for a more detailed discussion.) In fact, this mass and spring system with mass 170 g and spring
power-law is reflected in the force drops after slip constant 160 N/m oscillates with frequency 4.88 Hz,
events, as shown in Fig. 6. The deviations from a or period 0.205 s, referred to here as inertial oscil-
power law at the lower and higher ends of the data are lations. Note that this frequency may be modified
because of measurement resolution and finite slider because of the elasticity of the grains, that is, there is
mass, respectively. an effective spring constant for the spring plus grains,
ke = kskg/(ks ? kg), that is lower than the bare spring
constant alone. The slope of the force versus distance
4.2. Granular Bed and Solid: Stick–slip
curve for the granular solid in the stick–slip low
to Steady Sliding
velocity regime (without vibrations) provides an
Figures 7, 8 and 9 present pulling force, Fp, ver- effective spring constant for the system, k = 100 N/m,
sus puller displacement in steps of 0.1 mm, for a which incorporates the effects of the elasticity of the
170 g slider pulled along by a 160 N/m spring, for, photoelastic disks. This implies a grain spring
respectively, the granular bed, the granular solid with constant of *250 N/m, which is consistent with an
no applied vibration, and the granular solid under the independent calibration made by measuring the grain
influence of external 11 Hz vibrations. In each case, displacement in response to an applied force. Taking
we show a puller displacement of 10 cm. Note that M = 0.17 kg, and k = 100 N/m, gives a frequency of
because the puller is moving uniformly, puller 3.86 Hz, and a period of 0.259 s. Interestingly, the
displacement can be converted simply to elapsed effective spring constant seems to be slightly higher for
time. We use a spring constant that is 60% larger than the granular solid with vibrations, i.e. ke ’ 120N/m,
Vol. 168, (2011) Transition to Steady Sliding in a 2D Granular Medium 2269

Figure 7
Granular bed, (no external vibration): pulling force versus distance in steps (1 step = 0.1 mm) for a 170 g slider and 160 N/m spring at pulling
speeds ranging from 0.2 to 5 cm/s. Stick–slip events are not observed above 1 cm/s. Modulated inertial oscillations, characteristic of the
crossover to steady sliding, are observed at 5 cm/s. Their frequency is consistent with inertial oscillations of a 170 g mass on a 160 N/m
spring, 4.88 Hz (period = 0.205 s)

giving a slightly higher frequency of 4.22 Hz, with a speeds, is much less sharp than for the slower speeds
slightly shorter period of 0.236 s. For the granular and weaker spring constants discussed earlier. Nev-
bed, the data in the stick–slip regime are noisier, but ertheless, for pulling speeds of 0.5–1 cm/s we see
we estimate ke ’ 68N/m, hence an inertial frequency relatively organized slips with a typical size scale that
of 3.18 Hz and period of 0.314 s. is roughly comparable with the size of a particle (or
Figure 7 shows data for the granular bed without the size of the cut-outs on the bottom of the slider).
vibration for pulling speeds ranging from 0.2 cm/s to For a pulling speed of 0.5 cm/s, the typical peak-to-
5 cm/s. The stick–slip character, even for the lower peak distance in the force-displacement data is
2270 J. Krim et al. Pure Appl. Geophys.

Figure 8
Granular solid, (no external vibration): pulling force versus distance in steps (1 step = 0.1 mm) for a 170 g slider and 160 N/m spring at
pulling speeds ranging from 0.2 to 10 cm/s. Stick–slip events are not observed above 2 cm/s. A distinct change in behavior is observed at
5 cm/s. Modulated inertial oscillations, characteristic of the crossover to steady sliding, are observed at 10 cm/s. Their frequency is consistent
with inertial oscillations of a 170 g mass on a 160 N/m spring, 4.88 Hz (period = 0.205 s). Lattice/grain size dimensions (4.5 mm) are
displayed in the 0.2 cm/s plot. Slip events correspond, on average, to an integral number of lattice steps

