You are on page 1of 23

26/10/2017

PEF-5750
Estruturas Leves
Ruy Marcelo de Oliveira Pauletti

PARAMETRIC SURFACES
EQUILIBRIUM OF MEMBRANES
MINIMAL SURFACES

(plus some preliminaries:


CAUCHY STRESSES &
THEOREM OF VIRTUAL WORK)

24/10/2017

Cauchy Stresses:
f
z z
f
z
y y
O
y O O
x x x

b P t P
b
n dA
B
Ω B

(a) A solid under displacement constraints and external loads;


(b/c) Internal stresses in a constrained and loaded solid

F
Cauchy Stress (“True Stress”): t  t  P , n   lim   t  P , n 
A  0 A

It can be shown that t  P, n   T  P  n

where T is the Cauchy Tensor Field


In an implicit Cartesian basis ei i  1, 2, 3  T    ij  where  ij  ei  t  e j 

Over the boundary surface: Tn  f

1
26/10/2017

Equilibrium
For any finite portion of the body B

 f dA   b dV  0

 T n dA   b dV  0
 
(Integral Equilibrium Equation)

Using Gauss Divergence Theorem:  div dV    n dA


 

where  is either a scalar, vector or tensor field

we have   divT  b  dV  0,   B
T


(Differential
 divTT  b  0, P  B
and since this must hold for   Equilibrium
 T n  f ,  P  B Equation)
3 
where divTT  
ij
ei
i , j 1 x j

Equilibrium of moments around three coordinate axis provides T  TT

So the transposition symbol can be omitted in the equilibrium equation, which then
expresses both force and moment equilibrium!

Theorem of Virtual Work


 divT  b  0 P  B
A body is in equilibrium, if, and only f 
 T n  f P  Bf
Thus, for any virtual displacement u compatible with the boundaries  u  0 in Bu 

 u   divT  b   0 P  B

  u   divT  b  dV  0

 u

Applying Green’s divergence Theorem for the first term:

  u   divT  dV    u   T n  dA   T :  E dV
T

  

Thus equilibrium can be expressed by  T :  E dV    u  t dA    u  b dV ,


  
 u

That is  Wint   Wext ,  u


1
 E   F   F T    ET is the tensor of virtual displacements
2
   u  
 F   u    is the gradient of the virtual transformation
 x 

2
26/10/2017

Plane Stress State


f
f
B
 z  0  x  xy 0
 
b  xz   yz  0  T   yx y 0 
b b  f 0  0 0 
 z z
0
B
t
y

A solid under plane stress state

  x  xy
   bx  0
 x y
Equilibrium: P  B , 
  yx   y  b  0
 x y
y

Parametric Surfaces

A parametric surface can be described by a position vector field r  r 1 ,  2 

(we´ll refer to an implicit Cartesian coordinate system (x,y,z), so that no distinction will be done
between covariant and contravariant base vectors, as it would be required for more general curvilinear
coordinate systems…)

3
26/10/2017

Parametric Surfaces
Some solids can be descripted by parametric surfaces, plus a ‘thickness’ scalar
parametric field given

h  h  1 ,  2 

Such that: P  B : x P  r 1 ,  2   z e 3 1 ,  2 

where:  h h
z   , 
 2 2
P
and: 0  h   min h

xP r1,2)

MEMBRANES: h  1
O

Parametric Surfaces
g3 g2

dA
g1

Keeping   constant,   1, 2

we define an in infinite set of coordinate r


curves, with associate tangent vectors g  (In general, g  1 )
 

g1  g 2
A unit vector field, always normal to te surface, is given by g3 
g1  g 2

Differential area elements are given by dA  g1  g 2   g1  g 2   g 3  det G

Where G   aij  , aij  g i  g j , i, j  1, 2, 3 is the METRIC TENSOR

4
26/10/2017

Parametric Surfaces
A differential displacement over the surface is given by

r r 2
d 2   g d
dr
dr  d 1 
 1  2  1

r r
r + dr where g 
 
O

2
(squared) length of an infinitesimal displacement:  ds   g   g  d  d  
2
 dr  dr 
 
, 1

2 2
Denoting a  g  g   ds     a d  d  
2
 ,   1, 2 we have
 1  1

‘FIRST FUNDAMENTAL FORM


OF THE SURFACE’

Normal Curvature

We consider a curve drawn onto the surface, parametrized by the arc-length:

