You are on page 1of 28

Capstan Equation Generalised for Noncircular

Geometries
arXiv:2012.13353v3 [physics.class-ph] 19 Jan 2021

Kavinda Jayawardana∗

Abstract
In our analysis, we extend the capstan equation to noncircular geometries.
We derive a closed from solution for a membrane with a zero-Poisson’s ratio
(or a string with an arbitrary Poisson’s ratio) supported by a rigid prism,
and then a rigid cone, at limiting-equilibrium. As a comparison, we extend
Kikuchi and Oden’s model for Coulomb’s law of static friction [1] to curvi-
linear coordinates. We also conduct numerical experiments to see how close
Coulomb’s law of static friction is to the ordinary friction law implied by
our generalised capstan equation, in curvilinear coordinates. Our numerical
results indicate that increasing the curvature, the Poisson’s ratio and the
thickness of the elastic body increases the frictional force, for a constant co-
efficient of friction, and incompressible materials such as rubber can have a
high frictional forces, even under low coefficients of friction. Our analysis
implies that the coefficient of friction is model dependent.
Keywords: Capstan Equation, Coefficient Friction, Contact Mechanics,
Coulomb’s Law of Static Friction, Curvilinear Coordinates, Mathematical
Elasticity

1. Introduction
The capstan equation, otherwise known as Euler’s equation of tension
transmission, is the relationship that governs the maximum applied-tension,
Tmax , with respect to the minimum applied-tension, T0 , of an elastic string
(i.e. one-dimensional elastic body), wound around a rough (i.e. exhibiting
friction) cylinder [2, 3]. Thus, the governing equation is given by the following


Corresponding author
Email address: zcahe58@ucl.ac.uk (Kavinda Jayawardana)

January 20, 2021


equation,

Tmax = T0 exp(µF θ) , (1)

where θ is the contact angle and µF is the coefficient of friction: note that
the coefficient of friction is the physical ratio of the magnitude of the shear
force and the normal force between two contacting bodies. The capstan
equation is the most perfect example of a belt-friction model, which describes
behaviour of a belt-like object moving over a rigid obstacle subjected to fric-
tion [4], and due to its simplicity, it is widely used to analyse the tension
transmission behaviour of cable-like bodies in contact with circular profiled
surfaces, such as in rope rescue systems, marine cable applications, computer
storage devices (electro-optical tracking systems), clutch or brake systems in
vehicles, belt-pulley machine systems and fibre-reinforced composites [5].

One of the largest applications of the capstan equation can be found in the
filed of electronic cable drive systems, and such devices include printers, pho-
tocopiers and tape recorders [6]. The most recent applications can be found
in the field of robotics, as cable drive systems are fundamental in the design
and manufacture of high-speed pick-and-place robots (DeltaBot, BetaBot
and DashBot), wearable robot-assisted rehabilitation-devices (robotic pros-
thetics) [7], biologically-inspired humanoid-robots (ASIMO by Honda [8]),
and haptic devices [9–12]. Further applications for the capstan equation can
be found in the field of textiles, where tensioned fibers, yarns, or fabrics are
frequently in contact with cylindrical bodies [13–15].

There exist cases in the literature where authors investigated the belt-
transm-ission properties of noncircular pulley systems without the considera-
tion of frictional interactions [16–20]. However, should one include friction for
such cases then the ordinary capstan equation (1) or any generalised capstan
equations present in the literature (e.g. models by Baser and Konukseven [6],
Lu and Fan [21], Doonmez and Marmarali [22], Wei and Chen [23], Jung et
al. [5, 13, 24], Kim et al. [25], and Stuart [26]), they will fail to work as these
models assume that the rigid obstacle has a circular profile. Thus, to nu-
merically model such problems, one can possibly use Konyukhov’s [27] and
Konyukhov’s and Izi’s [28] model for ropes and orthotropic rough surface,
where the authors use the generalised orthotropic friction law by consider-
ing the coefficient of friction in pulling and dragging directions separately of

2
a rope. As an alternative approach, in our reading, generalise the capstan
equation to be valid for rigid noncircular obstacles with zero-Gaussian curva-
ture while still adhering to the standard friction law [3] in a static-equilibrium
(or steady-equilibrium) dry-friction setting [1, 29].

2. Derivation
Consider an elastic body on a rough rigid surface that is subjected to
external loadings and boundary conditions such that the body is at limiting-
equilibrium [2]. The governing equation for friction at the contact region can
be expressed as follows,

F = µF R , (2)

where R is the normal reaction force and F is the frictional force experi-
enced on the body, and µF is the coefficient of friction between the rough
rigid surface and the body at the contact region [2, 3]. Equation (2) is the
mathematical interpretation of Amontons’ three laws of friction, and in the
literature it is commonly referred to as the friction law [3].

To investigate the behaviour of an elastic body at the boundary, one


must know its stress at the boundary: note where stress is analogous to the
boundary force divided by its contact area. Now, let F α be the frictional
forces and F 3 be the normal reaction force at the contact region, where the
contact region described by the surface x3 = 0 in curvilinear coordinates
(x1 , x2 , x3 ) ∈ R3 , where Rk is the kth-dimensional curvilinear space. Then,
the friction law (2) can be expressed in the curvilinear space as Fα F α =
µ2F F3 F 3 . Dividing this equation by the contact area twice is analogous to
τ3α τα3 = µ2F τ33 τ33 , where τ3j are normal stresses at the contact surface. If one
assumes that one is dealing with a very thin structure (i.e. a true-membrane:
a two-dimensional elastic body), then the notion of normal stress becomes
meaningless (for more on true-membranes, we refer the reader to chapter 7 of
Libai and Simmonds [30]). However, note that the stress tensor is constant
throughout the normal direction for such thin objects, and thus, dividing
our stress based friction equation further by the thickness of the object twice
enable us to obtain a relation that is analogous in terms of force densities,
i.e. fα f α = µ2F f3 f 3 . This leads to our hypothesis:

3
Hypothesis 1. For a true-membrane over a rigid foundation, whose contact
area is described by (x1 , x2 , 0), at limiting-equilibrium, the force density field
at the contact region is governed by the following equation,
p p
frα frα = µF fr3 fr3 , (3)

where frα are the frictional force densities and fr3 is the normal reaction
force density of the true-membrane in the curvilinear space, and µF is the
coefficient of friction at the contact region.

Note that Einstein’s summation notation (see section 1.2 of Kay [31]) is
assumed throughout, bold symbols signify that we are dealing with vector
and tensor fields, and we regard the indices i, j, k, l ∈ {1, 2, 3} and α, β, γ, δ ∈
{1, 2}. Also, note that we usually reserve the vector brackets ( · )E for vectors
in the Euclidean space and ( · ) for vectors in the curvilinear space.