*0.47 cm, corresponding to a time of 0.94 s; for *1.8 cm, corresponding to a time of 0.36 s. This
1 cm/s, the peak-to-peak distance is *0.57 cm, time is close to the estimated period for inertial
corresponding to a time of 0.56 s. We will see below oscillations for the granular bed case, namely 0.314 s.
that for the granular bed the effect of the character- Figure 8 presents data for Fp versus puller
istic size of a particle/slider-roughness is even displacement for sliding on the granular solid without
clearer. For a pulling speed of 5 cm/s, periodic vibration, where the cart is moving at speeds of
dynamics occur, with a typical distance per period of 0.2–10 cm/s. Unlike the case of a disordered granular
Vol. 168, (2011) Transition to Steady Sliding in a 2D Granular Medium 2271

Figure 9
Granular solid, vibrated at 11 Hz with a peak amplitude of *0.2 mm: pulling force versus distance in steps (1 step = 0.1 mm) for a 170 g
slider and 160 N/m spring at pulling speeds ranging from 0.2 to 10 cm/s. Stick–slip events are not observed above 2 cm/s. Modulated inertial
oscillations, characteristic of the initial onset of the crossover to steady sliding, are observed at 5 and 10 cm/s. Their frequency is consistent
with inertial oscillation of a 170 g mass on a 160 N/m spring, 4.88 Hz (period = 0.205 s). The 0.2 cm/s data do not include the starting point
for the motion

bed, the stick–slip is more periodic, and the slip displacement and slip distances change from d to
distances are close to that of the grain dimensions or 2d or 3d. We note that the average slip distances are
multiples of it. (displayed in the 0.2 cm/s plot) As the closer to multiples of the periodicity roughness
pulling speed is increased from 0.2 to 2.0 cm/s, the profile of the slider than that of the grains. A
typical force drop and slip distance of an event transition/bifurcation to steady sliding is evident for
increases by approximately a factor of two in each 5 cm/s. The fast oscillations at the beginning of the
case. This occurs because the characteristic puller 5 cm/s run are apparently a transient, and the nearly
2272 J. Krim et al. Pure Appl. Geophys.

