Microscopic Mechanism of the Transition from Stick-slip Motion to Smooth Sliding: A Minimal Sliding Velocity

The smooth sliding regime is the desired one for most devices. But what is a minimal velocity  vc when the sliding remains smooth? In what follows we show that at the nanoscale,  vc  is of atomic-scale order, i.e.  vc ~ 1 m/s (for the stick-slip to smooth sliding transitions at the meso- and macro-scale, see )



Hysteresis goto top

As follows from simulation, the system always exhibits hysteresis provided the lubricant is not molten, T < Tm: the sliding starts at  f = fs, but stops when the driving force decreases below a threshold  f = fb< fs,  when the sliding velocity is v = vb= v(fb) > 0. Therefore, if now one attaches a spring to the top substrate and the driving velocity increases, one should observe the transition from stick-slip motion to smooth sliding similarly to that observed experimentally. In the simplest approach, if the free end of the spring moves with a velocity vspring, the motion will be smooth if vspring> v(fb),  while in the opposite case of  vspring< v(fb)  the stick-slip motion has to be observed. A real picture is more involved, because the shape of the function v(f) depends on the rate of  f  changing, and also  fs depends on the timelife of the stationary contact.

 

Soft lubricant goto top

The constant-force algorithm for the soft lubricant (Vll= 1/9) with  Nl=5:  vb~ 0.3

a36-modified.gif a09-modified.gif
  enlarged view

 

 

The spring algorithm for the soft lubricant (Vll=1/9) with  Nl=5  and  kspring= 3·104:  the stick-slip motion is due to the melting-freezing mechanism, and  0.03 < vcrit < 0.1

a21.gif draw01.gif
  enlarged view

 

 

Hard lubricant goto top

The constant-force algorithm for the hard lubricant (Vll=1, ideal structure of the lubricant film corresponded to the perfect sliding) with  Nl=1 or 5:  vb~ 0.2 to 0.3

fig010.gif a09-modified.gif
  enlarged view

 

 

The spring algorithm for the hard lubricant (Vll=1, "amorphous" structure of the lubricant) with  Nl=2 and 5,  and  kspring= 3·104:  the stick-slip motion is due to the inertia mechanism, and  0.1 < vcrit < 1

hard.gif draw01.gif
  enlarged view

 

 

In all these cases the transition from stick-slip to smooth sliding occurs at  v ~ atomic scale, i.e. v ~ 1 to 10 m/s, while experiment typically shows the transition at driving velocities as low as v ~ 1 μm/s (more than six orders lower!). Thompson and Robbins (1990) suggested that the large mass of the sliding block inhibits the freezing transition. However, the block as a whole will never stop abruptly, but only the bottom surface of the block stops (Persson 1994). Below we prove this statement rigorously.

 

The simplest model: one atom in sinusoidal substrate potential goto top

fig-a5h.gif
the rigid bottom substrate (external sinusoidal potential) +
one atom + the dc force is applied directly to the atom +
Langevin motion equation + underdamped external friction
fig-a5.gif
Figure: the velocity of the atom (red) versus time when the force (solid line) smoothly decreases

 

The sliding-to-locked threshold force  fb can be found by balancing the gain in energy due to the driving force and the energy loss due to dissipation. When the particle moves for the distance a (one period of the external potential), it gains the energy Egain= fa  and losses some energy Eloss,

Eloss =

τ

0 

dt ffric(t) v(t) =

τ

0 

dt Mηv2(t) = Mη

a

0 

dx v(x),

(*)

where τ is the "washboard period" (the time of motion for the distance a) and  ffric(t)=Mηv(t)  is the external frictional force that causes the energy losses. In the regime of steady motion, these energies must be equal each other, Egain= Eloss. Thus, the backward threshold force for the transition from the sliding (running) motion to the locked (pinned) state is determined by the equation  fb= min(Eloss)/a. The minimal losses are achieved when the particle has zero velocity on top of the total external potential  Vtot(x) = V(x) – fx.  In the limit  η → 0  when   f → 0,  analytics: From the energy conservation law,  (1/2)mv2 + (1/2)E [1 – cos(2πx/a)] = E,  we can find the particle velocity  v(x)  and then substitute it into Eq.(*); this yields
fb = Mη

a
æ
è
E

M
ö
ø
1/2

 

a

0 
dx é
ë
1 – cos æ
è
x

a
ö
ø
ù
û
1/2

 
= Cη(EM )1/2,
(**)
where  C ≡ (2π)–10 dy (1–cos y)1/2 = 2√2 /π ≈ 0.9. Equation (**) can be rewritten as  fb= Mηv = (2/π)Mηvm,  where  v = a–10a dx v(x),  and  vm= (2E/M)1/2 = πv/2  is the maximum velocity achieved by the particle when it moves at the bottom of the total external potential Vtot(x). The average particle velocity,  v = τ–10τ dt v(t) = a/τ,  tends to zero when  f fb,  because  τ → ∞  in this limit. Thus, this simple picture predicts that  fb <  fs= 1,  i.e. the hysteresis exists due to inertia mechanism, but  vb~ M 1/2  depends on the mass of the sliding object.