2.1. General Prism Case


Now, consider a general prism parametrised by the following map,

σ(x1 , x2 ) = (x1 , u(x2 ), v(x2 ))E , ∀ (x1 , x2 ) ∈ ω ,

where ω ⊂ R2 and u(·) and v(·) are C 1 (ω) 2π-periodic functions, C k (·) is
a space of continuous functions that has continuous first k partial deriva-
tives in the underlying domain. Note that ω is a simply-connected bounded
two-dimensional domain with a positively-oriented piecewise-smooth closed
boundary ∂ω such that σ forms an injection. The prism’s first fundamental
form tensor can be expressed as F [I] = diag(1, (u′ )2 + (v ′ )2 ) and the only
non-zero component of its second fundamental form tensor can be expressed
as follows,
(v ′ u′′ − u′ v ′′ )
F[II]22 = − q ,
′ 2 ′ 2
(u ) + (v )

where the prims represent derivatives. Note that this is a surface with a
zero-Gaussian curvature, i.e. det(F [II] ) = 0 where det(·) is the determinate
of a matrix operator.

Now, consider a rectangular membrane with a zero-Poisson’s ratio (or a


string with an arbitrary Poisson’s ratio), with a thickness h, that is in contact

4
with the prism and at limiting-equilibrium such that the boundary of this
membrane can be described as follows,
∂ω = ∂ωf ∪ ∂ωT0 ∪ ∂ωTmax ,
where
∂ωf = {(x1 , x2 ) | x1 ∈ {0, l} and θ0 < x2 < θmax } ,
∂ωT0 = {(x1 , x2 ) | 0 ≤ x1 ≤ l and x2 = θ0 } ,
∂ωTmax = {(x1 , x2 ) | 0 ≤ x1 ≤ l and x2 = θmax } .
Now, assert that the limits of x2 , is chosen such that F[II]22 < 0 in ω, i.e. the
contact region is a surface with a positive mean-curvature.

Consider the diffeomorphism Θ(x1 , x2 , x3 ) = σ(x1 , x2 )+x3 N (x1 , x2 ) with


respect to the map σ, where
∂1 σ × ∂2 σ
N=
||∂1 σ × ∂2 σ||
is the unit normal to the surface σ, ∂α are the partial derivatives with re-
spect to the curvilinear coordinates xα , × is the Euclidean cross product
and || · || is the Euclidean norm and x3 ∈ (−ε, ε) for some ε > 0 (theorem
4.1-1 of Ciarlet [32]). Now, with respect to the diffeomorphism Θ, the three-
dimensional Cauchy’s momentum equation in the curvilinear space can be
expressed as ∇ ¯ i T i + fj = 0, where T is the Cauchy’s stress tensor of the
j
true-membrane, is f is a force density field (see section 1 subsection 3.4 of
Morassi and Paroni [33]) and ∇ ¯ is the covariant derivative with respect to
the curvilinear coordinate system (x1 , x2 , x3 ) ∈ R3 (see section 6.4 of Kay
[31]). By definition, f is the sum of all the force densities, and thus, one can
re-express it as f = f r + g̃ r where g̃ r is some external loading (e.g. effects
due to gravity or centripetal force due to steady-equilibrium case), and g̃rj are
Lipschitz continuous (definition (iii) of appendix A section 3 of Evans [34]).

Now, assert that f 1 = 0 and the membrane is subjected to the boundary


conditions
Tβ1 ∂ω = 0 ,

T0
T22 ∂ω =

, (4)
T0 hl
Tmax (µF , σ(ω), g̃r , T0 )
T22 ∂ω

= , (5)
T max hl

5
where θ0 < θmax , and T0 and Tmax (µF , σ(ω), g̃r , T0 ) are forces applied at the
boundary (minimum and maximum applied-tension respectively) such that
T0 < Tmax (µF , σ(ω), g̃r , T0 ). Note that maximum applied tension is not arbi-
trary, as, for the membrane to remain at limiting-equilibrium, the maximum
applied tension must have a very specific value, which depends on the mini-
mum applied tension, the contact angle, the curvature, external loadings and
the coefficient of friction.

Due to conditions, which include the zero-Gaussian curvature, the zero-


Poisson’s ratio (i.e. ν = 0) and f 1 = 0, and the construction of the boundary
conditions (i.e. x1 independence), we find that the only non-zero component
of the stress tensor is T22 = T22 (x2 ). Thus, the Cauchy’s momentum equation
at x3 = 0 reduce to the following form,
∂2 T22 + fr2 + g̃r2 = 0 ,
F[II]22 T22 + fr3 + g̃r3 = 0 .
As friction opposes potential motion, i.e. fr2 < 0 (fr2 is decreasing as x2 is
increases for our case), and the normal reaction force is positive, i.e. fr3 > 0
(as we are considering a unit outward normal to the surface), hypothesis 1
implies that
1 1
∂2 T22 + µF (F[I]22 ) 2 F[II]22 T22 + g̃r2 + µF (F[I]22 ) 2 g̃r3 = 0 . (6)
Finally, integrate equation (6) with respect to boundary condition (4)
and multiply the resulting solution by lh to arrive at our first theorem:
Theorem 1. The tension T (·) of a membrane with a zero-Poisson’s ratio
(or a string with an arbitrary Poisson’s ratio) on a prism parametrised by
the map (x1 , u(x2 ), v(x2 ))E , subjected to an external force field (0, gr2 , gr3 )
in the curvilinear space, at limiting-equilibrium is
  ′ 
v (θ)
T (θ) = exp −µF arctan
u′(θ)
 Z θ   ′  
1
 v
× C− gr2 + µF (F[I]22 ) 2 gr3 exp µF arctan dx2 ,
θ0 u ′

where C = T0 exp(µF arctan(v ′/u′ ))|x2 =θ0 , T0 is the minimum applied-tension


at x2 = θ0 , grj are Lipschitz continuous, u(·) and v(·) are C 1 ([θ0 , θmax ]) 2π-
periodic functions and the interval [θ0 , θmax ] is chosen such that v ′ u′′ − u′ v ′′ >
0, ∀ x2 ∈ [θ0 , θmax ].

6
Proof. Please see above for the derivation. Note that × is the scaler multi-
plication and in this context.
Note that theorem 1 is not valid when u′ or v ′ are zero between the
limits θ0 and θmax . However, one can still evaluate such problems with some
thought as arctan(x) remains finite in the limit x → ∞.

Corollary 1. The tensile stress τ (·) of an infinitely long membrane with an


arbitrary Poisson’s ratio on a prism parametrised by the map (x1 , u(x2 ), v(x2 ))E
where |x1 | ≤ ∞, subjected to an external stress field (0, ḡr2 , ḡr3) in the curvi-
linear space, at limiting-equilibrium is
  ′ 
v (θ)
τ (θ) = exp −µF arctan
u′ (θ)
 Z θ   ′  
1
 v
× C− ḡr2 + µF (F[I]22 ) 2 ḡr3 exp µF arctan ′
dx2 ,
θ0 u

where C = τ0 exp (µF arctan(v ′ /u′ ))|x2 =θ0 , τ0 is the minimum applied tensile
stress applied at x2 = θ0 , ḡrj are Lipschitz continuous, u(·) and v(·) are
C 1 ([θ0 , θmax ]) 2π-periodic functions and the interval [θ0 , θmax ] is chosen such
that v ′ u′′ − u′ v ′′ > 0, ∀ x2 ∈ [θ0 , θmax ].

Proof. Simple case of noting that the solution in theorem 1 is invariant in


the x1 direction for all Poisson’s ratios given that the membrane is infinitely
long in the x1 direction.