   : r  r s

g3
1 = 10
ds
 


r0 r(s)
2 = 20
O

• dτ ν g
The curvature of  is given by τ   3
ds  

5
26/10/2017

Normal Curvature
g3
1 = 10
ds

  • dτ ν g
τ   3
r0 r(s)
2 = 20 ds  
O

O
dg3 C0
g3
g3 g3 + dg3
 
P0 ds  C dg3
d g3
P 
(s)  + d g3 = 
 d
r(s) ds
r(s) + dr C P
P0  + d
 

g3 + dg3
O C0
• dg 3 τ
And since g3  τ s  g 3  
ds 

• τ  1
So that, in any case g3  τ     τ 

  

Normal Curvature
1 •
may now be positive or negative, depending on the curve being
Thus   g3  τ concave or convex with respect to the surface orientation!

• dg 3 2
g 3 d   2
r d   2
d

dr
g3  τ    g 
ds  1   ds ds  1  ds  1 ds

1 • 1  2 2  g  
  g3  τ      g  3  d d  
  ds   
2
  1  1   

g 3
Denoting b  b  g   ,   1, 2
 
2 2

1




b d d 
1 1
‘’SECOND FUNDAMENTAL FORM OF THE SURFACE’
 
 2 2





a d d 
1 1
‘FIRST FUNDAMENTAL FORM OF THE SURFACE’

6
26/10/2017

Principal Curvatures

1 b11 d12  2b12 d1 d 2  b22 d 2 2


Expanding these forms  
 a11 d12  2 a12 d1 d 2  a22 d  2 2

 A  A  d 1 , d  2   a11 d 12  2a12 d 1 d  2  a22 d  2 2


Defining 
 B  B  d1 , d 2   b11 d 1  2b12 d 1 d  2  b22 d  2
2 2

We arrive at A  B  0 , which always holds!

(that is, for any direction defined by  d 1 , d  2  )


 A  B
Therefore  A  B   A  0  ,   1, 2
d   d   d   d  
  d1 , d 2 

Principal Curvatures
Furthermore, in the directions for which k is extremum:

0  ,   1, 2
d  

That is, in the directions for which k is extremum:

A B
 0  ,   1, 2
d   d  

Proceeding with the derivations:

 a11 d1  a12 d 2    b11 d1  b12 d  2  0



 a12 d1  a22 d 2    b12 d1  b22 d  2  0

7
26/10/2017

Principal Curvatures
  b11 b12   a11 a12    d 1   0 
Or, in matrix form:      (*)
 b12 b22   a12 a22    d  2   0 

  b11   a11 b12   a12  


This system has non-trivial solutions only if det   0
 b12   a12 b22   a22  

‘Characteristic equation’: a 11 22a  a12 2   2   a11b22  2a12 b12  b11 a22  k   b11b22  b12 2   0

The roots of the characteristic equations are:  I , II  K M  K M2  K G

b11b22  b122
where: K G   I  II  Is the GAUSSIAN CURVATURE
a11 a22  a122

1 1 a b  2 a12 b12  b11 a22 Is the MEAN CURVATURE


and where: KM   I   II   11 22
2 2 a11 a22  a122

Principal Curvatures
KG  P   0 KG  P   0 KG  P   0

  
The point is ELLIPTIC The point is PARABOLIC The point is HIPERBOLIC

K G  0 P   K G  0 P  
K G  0 P  
  

The surface is The surface is SIMPLE The surface is


SINCLASTIC or DEVELOPABLE ANTICLASTIC

Relationships between Gaussian  I   II  I   II  I   II


and Mean Curvatures:
KM     KG ; 
2 2  I  II 2

8
26/10/2017

Principal Directions & Curvature Lines


Substituting kI and kII in (*) we find the two mutually orthogonal direction I , II ,
for which the curvature radiuses are extrema;

These directions are called PRINCIPAL DIRECTIONS, and curves that follow the
principal directions are called CURVATURE LINES.

To show that indeed  I   II , we consider that along the coordinate lines d   0


(since   constant ), so for curvature lines we have
0 0
 a11 d1  a12 d  2   I  b11 d 1  b12 d  2  0  a12 I  b12  0
 That is 
 a12 d1  a22 d  2   II  b12 d1  b22 d  2  0  a12 II  b12  0
0 0 a12   I   II   0

Now, if  I   II  a12  g1  g 2  0  g1  g 2 , that is, indeed the curvature lines are


mutually orthogonal!