 

One-dimensional model goto top

Let the top substrate consists of  N  "layers", the first (bottom) layer moves in the external sinusoidal potential, and the dc force  F  is applied to the last (top) layer,

 

 

 ..  

x1

+ η

 .  

x1

+ η1 (

 .  

x1

 .  

x2

) + g (x1x2) + sin x1 = 0,

 

 

..

xl

+ ηl (2

 .

xl

 .

xl-1

 .     

xl+1

) + g (2xlxl–1xl+1) = 0,    l = 2,...,N–1,

 

 

..   

xN

+ ηN (

 .   

xN

.     

xN-1

) + g (xNxN–1) – F = 0.

 
model.gif
For the semi-infinite substrate we suppose that the damping  ηl  inside the block is zero at the interface and increases smoothly far from the interface,
ηl = ηm   hl h1

hN h1
,      hl = tanh æ
è
l Ld

ΔL
ö
ø
,      l = 1,...,N,
 
where Ld = 0.6N,  ΔL = N / 7  and  ηm= 10ωs. Thus, a wave emerged at the interface due to sliding, will propagate inside the substrate and will be damped there.

 

Simulation results goto top

lo-fig07.gif Figure: the transition from smooth sliding to the locked state at the force  f = 0.24  for the 1D top substrate of  N = 2048  "layers" for  g = 10  and  η = 0.1.

Inset: the velocities of some selected layers of the top substrate. One can see the wave emitted at the locking moment which propagates into the substrate

 

 

lo-fig03.gif

Figure 1D: v(f) for the 1D top substrate with g = 10 for three values of the external damping coefficient η = 0.03, 0.1 and 0.3. Open diamonds and dotted curves show the simulation results (N = 2048). One can see that for a semi-infinite substrate the sliding-to-locked transition is sharp.

Solid curves are for the approximate analytical results (dashed curves, for the unstable branches), dash-dotted lines describe the trivial contribution v = f/. One can see a good agreement of approximate analytics (only the first harmonic of ωwash has been included) with simulation

 

 

A finite slab goto top

When the top block corresponds to a thin slab (which may be glued to another large block so that a reflecting interface exists), then the wave emerged at the interface and propagated through the substrate, will be reflected from the top surface of the slab and go back, and a standing wave is excited. This wave does not allow the transition to the locked state, so the sliding state persists now for much smaller values of the dc force. This resonance effect depends on the width of the top block — the narrower is the slab, the larger is  vtop  and for smaller forces the sliding persists.
lo-fig08.gif lo-fig09.gif
Figure: the transition from sliding to the locked state as the force decreases with time for the 1D top substrate of N = 2048  layers with the constant damping ηl= 0.1 in the substrate (g = 10, η = 0.1) (enlarged view)

Figure: the averaged velocity of the top substrate versus the dc force for the 1D model with the constant damping ηl= 0.1 in the top substrate (g = 10, η = 0.1)

(enlarged view)

 

Analytics goto top

A general approach: let us present the total force as  F = F1 + F2 + F3,  where

F1  is the force because of the flux of energy into the top substrate,  F1 = Eloss /a,  while the sum

F2+F3 = mηv  is due to the external damping of the atoms in the lowest (contact) layer of the top substrate when it moves with respect to the bottom substrate, and consists of the "trivial" contribution F2 = mηv  and the "fluctuating" contribution

F3 =

mη


a

 

τ

0 

dt [v2(t) – v2] =

mη


v

1


τ

 

τ

0 

dt [v(t) – v]2.

 

Then we can use the following technique.

 

Green-function technique goto top

  1. The (top) substrate with internal (excitable) degrees of freedom is described by the causal (phonon) Green function  G(ω),  (ω2iωηl–D)G(ω) = 1,  where D is the elastic matrix of the semi-infinite (top) substrate and  ηl describes the damping inside the sliding block. Then the generalized susceptibility is  α(ω) = –G(ω)/m  (here ω must be real).
  2. When the top block moves with an average velocity  <v>,  its atoms in the lowest layer oscillate with the  washboard frequency  ω0 = (2π/a)<v> + higher harmonics.
  3. Linear response theory: when a system is excited by a small periodic force,  f(t) = Re f0 exp(iω0t)  (here  f0  is real), then its velocity will oscillate with the amplitude  v0=iω0x0,  where  x0=α(ω0)f0  (let  f0=fs).
  4. The rate of energy losses (the energy absorbed by the top block per one time unit) is
    R = 1

    2
    f02ω0 Im α(ω0) ,   or
     
    R = π

    4
      f02

    m
    ρ(ω0) ,
     
    where  ρ(ω)  is the density of phonon modes in the top substrate ( ∫0  ρ(ω) = 1 ),
    ρ(ω) = – 2

    π
    ω Im G(ω) .
     