2.2. General Cone Case


Now, consider a general cone parametrised by the following map,

σ(x1 , x2 ) = (x1 , x1 ū(x2 ), x1 v̄(x2 ))E , ∀ (x1 , x2 ) ∈ ω ,

where ω ⊂ R>0 × R, and ū(·) and v̄(·) are C 2 (ω) 2π-periodic functions.
Note that ω is a simply-connected bounded two-dimension domain with a
positively-oriented piecewise-smooth closed boundary ∂ω such that σ forms
an injection. Thus, cone’s first fundamental form can be expressed as follows,

1 + ū2 + v̄ 2 x1 ūū′ + x1 v̄v̄ ′


 
F [I] = 2 2 .
x1 ūū′ + x1 v̄v̄ ′ (x1 ū′ ) + (x1 v̄ ′ )

7
Also, only nonzero component of cone’s second fundamental form can be
expressed as follows,
(v̄ ′ ū′′ − ū′v̄ ′′ )
F[II]22 = −x1 q .
(ū′)2 + (v̄ ′ )2 + (v̄ū′ − ūv̄ ′ )2

Due to the non-diagonal nature of the first fundamental form of the cone,
it is difficult to find a simple friction law as we did with the prism case. But
note that this is a surface with a zero-Gaussian curvature, i.e. det(F [II] ) = 0.
Thus, Gauss’ Theorema Egregium (theorem 10.2.1 of Pressley [35]) implies
that there exists a map ϕ : (χ1 , χ2 ) 7→ ω such that the first fundamental form
tensor with respect to the isometry σ ◦ ϕ is the 2 × 2 identity matrix. With
some calculations, we can define the properties of this map ϕ as follows,

χ1 = x1 r̄ cos(φ) ,
χ2 = x1 r̄ sin(φ) ,

where

r̄ = 1 + ū2 + v̄ 2 ,
q
Z x2 (ū′ )2 + (v̄ ′ )2 + (v̄ū′ − ūv̄ ′ )2
φ(x2 ) = dθ .
0 r̄ 2
Also,

r̄ cos(φ) x1 r̄ ′ cos(φ) − x1 r̄φ′ sin(φ)


 
J= .
r̄ sin(φ) x1 r̄ ′ sin(φ) + x1 r̄φ′ cos(φ)

is the Jacobian matrix of the map ϕ. Furthermore, by the construction of


ϕ implies that det(J) > 0. With further calculations, one finds that the
first fundamental form tensor of the cone with respect to the isometry σ ◦ ϕ
ϕ
is F[I]αβ = δαβ and the second fundamental form tensor is can be expressed
follows,
(1 + ū2 + v̄ 2 ) (v̄ ′ ū′′ − ū′ v̄ ′′ ) sin2 (φ)
 
ϕ − sin(φ) cos(φ)
F [II] = − .
 3 − sin(φ) cos(φ) cos2 (φ)
x1 (ū′ )2 + (v̄ ′ )2 + (v̄ū′ − ūv̄ ′ )2 2

Now, consider an isosceles-trapezium membrane with a zero-Poisson’s


ratio (or a string with an arbitrary Poisson’s ratio), with a thickness h and

8
a length l separating its parallel sides that is in contact with the cone such
that it is at limiting-equilibrium. Thus, the boundary of this membrane has
the following form,
∂ω = ∂ωf ∪ ∂ωT0 ∪ ∂ωTmax ,
where
∂ωf = {(χ1 , x2 ) | χ1 ∈ {d, d + l} and θ0 < x2 < θmax } ,
∂ωT0 = {(χ1 , x2 ) | d ≤ χ1 ≤ d + l and x2 = θ0 } ,
∂ωTmax = {(χ1 , x2 ) | d ≤ χ1 ≤ d + l and x2 = θmax } ,
and d is the distance between the membrane and the apex of the cone at
x2 = 0. Note that d must always be a positive constant, and, just as it is for
the prism case, assert that the limits of x2 , i.e. limits of the contact interval,
ϕ2
is chosen such that F[II]2 < 0 in ω, i.e. the contact area is surface with a
positive mean-curvature.

Now, consider the diffeomorphism


Θ(χ1 , χ2 , x3 ) = σ ◦ ϕ(χ1 , χ2 ) + x3 N ϕ (χ1 , χ2 )
with respect to the map σ ◦ ϕ, where N ϕ is the unit outward normal to the
surface and x3 ∈ (−ε, ε), for some ε > 0. Note that N ϕ (χ1 , χ2 ) = N (x1 , x2 ),
i.e. unit normal to the surface is unchanged under the mapping σ ◦ ϕ, and
thus, the normal reaction force density remains unchanged under the new
coordinate system. Now, with respect to the diffeomorphism Θ, the three-
dimensional Cauchy’s momentum equation in the curvilinear space can be
expressed as ∇ ¯ i T i + fj = 0 where is f is a force density field. Unfortunately,
j
due to the geometry of the cone, we cannot impose a simple physically-
realistic external-loading as we did with the prism case. Thus, we omit the
external loading field g r form the calculations, i.e. now we have f j = frj .

Now, assert that f 1 = 0 and the membrane is subjected to the boundary


conditions
Tβ1 ∂ω = 0 , ∀ β ∈ {1, 2} ,

T0
cos(φ)T22 ∂ω =

, (7)
T0 hl
Tmax (µF , σ(ω), T0 )
cos(φ)T22 ∂ω

= ,
T max hl

9
where θ0 < θmax , and T0 and Tmax (µF , σ(ω), T0 ) are forces applied at the
boundary such that T0 < Tmax (µF , σ(ω), T0 ). Comments that we made re-
garding Tmax (µF , σ(ω), T0 ) for the prim case applies to this case also, with
the exception of the external loadings.

Due, to conditions which include the zero-Gaussian curvature, the zero-


Poisson’s ratio and f 1 = 0, and the construction of the boundary conditions,
we find that the only nonzero component of the stress tensor is T22 = T22 (χ2 ).
Thus, the Cauchy’s momentum equation at x3 = 0 reduces to following form,
∂2 T22 + fr2 = 0 ,
ϕ2 2
F[II]2 T2 + fr3 = 0 .

As friction opposes potential motion, i.e. fr2 < 0 (fr2 is decreasing as χ2 is


increases for our case), and the normal reaction force is positive, i.e. fr3 > 0
(as we are considering a unit outward normal to the surface), hypothesis 1
implies that
ϕ2 2
∂2 T22 + µF F[II]2 T2 = 0 . (8)

Despite the fact that equation (8) provides one with a simple relation for
friction, it is near impossible to integrate with respect to χ1 . But notice that
χ1 is related to χ2 by the following equation,
χ2 = χ1 tan(φ) . (9)
Thus, in accordance with the Fubini’s theorem (theorem 3-10 of Spivak [36]),
one may keep χ1 fixed and take the differential of equation (9) to find the
following,
q
2 1
(ū′ )2 + (v̄ ′ )2 + (v̄ū′ − ūv̄ ′ )2
dχ = χ sec2 (φ) dx2 . (10)
1 + ū2 + v̄ 2
Now, with the use of equation (10), one can express equation (8) purely in
terms of x2 as follows,

∂ 2 1 + ū2 + v̄ 2 (v̄ ′ ū′′ − ū′ v̄ ′′ )
log(T 2 ) − µ F cos(φ) = 0 . (11)
∂x2 (ū′ )2 + (v̄ ′ )2 + (v̄ū′ − ūv̄ ′ )2
Finally, integrate equation (11) with respect to boundary condition (7) and
multiply the result by hl cos(φ(θ0 )) to arrive at our second theorem:

10
Theorem 2. The tension T (·) of a membrane with a zero-Poisson’s ratio
(or a string with an arbitrary Poisson’s ratio) on a cone parametrised by the
map (x1 , x1 ū(x2 ), x1 v̄(x2 ))E at limiting-equilibrium is
Z θ√ !
1 + ū2 + v̄ 2 (v̄ ′ ū′′ − ū′v̄ ′′ ) 2
T (θ) = T0 exp µF 2 2 2 cos(φ) dx ,
′ ′ ′ ′
θ0 (ū ) + (v̄ ) + (v̄ū − ūv̄ )

where
q
Z x2 (ū′ )2 + (v̄ ′ )2 + (v̄ū′ − ūv̄ ′ )2
φ(x2 ) = dθ ,
0 1 + ū2 + v̄ 2

and where T0 is the minimum applied-tension at x2 = θ0 , ū(·) and v̄(·) are


C 2 ([θ0 , θmax ]) 2π-periodic functions, and the interval [θ0 , θmax ] is chosen such
that v̄ ′ ū′′ − ū′ v̄ ′′ > 0, ∀ x2 ∈ [θ0 , θmax ].
Proof. Please see above for the derivation.
Note that, if one expresses the map of the cone as follows,

x1 ϑ(x2 ) = (x1 , x1 ū(x2 ), x1 v̄(x2 ))E ,

then one can alternatively express the above theorem as:


Corollary 2. The tension T (·) of a membrane with a zero-Poisson’s ratio
(or a string with an arbitrary Poisson’s ratio) on a cone parametrised by the
map x1 ϑ(x2 ), where ϑ′′ · (ϑ′ × ϑ) > 0, ∀ x2 ∈ [θ0 , θ], at limiting-equilibrium
is
 Z θ
ϑ′′ · (ϑ′ × ϑ)

2
T (θ) = T0 exp µF ||ϑ|| cos(φ) dx ,
θ0 ||ϑ′ × ϑ||2

where
x2
||ϑ′ × ϑ||
Z
2
φ(x ) = dθ ,
0 ||ϑ||2

and where T0 is the minimum applied-tension at x2 = θ0 .


Proof. Let ϑ(x2 ) = (1, ū(x2 ), v̄(x2 ))E , then the result follows from theorem
2.

11
2.3. Explicit Solutions
Consider a cylinder with a radius a. This cylinder can easily be parametrised
by the map σ(x1 , θ) = (x1 , a sin(θ), a cos(θ))E . Now, consider a membrane
with a zero-Poisson’s ratio (or a string with an arbitrary Poisson’s ratio) over
the cylinder at limiting-equilibrium. Given that we are applying a minimum
tension T0 at θ0 = 0 and the membrane is not subject to an external loading,
then in accordance with theorem 1, we find tension on the membrane to be
of the following form,
T (θ) = T0 exp(µF θ) , (12)
which is just the ordinary capstan equation (1). Now, if one assumes that the
membrane (or the string) is subjected to the force of gravity g = (0, 0, −g)E ,
where g is the acceleration due to gravity, then we find that the covariant
force density due to gravity in the curvilinear space with respect to the map
σ to be g r = (0, a̺g sin(θ), − ̺g cos(θ)), where θ is the acute angle that
the vector (0, 0, 1)E makes with the vector (0, a, 0). Thus, theorem 1 implies
that the tension observed on the membrane has the following form,
1 − µ2F
 
Tg (θ) = T0 − ahl̺g exp(µF θ)
1 + µ2F
1 − µ2F
 
2µF
+ ahl̺g cos(θ) + sin(θ) . (13)
1 + µ2F 1 + µ2F
Note that the properties of equation (13) in the steady-equilibrium case is
extensively analysed in chapter 6 of Jayawardana [37] and most notable re-
sults can be found in Jayawardana et al. [38].

For an elliptical-prism case, consider a prism with a horizontal diameter


of 2a and the vertical diameter of 2b. This prism can easily be parametrised
by the map σ(x1 , θ) = (x1 , a sin(θ), b cos(θ))E , where θ is the acute angle
that the vector (0, 0, 1)E makes with the vector (0, ϕ(θ), 0), and where ϕ(θ) =
1
(b2 sin2 (θ) + a2 cos2 (θ)) 2 . Now, consider a membrane with a zero-Poisson’s
ratio over the cylinder at limiting-equilibrium. Given that we are applying a
minimum tension T0 at θ0 = 0 and the membrane is not subject to an external
loading, theorem 1 implies that the tension observed on the membrane has
the following form,
  
b
Telliptical (θ) = T0 exp µF arctan tan(θ) . (14)
a

12
Note that θ must not exceed the value 21 π, as at 12 π, as tan(·) is singular.
However, this is still not a problem as arctan((b/a) tan(·)) remains finite at
1
2
π, i.e. arctan x → ± 21 π as x → ±∞. For example, assume that the contact
angle is 12 π + α where 0 < α < 12 π. Thus, by considering the finiteness of
arctan(·) and considering elementary trigonometric identities, one finds that
the solution to this problem has the following from,
     
1 1 b
Telliptical π + α = T0 exp µF π + arctan tan(α) .
2 2 a

2.8

2.6

2.4

2.2
δτ

1.8

1.6

1.4
0.5 1 1.5
δb

Figure 1: Tension ratio against δb.

On a different note, equation (14) implies that the maximum applied-


tension, Tmax , is dependent on the mean-curvature of the rigid prism. To
investigate this matter further, consider a membrane (infinitely long in x1
direction or otherwise) on a rough elliptical prism, at limiting-equilibrium,
between the contact angles θ0 = − 14 π and θmax = 14 π. Thus, theorem 1 (and
corollary 1) implies that
  
b
δτ = exp 2µF arctan , (15)
a

13
where δτ = τmax /τ0 = Tmax /T0 , which we call the tension ratio. As the reader
can see that for a fixed contact interval and for a fixed coefficient friction,
equation (15) implies a non-constant tension ratio, δτ , for varying δb, where
δb = b/a. As the mean-curvature of the prism is H(θ) = 21 ab(ϕ(θ))−3 , one
can see that the tension ratio is related to the mean-curvature by
" #!
δτ = exp 2µF arctan max (2aH(θ), 1) + min (2aH(θ), 1) − 1 .
θ∈[− 41 π, 41 π] θ∈[− 14 π, 41 π]

We plot figure 1 to visualise the tension ratio, against δb, which is calcu-
lated with µF = 21 and δb ∈ [ 21 , 23 ]. The figure shows that, for a fixed contact
interval, as δb increases (i.e. as the mean-curvature of the contact region
increases), the tension ratio also increases. This is an intuitive result as the
curvature of the contact region increases, the normal reaction force on the
membrane also increases, which in turn leads to a higher frictional force, and
thus, a higher tension ratio. Now, this is a fascinating result as this effect
cannot be observed with the ordinary capstan equation.