When curvatures lines are used as coordinate lines, b11 b22


xpressions for the principal curvatures are simplified: I  ;  II 
a11 a22

If  I   II  0 , the vicinity of the point is locally spherical,


and all directions are principal ones.

Curvature Lines
Determination of the Curvature Lines may be complicated, but, by assuming they are
known, they may be taken as coordinate lines, respect to which the equilibrium of
membranes is simpler to express!

We also keep the arc-length along the curvature lines as parameter, that is   s

And we define, at each point, a local Cartesian coordinate system such that

 r
 e  ,   1, 2 e3
 s e2
 e 3  e1  e 2
O
 e1

9
26/10/2017

Curvature Lines
• It can be shown that necessary  e 3
and sufficient conditions for    e   0
 
coordinates lines be curvature  ,   1, 2 ;   2,1
lines are:  e   e  0
  
3

e 3
• Reciprocal conditions are:  e  0 ;   1, 2 ;   2,1
 

Torsion
 e e 
The TORSION of a curve is defined such that    e 3
s  s
It expresses ‘how fast’ the osculating plane  e1 , e 2  rotates, in the vicinity of a point

e 2
e2 de 2  ds1   e 3 ds1
e3 s1
e 3  de 3 e 2
e3

e 2  de 2
ds1

Torsion
Curvature lines do not present any torsion, since the previous conditions require that,
for them:
 e
 0 ;   1, 2 ;   2,1
s 

(a) Lines with torsion (b) curvature lines (without torsion)

10
26/10/2017

A gridshell composed of metal plates that run along asymptotic lines


(lines of zero normal curvature) – TUM, 2017
E. Schling, D. Hitrec, J. Schikore, R. Barthel. DESIGN AND CONSTRUCTION OF THE ASYMPTOTIC PAVILION,
STRUCTURAL MEMBRANES 2017, Munich.

Euler’s Theorem & Curvature Tensor

The curvature of a line making an angle  with a curvature line is

   I cos 2    II sin 2 

Eulers Theorem highlights the tensor nature of the curvatures around a point…

The components of the Surface Curvature Tensor on a point, in an intrinsic


orthogonal coordinate system are given by:

 e
 K  g 3  ,   1, 2 
s 

 1  
that is K   K    
   2 

 I ,  II are the eigenvalues of K and  I ,  II its eigenvectors

11
26/10/2017

DIVERGENCE OF A SURFACE TENSOR


2 2
If T is resolved as a SURFACE TENSOR T  T  s1 , s2     T e  e 
 1  1

(where we consider orthogonal coordinates lines with arc-length parameters s


-- not necessarily curvature lines! )

2  T 
2
Then its divergence vector is divT   

1  1 s 
 e   K : T  e3
 
  

• where K is the Surface Curvature Tensor

• and where the scalar product between tensors K and T :

K : T  tr  K T T     K  T
2 2

 1  1

Equilibrium of Membranes - I

(b) A transversal cut showing


(a) External and boundary loads in a membrane,  33
the variation of
with h<<1 (constant);
(usually disregarded!)

Equilibrium of Forces:

 
h Tn ds   bA

 
Applying the divergence theorem   h  div T  b  dA  0 , 
 
 h  div T  b  0 P   (Equilibrium of Forces)

T  TT (Equilibrium of Momentum)

12
26/10/2017

Equilibrium of Membranes - I
Remembering the expression of the divergence of a surface tensor equation
And denoting the external loads as

b  b1e1  b2 e 2  pe 3

        22  
h   11  12  e1   12 
 s  e 2   K : T  e 3   f1e1  f 2 e 2  pe 3  0
 1  s 2  
 1 s  s 2  

  11  12  
 h    b1 
  s1 s 2  
  Similar to a plane stress state
  12  22  
 h    b2 
  s1 s 2  

 h K : T  p
  Surface stresses equilibrate transversal load by
means of curvature!