constant pulling force in the second part of the run are velocity of Ax = 1.4 cm, which is squarely in the
the steady state. Above this speed, the system middle of the range considered here, in which a
displays inertial oscillations, similar to those reported number of different transitions occur. It may there-
by NASUNO et al. (1997a, b) for granular layers fore be expected that the vibrated system would
consisting of glass spheres. The period of these display transitions at lower sliding speeds than the
oscillation, 0.205 s, is displayed in the 10 cm/s plot. unvibrated slider.
We can also compare this with the prediction of Eq. 7 Comparing the unvibrated granular solid data,
for the crossover velocity to steady sliding. Taking M Fig. 8, with the vibrated granular solid data, Fig. 9,
= 0.17 kg, and k = 100 N/m, and a characteristic several features stand out:
distance of one disk, the prediction of Eq. 7 for the
1. With vibration, a steady sliding regime is not
crossover velocity to steady sliding is approximately
observed for the sliding speeds studied.
10 cm/s. Here, we take c of order unity, Fs to be the
2. With vibration, the system transitions to inertial
static friction force, Fs = 1 N, and d to be half of the
oscillations at a lower sliding speed.
characteristic distance 5 mm. This estimate corre-
3. At the lower speeds, at least, vibration regularizes
sponds reasonably well to the observed velocity of
the slips, and reduces the typical change in force
5 cm/s for which we observe steady sliding.
following a slip.
Figure 9 presents data for the granular solid in the
4. The transition between slips of d and 2d, etc., is
presence of externally imposed horizontal vibration
much cleaner with vibration than without.
at a frequency of 11 Hz, close to the third overtone of
5. The presence of vibrations tends to elevate
the system’s inertial resonant frequency. In this
the transitions between stick–slips from nd to
figure, the effect of increased pulling speed is
(n ? 1)d.
qualitatively similar to what is seen in atomic scale
stick–slip models of moderate strength substrates as In some regards, the vibration affects the system
velocity is increased (NAKAMURA et al. 2005). In in a manner that is qualitatively similar to an increase
particular, at low sliding speeds, the slip length in the of disorder in molecular systems (NAKAMURA et al.
stick–slip corresponds to the particle/slider-roughness 2005; FAJARDO and MAZO 2010). However, the
length scale. There is then a transition from a region observed response does not exhibit the behavior
of single lattice-spacing sliding to a region of associated with single contact friction at increased
multiple lattice-spacing sliding as the pulling velocity temperature, i.e., a shift of the transition to higher
increases. Close to 5 cm/s, the system transitions to speeds (TSHIPRUT et al. 2009). This is likely to be
oscillations. The oscillations observed are consistent because of the extended incommensurate nature of
with inertial oscillations of the spring–mass system, the contact.
and are similar to those reported earlier by Kudrolli In Fig. 10, we compare sliding at 5 cm/s for the
and coworkers (NASUNO et al. 1997a, b). In the three high-velocity cases studied: sliding on the solid
granular solid case with vibration, we expect a bed in the presence and absence of vibration, and
characteristic period of 0.236 s. This value matches sliding on the granular bed. The vibrated and
the data reasonably well for sliding at 5 cm/s, where granular substrates are very similar at comparable
the period of oscillations is 0.263 s (Fig. 7). At pulling speeds in the regime of inertial oscillations.
10 cm/s, the period of the oscillations is comparable, The solid lattice exhibits entirely distinct behavior in
namely 0.249 s. Videos of the system while being that it requires a higher sliding speed to transition
vibrated reveal an amplitude of vibration of the order into the inertial oscillatory behavior. We interpret
of tenths of millimeters, which adds approximately this as, on average, an effectively smaller amount of
8lJ of energy to the 0.17 kg slider. To place this energy needed to lift the slider over the grains in the
in context, this amount of energy, converted to case of the granular bed vs. the solid bed. For the
gravitational energy, would correspond to a height granular solid the addition of vibration provides
increase of 0.024 mm. An oscillatory amplitude of energy to the slider that effectively lowers the energy
0.2 mm at 11 Hz corresponds to a peak (horizontal) barrier.
Vol. 168, (2011) Transition to Steady Sliding in a 2D Granular Medium 2273

behavior. The values at which we observe the


crossover are in quantitative agreement with esti-
mates based on current molecular models.

Acknowledgements

This work was supported by the US Army Research


Office (grant W911NF-07-1-0131-00), by LANL
Subcontract Number: 64898-001-08 (with P. John-
son and C. Marone), and by NSF-DMR0906908 and
NSF-DMR0805204. We appreciate insightful discus-
sions with M.O. Robbins.