On a final note, to find the tension observed on a membrane over a right-


circular cone with a 2α-aperture (i.e. angle between two generatrix lines),
consider the map σ(x1 , θ) = (x1 , x1 tan(α) sin(θ), x1 tan(α) cos(θ))E . Thus,
in accordance with theorem 2, we find tension on the membrane can be
expressed as follows,
Tcone (θ) = T0 exp (µF cot[α] sin (sin[α]θ)) .

3. Comparison Against Kikuchi and Oden’s Work


The most comprehensive mathematical study of friction that we are aware
of is the publication by Kikuchi and Oden [1], where the authors present a
comprehensive analysis of the Signorini’s problem, Coulomb’s law of static
friction and non-classical friction laws. The work includes meticulous docu-
mentation of the existence, the uniqueness and the regularity results for the
given mathematical problems, which also includes finite-element modelling
techniques. There, the authors present regularised Coulomb’s law of static
friction as follows,
 uT
νF σn
 , if |uT | ≥ ε ,
|uT |
σT (u) = (16)
 νF σn uT , if |uT | < ε ,

ε
14
where νF is the (Coulomb’s) coefficient of friction, σ T (u) is the tangential
stress tensor, σn < 0 is the normal stress, u is the displacement field and
uT is the tangential displacement field at the contact boundary, ε > 0 is
the regularisation parameter, and where this friction law is considered to be
an alternative mathematical representation of the Amontons’ three laws of
friction. In this section, we extend Kikuchi and Oden’s model for Coulomb’s
law of static friction [1] to curvilinear coordinates, and we conduct several
numerical examples to compare the results against our generalised capstan
equation.

Let Ω ⊂ R3 be a simply connected open bounded domain and let ∂Ω =


∂Ω0 ∪ ∂Ωf ∪ ∂ΩT be the sufficiently smooth boundary of the domain, where
meas(∂Ω0 ; R2 ), meas(∂Ωf ; R2 ), meas(∂ΩT ; R2 ) > 0, and where meas(·; Rk )
is standard Lebesgue measure in Rk (see chapter 6 of Schilling [39]). Now,
let X̄ : Ω → E3 be a diffeomorphism where X̄(x1 , x2 , x3 ) = σ(x1 , x2 ) +
x3 N (x1 , x2 ), σ ∈ C 2 (∂Ω0 ; R2 ) is an injective immersion and Ek is the kth-
dimensional Euclidean space. Note that by construction, we either have
x3 > 0 or x3 < 0 in Ω, but never both.

Now, assume that Ω describes the domain of an elastic body such that
∂Ω0 describes the region where the body is in contact with a rough rigid sur-
face, ∂Ωf describes the stress-free boundary and ∂ΩT describes the bound-
ary with traction. Let v ∈ C 2 (Ω; R3 ) be the displacement field of the
body. Given that f ∈ C 0 (Ω; R3 ) is an external force density field and
τ 0 ∈ C 0 (∂Ω0 ; R3 ) is a traction field (i.e. applied boundary-stress) at ∂ΩT ,
we can express the equations of equilibrium in curvilinear coordinates as
∇¯ i T i (v) + fj = 0, where T ij (v) = Aijkl Ekl (v) is the second Piola-Kirchhoff
j
stress tensor, Ekl (v) = 21 (∇ ¯ i vj + ∇
¯ j vi ) is the linearised Green-St Venant strain
ijkl
tensor and A = λg g + µ(g ik g jl + g il g jk ) is the isotropic elasticity ten-
ij kl

sor in curvilinear coordinates, λ = (1 − ν − 2ν 2 )−1 νE is the first Lamé’s


parameter, µ = 12 (1 + ν)−1 E is the second Lamé’s parameter, E ∈ (0, ∞)
is the Young’s modulus and ν ∈ (−1, 12 ) is the Poisson’s ratio of the of the
elastic body. The trivial boundary conditions are n̄i Tji (v) = 0 on ∂Ωf and
n̄i Tji (v) = τ0j on ∂ΩT , where n̄ is the unit outward normal to ∂Ω and ∇ ¯
is the covariant derivative operator. Note that for a thorough mathematical
analysis of the linear elasticity equations in a curvilinear setting, we refer the
reader to chapter 3 of Ciarlet [32].

15
To investigate the behaviour at the boundary ∂Ω0 , recall Kikuchi and
Oden’s model for Coulomb’s law of static friction [1]. Now, assume that X̄
is constructed such that v 3 describes the normal displacement and v β describe
the tangential displacements at the boundary ∂Ω0 . Now, define the boundary
as ∂Ω±0 = limx3 →0± Ω where sign depends on the construction of Ω. Noticing
that on ∂Ω− 0 , the form of the governing equations will remain identically to
equation (16) and on ∂Ω+ 3
0 , we have T3 (v)|∂Ω+
0
= −σn (v), simply re-express
the friction equation (16) in curvilinear coordinates to obtain the following,
v 3 |∂Ω±0 = 0 , (17)
 1
νF (g33 ) 2 v β 3 1
T3 (v)|∂Ω±0 , if (vα v α ) 2 |∂Ω±0 ≥ ε ,

∓

 1

T3β (v)|∂Ω±0 = (vα v α ) 2 (18)


1
 νF (g33 ) 2 v β

1
∓ T33 (v)|∂Ω±0 , if (vα v α ) 2 |∂Ω±0 < ε ,

ε
where νF is the coefficient of friction with respect to Coulomb’s law of static
friction, T33 (v)|∂Ω±0 is the purely normal stress and T3β (v)|∂Ω±0 are the shear
stresses. Note that in this framework we have g33 = 1, where gij = ∂i X̄k ∂j X̄ k
1 √
is the covariant metric tensor, and by convention (vα v α ) 2 = v1 v 1 + v2 v 2 .

Note that the model we derived, equations (17) and (18), explicitly de-
pends on the mechanical and geometrical properties of the elastic body. Now,
recall the standard friction law (2), which is a very simple law that does not
explicitly depend on any mechanical and geometrical properties of the body
in question. This raises the question that what is the relationship between
Coulomb’s law and the standard friction law for different values of geomet-
rical and elastic properties?

To investigate this matter, consider the map of a rigid semi-prism (x1 ,


a sin(x2 ), b cos(x2 ))E , where x2 ∈ [− 12 π, 12 π], a is the horizontal radius and b
is the vertical radius. Now, assume that an elastic body is over this prism
and one is applying a traction τ0 at x2 = − 21 π and a traction τmax at x2 = 12 π.
Assume further that the body in question has a thickness h, infinitely long
and in contact with an infinitely long semi-prism. This leads to the following
map of the unstrained configuration,
X̄(x1 , x2 , x3 ) = (x1 , a sin(x2 ), b cos(x2 ))E
+ x3 (ϕ(x2 ))−1 (0, b sin(x2 ), a cos(x2 ))E ,

16
1
where ϕ(x2 ) = (b2 sin2 (x2 ) + a2 cos2 (x2 )) 2 , x1 ∈ (−∞, ∞), x2 ∈ (− 21 π, 21 π)
and x3 ∈ (0, h). Now, let v = (0, v 2(x2 , x3 ), v 3(x2 , x3 )) be the displace-
ment field of the elastic body and let δv = (0, δv 2(x2 , x3 ), δv 3(x2 , x3 )) be a
perturbation of the displacement field. Now, we can express the governing
equations as follows,

(λ + µ)∂ 2 ∇ ¯ i v i + µ∆v
¯ 2=0,


(λ + µ)∂ 3 ∇ ¯ i v i + µ∆v
¯ 3=0,


(λ + µ)∂ 2 ∇ ¯ i δv i + µ∆δv¯ 2=0,




(λ + µ)∂ 3 ∇ ¯ i δv i + µ∆δv¯ 3=0,




where ∆¯ = ∇ ¯ i∇
¯ i is the vector-Laplacian operator in the curvilinear space
(see page 3 of Moon and Spencer [40]) with respect to ΩNew .