  5. The energy absorbed by the top substrate during its motion for one period of the external potential is equal to Eloss(1) = =2πR/ω0=Ra /v,  and the contribution  F1  is  F1= Eloss(1)/a.
  6. If we take into account the lowest harmonic of the washboard frequency only,  v(t) = v+ |v0| cos(ω0t + φ),  then  |v0| = ω0|a(ω0)| f0,
    F1= π

    4
      f02

    mv
    ρ(ω0) ,  and
    (F1)
    F3= 1

    2
      mη

    v
    |v0|2 = η f02

    2m
      æ
    è
    2π

    a
    ö
    ø
    2

     
    |G(ω0)|2 v
     
    Note that for a single atom, the backward threshold force is determined by the F3 contribution only. However, when the moving block has internal degrees of freedom, then  F1> 0, and the backward velocity may be nonzero, so that  F2> 0  too.
  7. In a general case, one has to take into account all harmonics ωn= (n+1)ω0,  n = 0,...,∞,  and perform the self-consistent calculation.  Namely, for a given velocity v we have to start with some approximate shape of  x(t)  [e.g.,  x(t)=vt]  and then calculate step by step:
    1. the force acting on the atoms of the top block,
      f(t) = – sin x(t) – m η dx(t)/dt ,
       
    2. make its Fourier transform,
      fn =

      τ

      0 
      dt eiωnt f(t) ,
       
    3. find the velocity in response to this force,
      vn= –n fnG(ωn) ,
       
    4. then perform the backward Fourier transform,
      v(t) = (2π)–1ω0

      n=0 
      ent vn ,
       
    5. and calculate x(t) as the integral of v(t) over time; the output trajectory must coincide with the input one;

    6. finally, when self-consistency has being achieved, calculate

      F1 = –

      1


      2mv

       



      n=0 

      ωn |fn|2 Im G(ωn) = –

      π


      am

       



      n=0 

      (n+1) |fn|2 Im G(ωn)     and

       
      F3 = mη

      2v
       

      n=0 
      |vn|2 πmηv

      a
       

      n=0 
      (n+1)2 |fnG(ωn)|2.
       

One-dimensional model goto top

(Only the first harmonic is taken into account in what follows). Let the top substrate is modeled by the semi-infinite one-dimensional chain of atoms oriented perpendicularly to the bottom substrate as in the simulation presented above. The "substrate" atoms are coupled by harmonic springs with the elastic constant  g  and the lattice constant  a;  therefore the phonon spectrum of the substrate is  ω(k)=ωmsin(ak/2),  where  ωm=2√(g/m),  and the Green function is
Im G(ω) = – 2

ωm2
  æ
è
ωm2

ω2
– 1 ö
ø
1/2

 
 
(G)
inside the phonon zone,  |w| < ωm,  and   Im G(ω)=0  outside the zone,  |ω|ωm.  The real part of the Green function is constant inside the zone,  Re G(ω)=2/ωm2  for  |ω|ωm,  while outside the zone,  |ω| > ωm,  the real part is equal to
Re G(ω) = 2

ωm2
  é
ë
1– æ
è
x–1

x+1
ö
ø
1/2

 
ù
û
,
 
where  x = 2ω2/ωm2 – 1.  The "surface" density of phonon states is equal to
ρ(ω) = 4

πωm
  æ
è
1 – ω2

ωm2
ö
ø
1/2

 
,      |ω|ωm.
(rho)
The substitution of this expression into Eq.(F1) yields
F1 = 1

4
æ
è
2π

a
ö
ø
f02

g
é
ë
æ
è
a

2π
ö
ø
2

 
4g

mv2
 – 1 ù
û
1/2

 
(F1a)
provided the washboard frequency is inside the phonon spectrum,  ω0< ωm.
In the low velocity limit,  ‹v→ 0,  Eq.(F1a) leads to  F1 f02/v(mg)1/2Thus, the phonon contribution to the frictional force tends to infinity at  v→ 0,  because the density of phonon states, Eq.(rho), is nonzero at ω=0  for the one-dimensional system. From Eq.(G) we have  |v0| = 2f0/m  and
|G(ω0)|2 = 4

ωm2ω02
= 4

ωm2
æ
è
a

2π
ö
ø
2
 
 
1

v2
 ,
 
so that the contribution  F3  is equal to
F3 = ηf02

2gv
 .
 