REFERENCES

AHARANOV, E. and SPARKS, D. (2003) Stick–slip in granular mate-


rial, J. Geophys. Res. 109, art #B09306
BAREL, I. and URBAKH, M. (2010), Multibond dynamics of Nano-
scale friction: the role of temperature Phys. Rev. Lett. 104, art #
066104
Figure 10 BAUMBERGER, T., BERTHOUD, P. and COROLI, C. (1999), Physical
Pulling force versus distance in steps (1 step = 0.1 mm) for a 170 g analysis of the state- and rate-dependent friction law. II.
slider and 160 N/m spring pulled at 5cm/s. This figure compares Dynamic friction Phys. Rev. B 60, 3928-3939
three cases: granular solid without vibration (top), granular solid BEHRINGER, R.P., CHAKRABORTY, D. BI, B., HENKES, S. and HARTLEY,
with 11 Hz vibration applied to the slider (middle); and granular R. (2008), Title: Why Do Granular Materials Stiffen with Shear
bed (bottom). The force curve for the vibrated granular solid and Rate? Test of Novel Stress-Based Statistics, Phys. Rev. Lett. 101,
the granular bed are very similar at 5 cm/s, in distinct contrast with art #268301
the unvibrated granular solid (see text) BEN-ZION, Y (2001), Dynamic ruptures in recent models of earth-
quake faults, J. Mechanics and Physics of Solids 49, 2209-2244
BEN-ZION, Y. (2008) Collective Behavior of Earthquakes and
5. Conclusions Faults: Continuum-Discrete Transisitions, Porgressive Evolu-
tionary Changes, and Different Dynamic Regimes Reviews of
Geophysics 46, RG406, 1-70
To summarize, we have performed an experi- BOROVSKY, B., KRIM, J., SYED Asif, S.A. and WAHL, K. (2001)
mental study of the stick–slip and stick–slip to steady Measuring Nanomechanical Properties of a Dynamic Contact
sliding behavior of a slider that is pulled through Using an Indenter Probe and Quartz Crystal Microbalance,
J. Appl. Phys. 90 6391-6396
a spring across 2D granular beds of photoelastic BOWDEN, F. P. and TABOR, D. (1954) , The Friction and Lubrication
disks. The disks are either fixed in a lattice or of Solids, Oxford University Press
unconstrained. We find more overlap than might be BRAIMAN, Y., HENTSCHEL, H.G.E., FAMILY, F., MAK, C., AND KRIM, J.
(1999), Tuning Friction with Noise and Disorder Phys. Rev. E
expected, despite the difference in scales, between
59(5), R4737-R4740.
molecular, granular and geophysical friction. To our BRAUN, O.M., PEYRARD, M., BORTOLANI, V., FRANCHINI, A. and
knowledge, this similarity has not been reported VANOSSI (2005),Transition from smooth sliding to stick–slip
previously in the experimental literature. Despite motion in a single frictional contact Phys. Rev. E 72, art #
056116
having minimal impact on mean effective friction BURRIDGE, R. and KNOPOFF, L. (1967), Model and Theoretical
coefficients, both external vibration and granular Seismicity Bull. Seismol. Soc. Am. 57, 3411
‘‘disorder‘‘ affect transitions at higher pulling speeds. CAPOZZA, R., VANOSSI, A., VEZZANI, A., AND ZAPPERI, S. (2009)
Suppression of Friction by Mechanical Vibrations, Phys. Rev.
Specifically, they result in a reduction in the cross- Lett. 103, art # 085502
over speed to steady sliding and an increase in the CARLSON, J. M., LANGER, J. S. and SHAW, B. E. (1994) Dynamics of
crossover from single to double jump stick–slip Earthquake faults, Rev. Mod. Phys. 66, 657-670
2274 J. Krim et al. Pure Appl. Geophys.