Eliminating x1 dependency, we can express the remaining boundaries as


follows,
New New
∂ΩNew = ∂ΩNew
0 ∪ ∂ΩNew
f ∪ ∂ΩT0 ∪ ∂ΩTmax ,

where
1 1
∂ΩNew
0 = {(− π, π) × {0}} ,
2 2
1 1
∂ΩNew
f = {(− π, π) × {h}} ,
2 2
1
∂ΩNew
T0 = {{− π} × (0, h)} ,
2
1
∂ΩNew
Tmax = {{ π} × (0, h)} .
2

17
Thus, the boundary conditions reduce to the following,

v 3 |∂ΩNew = 0 (zero-Dirichlet),
 2 2 2 2 3
 3
 0
(λ + 2µ) ∂2 v + Γ̄22 v + Γ̄23 v + λ∂3 v |∂ΩNew = τ0 (traction),
T0
2 2 2 2 3 3
  
(λ + 2µ) ∂2 v + Γ̄22 v + Γ̄23 v + λ∂3 v |∂ΩNew = τmax (traction),
Tmax
2 2 3
 
(ψ̄2 ) ∂3 v + ∂2 v |∂ΩNew
f ∪∂ΩNew New
T0 ∪∂ΩTmax
= 0 (zero-Robin),
λ ∂2 v 2 + Γ̄22
2 2 2 3
v + (λ + 2µ)∂3 v 3 |∂ΩNew
  
v + Γ̄23 = 0 (zero-Robin),
f

δv 2 |∂ΩNew ∪∂ΩNew ∪∂ΩNew = 0 ,


f T0 Tmax
3
δv |∂ΩNew = 0 ,
2 2
where Γ̄22 and Γ̄23 are the Christoffel symbols of the second kind, and ψ̄2 =
2 3 2 −2
ϕ(x ) + x ab(ϕ(x )) .

Now we can find the fiction laws governing the boundary conditions at
New
the boundary ∂Ω0 , which are:
If ψ̄2 |v 2 ||∂ΩNew
0
≥ ǫ, then

µψ̄2 ∂3 v 2 + νF sign(v 2 )T33 (v) |∂ΩNew


 
0
=0;

If ψ̄2 |v 2 ||∂ΩNew
0
< ǫ, then

µψ̄2 ∂3 δv 2 + νF ǫ−1 ψ̄2 v 2 T33 (δv)




+ νF ǫ−1 ψ̄2 δv 2 T33 (v) + µψ̄2 ∂3 v 2 + νF ǫ−1 ψ̄2 v 2 T33 (v) |∂ΩNew

0
=0,

where T33 (v) = λ(∂2 v 2 + Γ̄22


2 2 2 3
v + Γ̄23 v ) + (λ + 2µ)∂3 v 3 .

Despite the fact that the original problem is three-dimensional, as a result


of problem’s invariance in the x1 direction, it is now to a two-dimensional
problem as the domain resides in the set {(x2 , x3 ) | (x2 , x3 ) ∈ [− 12 π, 12 π] ×
[0, h]}. We are fully aware that Kikuchi and Oden’s model [1] is only de-
fend for bounded domains; however, we now show that the reduced two-
dimensional problem is numerically sound.

To conduct numerical experiments, we use the second-order accurate


iterative-Jacobi finite-difference method with Newton’s method for nonlinear

18
systems (see chapter 10 of Burden et al. [41]). The grid dependence is intro-
duced in the discretisation of the (reduced two-dimensional) domain implies
that the condition ψ0 ∆x2 ≤ ∆x3 , ∀ ψ0 ∈ {ψ̄2 (x2 , x3 ) | x2 ∈ [− 12 π, 21 π] and x3 ∈
[0, h]} must be satisfied, where ∆xj is a small increment in xj direction.
For our purposes, we use ∆x2 = N 1−1 π and ψ0 = ψ̄2 ( 14 π, h), where N =
250. Note that for all our examples, we fix the values νF = 21 , τ0 = 1
and a = 2 for our constants, and we assume the default values τmax = 2,
b = 2, h = 1, E = 103 , ν = 14 and ε = 10−5 for our variables, un-
less it strictly says otherwise. Also, all numerical codes are available at
http://discovery.ucl.ac.uk/id/eprint/1532145.

×10-3 ×10-4

0
1.5
-1
1
Azimuthal Displacement

-2
Radial Displacement

0.5 -3

0 -4

-5
-0.5
-6
-1
-7
-1.5 -8

-2 -9
2 1

1.5 1.5

1 2 -1 -2
1 -1 0 2 1 0
Radius -2 Radius 2
θ θ

Figure 2: Displacement field of the modified Kikuchi and Oden’s model.

Figures 2 and 3 are calculated with the values τmax = 1, b = 2, h = 1,


E = 103 , ν = 41 , ε = 10−5 and with a grid of 250 × 41 points. Figure
2 shows the azimuthal (i.e v 2 ) and the radial (i.e v 3 ) displacements. The
maximum azimuthal displacements are observed at (θ =) x2 = ± 21 π, with
respective azimuthal displacements of v 2 = ±1.72 × 10−3. The maximum
radial displacement is observed at x2 = ± 21 π, with a radial displacement of
v 3 = −8.24 × 10−4 . Also, figure 3 shows the behaviour of the elastic body at
New
the boundary ∂Ω0 . It implies that in the region [− 21 π, −0.0694] the body
slid in the negative (i.e. decreasing) azimuthal direction, and in the region
[0.0694, 12 π] the body slid in the positive (i.e. increasing) azimuthal direction.
The region (−0.0201, 0.0201) describes the azimuthal region of the body that

19
1

0.8

0.6

-1,+1 = Slip, 0 = No Slip


0.4

0.2

-0.2

-0.4

-0.6

-0.8

-1
-2 -1 0 1 2
θ

Figure 3: Slip and stick regions of the modified Kikuchi and Oden’s model.

is bonded to the rigid obstacle, i.e. magnitude of the displacement is below


the threshold of sliding determined by the Coulomb’s friction condition. Note
that this ε is not a physical parameter as it has no real life significance, i.e. it
is merely introduced make Coulomb’s friction law non-singular in the bonded
region. Further note that if the applied traction at x2 = 21 π is larger than
the traction at x2 = − 12 π, then figure 3 will no longer be symmetric in the
azimuthal direction, and for this case the right side (i.e. where slip = 1)
will be much greater than the left side (i.e. where slip = −1). Thus, if one
increases the traction at x2 = 12 π enough, then the entire body will slip in
the positive azimuthal direction, i.e. slip = 1, ∀ θ. Thus, the subject of of
the remainder of this section to investigate how this slip region behaves for
a given set of variables, as this will infer the relationship between Coulomb’s
law and the standard friction law.