Figure: the frictional force (solid curve and solid diamonds) as a function of the velocity of the 1D top substrate with  g = 10 (ωm≈ 6.32)  and  η = 0.1, the contribution  F1 due to radiation into the top substrate (dotted curve and open diamonds), the trivial contribution  F2= mηv (solid line), and the fluctuating contribution F3 (dashed curve and crosses). lo-fig06.gif

The function  f(v)  has a minimum at  v = vb  where  f(vb) = fb. Thus, if the dc force applied to the top substrate decreases, then the transition from the sliding regime to the locked state takes place at   f = fb  when the average velocity of the top block is nonzero, vb> 0.

The comparison of the analytical and simulation results was presented above in Fig.1D. One can see that the agreement is quite good (recall that in analytical approach we took into account the lowest harmonic of the washboard frequency only). Emphasize also that  the analytical approach corresponds to the constant-velocity algorithm of MD simulation, while the numerical results were obtained with the constant-force algorithm; both approaches lead to the same result.

 

 

 

Figure: (a) the threshold force  fb  and (b) the velocity  vb  for the sliding-to-locked transition as functions of the external damping  η  in the 1D model of the top substrate for two values of the substrate elastic constant  g = 10 (ωm≈ 6.32) and  g = 100 (ωm= 20).

 

The threshold values do not depend on the total mass of the sliding (top) block, but do depend on the elasticity of the block: a larger is  g, the lower are both threshold values  fb  and  vb.

lo-fig04.gif
Thus, the Green-function technique works well for the 1D model and, therefore, it may be applyed for the 3D model as well, where MD simulation (with few thousand of substrate layers) is problematic.

 

Three-dimensional model goto top

For the semi-infinite 3D substrate the Green function can be approximated by (Braun 1989)

Im G(ω) = –

16


ωm6

ω (ωm2ω2)3/2 = –

16


ωm2

 ξ (1 ξ2)3/2,

 

where  ξ = ω/ωm|ξ| ≤ 1. The real part of Green function may be found with the help of Kramers-Kronig relation,

Re G(ω) =

2


π



0 

1

ω1


ω12ω2

 Im G(ω1) =

32


πωm2

1

0 

dy

y2 (1– y2)3/2


(ω /ωm)2y2

 .

 

The "surface" density of phonon states is equal to

ρ(ω) =

32


πωm6

ω2(ωm2 ω2)3/2.

 

The substitution of this expression into Eq.(F1) yields

F1 =

8 f02


m3

æ
è

2π


a

ö
ø

2

 

æ
è

1 –

ω02


ωm2

ö
ø

3/2

 

v.

 

In the limit  v→ 0  we obtain

|v0| =

6 f0


m2

 

æ
è

2π


a

ö
ø

v      and       F3 =

18 η f02


m4

æ
è

2π


a

ö
ø

2

 

v.

 

Thus, in the approximation when only the main harmonic of the washboard frequency is taken into account, in the limit  v→ 0  we obtain  F1=(8/ωm3)v,  F2=ηv  and  F3=(18η/ωm4)v,  which all together gives  Fb= 0  and  vb= 0.  However at low velocities of the top block, the motion of the atoms is highly anharmonic, and the higher-order harmonics of the washboard frequency must be taken into account. The self-consistent calculation leads to the results presented in the figures below.

lo-fig10.gif

lo-fig11.gif

The dependence F(v) and its contributions for the 3D model of the top substrate (enlarged view). The dependence F(v) for the 3D model for different model parameters (enlarged view).

Conclusion goto top

In all cases, the microscopic transition from stick-slip motion to smooth sliding occurs at  v ~ atomic scale,  i.e. v ~10 m/s, while “macroscopic” experiments show the transition at spring velocities as low as  v ~1 μm/s  (more than six orders lower). A large mass of the sliding block does not inhibit the freezing transition, because the block as a whole never stop abruptly – only the bottom surface of the block stops. Thus, nothing in common: the experimentally observed macroscopic smooth sliding corresponds in fact to the microscopic (atomic-scale) stick-slip motion, while the experimental (macroscopic) stick–slip motion is due to the fs(t) dependence (aging of contacts). The corresponding theory is similar to that used in earthquakes-like models.

 

Next: Earthquake-like model

goto main Back to main page or tribology page

 


Last updated on September 30, 2008 by Oleg Braun.             Translated from LATEX by TTH