COULOMB, C.A. (1773) Memoires deMathematique & de Physique, KRIM, J and BERHINGER, R.P. (2009), Friction, force chains, and
presentes a l’Academie Royale des Sciences par divers Savans et falling fruit, Physics Today 62, 66-67
lus dan ses Assemblees, 7, 343-382 LOSERT, W., GEMINARD, J.-C., NASUNO, S. AND GOLLUB, J. P., (2000),
DANIELS, K.E. and BEHRINGER, R.P. (2005), Hysteresis and com- Mechanisms for slow strengthening in granular materials,
petition between disorder and crystallization in sheared and Phys.Rev. E 61 4060-4068
vibrated granular flow., Phys. Rev. Lett. 94, art# 168001 LUAN, B. and ROBBINS, M.O. (2004), Effect of inertia and elasticity
DANIELS, K. E. and HAYMAN, N.W. (2008), Force chains in on stick–slip motion, Phys. Rev. Lett. 93, art # 036105
seismogenic faults visualized with photoelastic granular shear MAK, C. and KRIM, J. (1997), Quartz crystal microbalance studies
experiments Journal of Geophysical Research 113, art # B11411 of disorder-induced lubrication, Faraday Discussions 107,
DAWSON, B.D., LEE, S.M. and KRIM, J. (2009), Tribo-Induced 389-397
Melting Transition at a Sliding Asperity Contact, Phys. Rev. Lett. MAKSE, H. A. and KURCHAN J. (2002) Testing the thermodynamic
103, art # 205502 approach to granular matter with a numerical model of a deci-
DE PAOLA, N., HIROSE, T., MITCHELL, T., DI TORO, G., VITI, C. and sive experiment Nature 415, 614-617
SHIMAMOTO, T. (2011) Fault lubrication and earthquake propa- MAJMUDAR, T. S. and BEHRINGER,R. P. (2005), Title: Contact force
gation in thermally unstable Rocks, Geology 39, 35-38 measurements and stress-induced anisotropy in granular mate-
DIETERICH, J. H. (1972), Time-Dependent Friction in Rocks, rials, Nature 435, 1079-1082
J. Geophysical Research 77, 3690 MAJMUDAR, T. S., SPERL, M., LUDING, S. and BEHRINGER, R. P.
EDWARDS, S.F. and OAKESHOTT, R.B.S. (1989), Theory of Powders, (2007), Jamming transition in granular systems, Phys.Rev. Lett.
Physica A 157, 1080-1090 98, art #058001
ELMER, F.-J. (1997), Nonlinear dynamics of dry friction, J. Phys. MARONE, C., RALEIGH, C. B. and SCHOLZ, C. H. (1990), Frictional
A. Math Gen. 30, 6057-6063 behavior and constitutive modeling of simulated fault gouge,
EYRING, H., (1936), Viscosity, plasticity, and diffusion as examples J. Geophysical Research 95, 7007-7025
of absolute reaction rates, J. Chem. Phys. 4, 283-291 MCFADDEN, C.F. and GELLMAN, A.J., (1997) Metallic friction: the
FAJARDO, O.Y. and MAZO, J. J (2010), Title: Effects of surface influence of atomic adsorbates at submonolayer coverages, Surf.
disorder and temperature on atomic friction, Phys. Rev. B 82, art Sci. 391, 287-299
#035435 MELOSH, H.J., (1979) Acoustic Fluidization-New Geologicprocess,
FILIPPOV, A.E., KLAFTNER, J., and URBAKH, M. (2004), Friction J. Geophysical Research 84, 7513-7520
through dynamical formation and rupture of molecular bonds, MUSER, M. H., URBAKH, M. and ROBBINS, M. O. (2003), Statistical
Phys. Rev. Lett. 92, art #135503 mechanics of static and low-velocity kinetic friction, Advances in
GENG, J., BEHRINGER, R. P., REYDELLET, G., and CLEMENT, E. (2003), Chemical Physics 126, 187-272
Green’s function measurements of force transmission in 2D NAKAMURA, J., WAKUNAMI, S. and NATORI, A. (2005) Double-slip
granular materials, Physica D 182, 274303 mechanism in atomic-scale friction: Tomlinson model at finite
GOLDHIRSCH, I., and ZANETTI, G. (1993), Clustering Instabilty in temperatures, Phys. Rev. B 72, art # 235415
Dissipative Gases, Phys. Rev. Lett. 70, 1619-1622 NASUNO, S., KUDROLLI, A. and GOLLUB, J. P., (1997a), Friction in