Figure 4a shows how close the elastic body is to its limiting-equilibrium


for varying tension ratio. We see that as the tension ratio increases, the elas-
tic body gets closer and closer to fully debonding. In fact, when δτ = 2.75,
our elastic body is fully debonded from the rigid foundation and sliding in
the positive x2 direction. To compare it against the capstan equation, we
invoke corollary 1 with δτ = 2.75, which implies that the capstan coeffi-
cient of friction is µF = 0.322, regardless of the Poisson’s ratio of the body.
This result implies that the capstan coefficient of friction is an underesti-

20
1 0.9
Coulombs
0.9
0.85
0.8
0.8
0.7
0.75
0.6
Slip

Slip
0.5 0.7

0.4
0.65
0.3
0.6
0.2
Limiting Equilibrium 0.55
0.1
Coulombs

0 0.5
1 1.5 2 2.5 3 0.5 1 1.5
δτ δb

(a) Slip for δτ . (b) Slip for δb.


0.75 0.717
Coulombs
0.716
0.7
0.715

0.65 0.714

0.713
Slip

Slip

0.6
0.712

0.55 0.711

0.71
0.5
0.709
Coulombs

0.45 0.708
0 0.2 0.4 0.6 0.8 1 0 0.05 0.1 0.15 0.2 0.25 0.3
h ν

(c) Slip for h. (d) Slip for ν.

Figure 4: Coulomb’s law of static friction limiting-equilibrium for varying


δτ , δb, h and ν.

mate of Coulomb’s coefficient of friction, i.e. µF ≤ νF . Note that capstan


equation underestimating the actual coefficient of friction is a documented
phenomenon in the literature [42].

Figure 4b shows how close the elastic body is to its limiting-equilibrium


for varying curvature of the contact region. We see that as the curvature
increases, the modified Kikuchi and Oden’s model moves away from limiting-
equilibrium. For the contact interval (− 12 π, 21 π), the observed result coincides
with capstan equation for an elliptical prism case (14). Note that we inter-
pret δb as the curvature as δb is correlated with the mean-curvature (i.e.

21
H(x2 ) = 21 ab(ϕ(x2 ))−3 ) in the set x2 ∈ (− 21 π, 12 π).

Figure 4c shows how close the elastic body is to its limiting-equilibrium


for varying thickness of the overlying body. We see that as h decreases, the
modified Kikuchi and Oden’s model moves away from limiting-equilibrium,
i.e. as the thickness decreases, Coulomb’s law of static friction behaves more
like the standard friction law. This possibly imply a convergence between
both model for membranes (and strings).

Another numerical experiment that we conduct examines the behaviour


of the body under varying Young’s modulus, E, and for this experiment we
consider E ∈ [500, 1500]. We find that the body has a constant slip of 71.2%
in the positive x2 direction for all values of Young’s moduli. This is intuitive
as, whatever the value of Young’s modulus is (given that it is not zero or
infinite), one can always rescale the displacement field without affecting the
final form of the solution. This result also tends to coincide with the capstan
equation, as the capstan equation is also invariant with respect to Young’s
modulus of the elastic body. Note that for values above E = 1010 , we ob-
serve total bonding. We hypothesise that this is due to the magnitude of the
regularisation parameter, ε, and not to do with any physical realistic proper-
ties as the displacement field can be rescaled for any given value of Young’s
modulus. Our hypothesis is further justified as when the magnitude of the
regularisation parameter is decreased, we did not observe total bonding for
the case E = 1010 .

Figure 4d shows how close to the elastic body is to its limiting-equilibrium


for varying Poisson’s ratio. We see that as ν increases, the modified Kikuchi
and Oden’s model moves away from the limiting equilibrium, i.e. as the
body becomes incompressible, one needs to apply more force to debond the
body from the rigid surface. This is a surprising result as this tends to
contradict the capstan equation, as the modified capstan equation for the
infinite width case (corollary 1) is invariant with respect to Poisson’s ratio
of the elastic body. This effect is small in comparison to other results, but it
is still numerically observable.

22
4. Conclusions
In our analysis, we taken the standard friction law (2) and extend it to
model thin objects on rough rigid surfaces in an attempt to extend the cap-
stan equation (1) to more general geometries. First, we derived closed from
solutions for a membrane with a zero-Poisson’s ratio (or a string with an
arbitrary Poisson’s ratio) supported by a rigid prism at limiting-equilibrium
(static-equilibrium or steady-equilibrium), and then a rigid cone at limiting-
equilibrium (static-equilibrium case only). Although, we used the standard
friction law to generalise the capstan equation to noncircular geometries,
what we derived can be regarded as a extended case of Konyukhov’s [27] and
Konyuk-hov’s and Izi’s [28] model; however, what is unique about our work
is that we are considering a true-membrane over a rigid obstacle, and we are
able incorporate an external loading field (prism case only).

As a comparison, we extended Kikuchi and Oden’s [1] model for Coulomb’s


law of static friction to curvilinear coordinates. For this model, we conducted
numerical experiments to observe the relationship between Coulomb’s law of
static friction and the ordinary friction law (2) implied by our generalised
capstan equation, in curvilinear coordinates. To do this, we modelled an
elastic body over a rigid rough prism with an elliptical cross section. Our
numerical results indicate the following: for a fixed coefficient of friction (i)
the capstan coefficient of friction is an underestimate of Coulomb’s coeffi-
cient of friction, i.e. µF ≤ νF ; (ii) as the curvature of the contact region
increases, one require a larger force to debond the body, which is a result
that coincides with the modified capstan equation; (iii) as the thickness of
the body decreases, one require a larger force to debond the body, yet this
force is still an underestimate to what is predicted by the modified capstan
equation; (iv) Young’s modulus of the body does not affect the governing
equation of the contact region, which is another result that coincides with
the capstan equation (modified or otherwise); and (v) as Poisson’s ratio of
the body increases, one requires a larger force to debond the body. The last
result implies that incompressible elastic bodies, such as rubber, tend to be
more difficult to debond from a rigid surface relative to a compressible ob-
ject with the same coefficient of friction acting on the contact region. Now,
this result is a surprising as this behaviour cannot predicted by the modified
capstan equation as the modified capstan equation is invariant with respect
to Poisson’s ratio of the body (for the problem that we considered).

23
Conclusion inferred by our numerical modelling can be summarised as
follows: for the same value of coefficient of friction, different models predict
vastly different different limiting-equilibriums, which, in turns, implies that
calculating the coefficient of friction is model dependent. Thus, our analysis
shows that modelling friction is not a well understood problem as different
models (that supposedly model the same physical phenomena with common
roots stretching back to Amontons’ laws of friction) predict different out-
comes.

References
[1] N. Kikuchi, J. Oden, Contact Problems in Elasticity: A Study of Vari-
ational Inequalities and Finite Element Methods, Studies in Applied
Mathematics, Society for Industrial and Applied Mathematics, 1988.

[2] D. Quadling, Mechanics 1, Cambridge Advanced Level Mathematics,


Cambridge University Press, 2004.

[3] K. L. Johnson, K. L. Johnson, Contact mechanics, Cambridge university


press, 1987.