HARTLEY, R. R. and BEHRINGER, R. P. (2003), Logarithmic rate granular layers: Hysteresis and precursors, Phys. Rev. Lett. 79
dependence of force networks in sheared granular materials, 949-952
Nature 421, 928-930 NASUNO, S., KUDROLLI, A. and GOLLUB, J. P., (1997) Sensitive force
HENKES, S., O0 HERN, C.S. and CHAKRABORTY, B. (2007), Entropy and measurements in a sheared granular flow with simultaneous
temperature of a static granular assembly: An ab initio imaging, Powders & Grains 97,329-332, R. P. Behringer and
approach. Phys. Rev. Lett. 99, art# 038002 J. T. Jenkins, eds., Balkema, Rotterdam, 1997
HESLOT, F., BAUMBERGER, T., PERRIN, B., CAROLI, B. and CAROLI, NOWAK, E.R., KNIGHT, J.B., POVINELLI, M.L., JAEGER, H.M. and
C.(1994) Creep, Stick–slip, and Dry-Friction dynamics- Exper- NAGEL, S.R. (1997), Reversibility and irreversibility in the pack-
iments and a Heuristic Model,Phys. Rev. E 49, 4973-4988 ing of vibrated granular material Powder Technol. 94, 79-83
HOWELL, D.W., BEHRINGER, R.P., and VEJE, C.T. (1999), Fluctua- PANELLA, V., CHIARELLO, R. and KRIM, J. (1996), Adequacy of the
tions in granular media, Chaos 9, 559-572 Lifshitz theory for certain thin adsorbed films, Phys. Rev. Lett.
JANSEN, L., HOLSCHER, H., FUCHS, H. and SCHIRMEISEN, A. (2010) 76, 3606-3609
Temperature Dependence of Atomic-Scale Stick–slip Friction, ROBBINS. M.O., and KRIM, J. (1998), Energy Dissipation in inter-
Phys. Rev. Lett. 104, art # 256101 facial friction, MRS Bulletin 23, 23-25
JINESH, B., YU, S., KRYLOV, H., VALK, DIENWIEBEL, M. and FRENKEN, RUINA, A. (1983), Slip instability and state variable friction laws,
J.W.M. (2008), Thermolubricity in atomic-scale friction, Phys. Jour. Geophysical Research 88, 359-370
Rev. B 78 art # 155440 SIAVOSHI, S., ORPE,A. V. and KUDROLLI, A. (2006), Friction of a
JOHNSON, P.A. and JIA X. (2005), Nonlinear dynamics, granular slider on a granular layer: Nonmonotonic thickness dependence
media and dynamic earthquake triggering, Nature 437, 871- and effect of boundary conditions, Phys. Rev. E 73, art
874 #010301(R)
JOHNSON, P., SAVAGE, H., KNUTH, M., GOMBERG, J., and MARONE, C. SOCOLIUC, A., GNECCO, E., MAIER, S., PFEIFER, O., BARATOFF, A.,
(2008), Effects of acoustic waves on stick–slip in granular media BENNEWITZ,R. and MEYER, E. (2006), Atomic-scale control of
and implications for earthquakes, Nature 451, 47-60 friction by actuation of nanometer-sized contacts Science 313,
KRIM, J. (1996) The Atomic-scale Origins of Friction, Langmuir, 207-210
12, 4564-4566 TSAI, J.-C., VOTH, G. A., and GOLLUB, J.P. (2003), Internal Gran-
KRIM, J. (2002), Friction at Macroscopic and Microscopic Length ular Dynamics, Shear-Induced Crystallization, and Compaction
Scales, Amer. J. Phys. 70, 890-897 Steps, Phys. Rev. Lett 91, 064301
Vol. 168, (2011) Transition to Steady Sliding in a 2D Granular Medium 2275

TSHIPRUT, Z., ZELNER, S. and URBAKH, M. (2009) Temperature- YU, P., SHANNON, T., UTTER, B. and BEGRINGER, R.P., (2011) Stick–
Induced Enhancement of Nanoscale Friction, Phys. Rev. Lett slip in a 2D granular Medium, Preprint.
102, art # 136102 ZHANG, J., MAJMUDAR, T. S., TORDESILLAS, A. and BEHRINGER, R. P.
YU, PEIDONG and BEHRINGER, R.P. (2005), Granular Friction: (2010) Statistical properties of a 2D granular material subjected
A Slider Experiment Chaos 15 041102 to cyclic shear, Granular Matter. 12, 159-172

(Received September 2, 2010, revised June 8, 2011, accepted June 10, 2011, Published online August 9, 2011)

You might also like