[4] C. L. Rao, J. Lakshinarashiman, R. Sethuraman, S. M. Sivakumar, En-


gineering Mechanics: Statics and Dynamics, PHI Learning Pvt. Ltd.,
2003.

[5] J. H. Jung, N. Pan, T. J. Kang, Capstan equation including bending


rigidity and non-linear frictional behavior, Mechanism and Machine
Theory 43 (2008) 661–675.

[6] O. Baser, E. I. Konukseven, Theoretical and experimental determination


of capstan drive slip error, Mechanism and Machine Theory 45 (2010)
815–827.

[7] S. Kang, H. In, K. J. Cho, Design of a passive brake mechanism for


tendon driven devices, International Journal of Precision Engineering
and Manufacturing 13 (2012) 1487–1490.

24
[8] G. Stellin, G. Cappiello, S. Roccella, M. C. Carrozza, P. Dario, G. Metta,
G. Sandini, F. Becchi, Preliminary design of an anthropomorphic dex-
terous hand for a 2-years-old humanoid: towards cognition, Hand 20
(2006) 9.

[9] S. Behzadipour, A. Khajepour, Cable-based robot manipulators with


translational degrees of freedom, INTECH Open Access Publisher, 2006.

[10] S. J. Ball, I. E. Brown, S. H. Scott, A planar 3dof robotic exoskeleton for


rehabilitation and assessment, in: Engineering in Medicine and Biology
Society, 2007. EMBS 2007. 29th Annual International Conference of the
IEEE, IEEE, pp. 4024–4027.

[11] S. K. Mustafa, G. Yang, S. H. Yeo, W. Lin, Kinematic calibration of


a 7-dof self-calibrated modular cable-driven robotic arm, in: Robotics
and Automation, 2008. ICRA 2008. IEEE International Conference on,
IEEE, pp. 1288–1293.

[12] J. C. Perry, J. Rosen, Design of a 7 degree-of-freedom upper-limb pow-


ered exoskeleton, in: Biomedical Robotics and Biomechatronics, 2006.
BioRob 2006. The First IEEE/RAS-EMBS International Conference on,
IEEE, pp. 805–810.

[13] J. H. Jung, T. J. Kang, J. R. Youn, Effect of bending rigidity on the


capstan equation, Textile research journal 74 (2004) 1085–1096.

[14] R. Christensen, Log-linear models and logistic regression, Springer Sci-


ence & Business Media, 2006.

[15] P. Grosberg, D. E. A. Plate, 19—capstan friction for polymer monofila-


ments with rigidity, Journal of the Textile Institute 60 (1969) 268–283.

[16] G. Endo, H. Yamada, A. Yajima, M. Ogata, S. Hirose, A passive weight


compensation mechanism with a non-circular pulley and a spring, in:
Robotics and Automation (ICRA), 2010 IEEE International Conference
on, IEEE, pp. 3843–3848.

[17] B. Kim, A. D. Deshpande, Design of nonlinear rotational stiffness us-


ing a noncircular pulley-spring mechanism, Journal of Mechanisms and
Robotics 6 (2014) 041009.

25
[18] P. Krawiec, et al., Analysis of generation capabilities of noncircular
cogbelt pulleys on the example of a gear with an elliptical pitch line,
Journal of Manufacturing Science and Engineering 133 (2011) 051006.

[19] D. Shin, X. Yeh, O. Khatib, Variable radius pulley design methodology


for pneumatic artificial muscle-based antagonistic actuation systems, in:
Intelligent Robots and Systems (IROS), 2011 IEEE/RSJ International
Conference on, IEEE, pp. 1830–1835.

[20] E. Zheng, F. Jia, H. Sha, S. Wang, Non-circular belt transmission design


of mechanical press, Mechanism and Machine Theory 57 (2012) 126–138.

[21] Y. Lu, D. Fan, Transmission backlash of precise cable drive system,


Proceedings of the Institution of Mechanical Engineers, Part C: Journal
of Mechanical Engineering Science 227 (2013) 2256–2267.

[22] S. Döonmez, A. Marmarali, A model for predicting a yarn’s knittability,


Textile research journal 74 (2004) 1049–1054.

[23] M. Wei, R. Chen, An improved capstan equation for nonflexible fibers


and yarns, Textile research journal 68 (1998) 487–492.

[24] J. H. Jung, N. Pan, T. J. K., Generalized capstan problem: Bending


rigidity, nonlinear friction, and extensibility effect, Tribology Interna-
tional 41 (2008) 524–534.

[25] Y. S. Kim, M. K. Jain, D. R. Metzger, A finite element study of cap-


stan friction test, in: AIP Conference Proceedings, volume 712, IOP
INSTITUTE OF PHYSICS PUBLISHING LTD, pp. 2264–2269.

[26] I. M. Stuart, Capstan equation for strings with rigidity, British Journal
of Applied Physics 12 (1961) 559.

[27] A. Konyukhov, Contact of ropes and orthotropic rough surfaces,


ZAMM-Journal of Applied Mathematics and Mechanics/Zeitschrift für
Angewandte Mathematik und Mechanik 95 (2015) 406–423.

[28] A. Konyukhov, R. Izi, Introduction to computational contact mechanics:


a geometrical approach, John Wiley & Sons, 2015.

26
[29] R. M. Overney, Introduction to tribology – fric-
tion, http://depts.washington.edu/nanolab/ChemE554
/Summaries%20ChemE%20554/Introduction%20Tribology.htm, 2007.

[30] A. Libai, J. G. Simmonds, The nonlinear theory of elastic shells, Cam-


bridge university press, 2005.

[31] D. Kay, Schaum’s Outline of Tensor Calculus, McGraw Hill Professional,


1988.

[32] P. G. Ciarlet, An introduction to differential geometry with applications


to elasticity, Journal of Elasticity 78 (2005) 1–215.

[33] A. Morassi, R. Paroni, Classical and Advanced Theories of Thin Struc-


tures: Mechanical and Mathematical Aspects, CISM International Cen-
tre for Mechanical Sciences, Springer Vienna, 2009.

[34] L. C. Evans, Partial Differential Equations, Graduate studies in math-


ematics, American Mathematical Society, 2010.

[35] A. N. Pressley, Elementary differential geometry, Springer Science &


Business Media, 2010.

[36] M. Spivak, Calculus on manifolds, volume 1, WA Benjamin New York,


1965.

[37] K. Jayawardana, Mathematical Theory of Shells on Elastic Foundations:


An Analysis of Boundary Forms, Constraints, and Applications to Fric-
tion and Skin Abrasion, Ph.D. thesis, UCL (University College London),
2016.

[38] K. Jayawardana, N. C. Ovenden, A. Cottenden, Quantifying the fric-


tional forces between skin and nonwoven fabrics, Frontiers in Physiology
8 (2017).

[39] R. L. Schilling, Measures, integrals and martingales, volume 13, Cam-


bridge University Press, 2005.

[40] P. Moon, D. E. Spencer, Field theory handbook: including coordinate


systems, differential equations and their solutions, Springer, 2012.

27
[41] R. Burden, J. Faires, A. Burden, Numerical analysis, Nelson Education,
2015.

[42] A. J. P. Martin, R. Mittelmann, 18—some measurements of the friction


of wool and mohair, Journal of the Textile Institute Transactions 37
(1946) T269–T280.

28

You might also like