  1    11  12     11  22 
htr     h  1 11   2 22  2 12   h    2 12   p
    2   12  22   
 1  2 

Equilibrium of Membranes - II
The previous derivation might seems too much abstract, so we recast the
membrane equilibrium in a more ‘engineering’ approach:

2

21 11

22

(a) External and boundary loads (b) a membrane element under a


in a membrane; surface stress field

We consider   and   constant through thickness h. and


define the Surface Stresses [N/m]:

N   h  ; N   N   h 

13
26/10/2017

Equilibrium of Membranes - II
2

N 2   N 02  dN 2 We again consider coordinates lines with


 e
0
11 1
b length parameters s
(not necessarily curvature lines!)
 120 e2 ds1

N10
 210 e1 N1  N10  dN1  N1  N11e1  N12 e 2
ds2 We define: 
 N 2  N 21e1  N 22 e 2
 220 e 2
N 02

1

Forces: dN1 ds2  dN 2 ds1  bds1ds2  0 (*)


And express Equilibrium:
Moments: N12  N 21

d N1 d N 2 (**)
Dividing (*) by ds1 ds2 :  b  0
ds1 ds2

Equilibrium of Membranes - II

Now:
d N1 d N e N e
  N11e1  N12 e 2   11 e1  N11 1  12 e 2  N12 2
ds1 ds1 s1 s1 s1 s1

e1 e 2
But  1e 3 and   e 3
s1 s1

N1 N11 N
Therefore  e1  12 e 2   k1 N11   N12  e3
ds1 s1 ds1

Likewise
N 2 N 21 N 22
 e1  e 2   k 2 N 22   N 21  e3
ds2 s 2 ds2

Substituting in (**):

N11 N N 21 N 22
e1  12 e 2   k1 N11   N12  e3  e1  e 2   k 2 N 22   N 21  e3  b  0
s1 ds1 s 2 ds2

14
26/10/2017

Differential Equilibrium in Membranes


Denoting b  b1e1  b2 e 2  pe 3 , considering that N12=N21
and collecting the terms according to the intrinsic basis:

 N11 N12 
   b1 
s1 s 2  Similar to a plane
  stress state!
 N12 N 22 
   b2 
 s1 s 2
N N
 11  22  2 N12  p “General Membrane Equation”
 1 2

If  are principal directions:  0

 N11 N12 
   b1 
1  2  Similar to a plane
  stress state!
 N12 N 22 
   b2 
 1  2
 N11 N 22 “General Membrane Equation
   p in Principal Directions”
 I  II

Differential Equilibrium in Membranes


If the membrane is under a uniform and isotropic stress field

 b1  b2  0
 N11  N 22   0 
    1 1  Equation of Laplace-Young or
 N12  N 21  0  h 0       p “Soap Films Equation”
  I II 

    II 
Rewriting: h 0  I   2 h 0 K M  p
  I  II 

In particular, for a spherical membrane, of radius r and thickness t:

1 𝑝𝑟
KM  ⇒ 𝜎0 =
r 2𝑡

15
26/10/2017

Minimum Surfaces (Plateu’s Problem)


The Area of a Surface  with a prescribed closed boundary is given by
A   dA   g1  g 2 d 1 d  2
 

We seek a surface  Spanned by a vector field x


* *
Such that A
*
Is a minimum.

In other words, for every compatible perturbation u around x*


A
 A*   u  0,  u
x x*

Thus the 1st order condition for x * To be a minimum is

A
0
x x*

With the equality constraint  x P  x P   0, P  


Where xP Spans the prescribe coordinates

Analytical solutions for this nonlinear equation maybe rather difficult!

Soap Film Analogy


We recast the problem of area minimization in a more nonlinear mechanics fashion:

 A     J  dA  0,  u around x *

is the Jacobian of the   u 


where  J  tr   F  F  I  from x* to x*   u
deformation gradient x *
Measured with respect to x*

The configuration x * That minimizes the functional A also fulfills the above equation  u

We can always consider coordinate systems for which e3  g3 , normal to *

So that  F33  0

And therefore A  tr  F  dA  0    F 11   F22  dA  0  u


* *

16
26/10/2017

Soap Film Analogy


We now consider the problem of finding the membrane geometry x

Compatible with a self-equilibrated Cauchy surface stress field T


(zero external loads, except reactions along the fixed boundaries)

Static equilibrium always requires  Wint   Wext


But since we assume that external loads are zero (except at fixed boundaries):  Wext  0

Therefore, for self-equilibrated surfaces:  Wint  0,  u

When deformations are measured from the equilibrium configuration x

This condition can be recast as  Wint  h  T :  F dA  0,  u



  u 
Where h is the membrane thickness and F  I 
x

Soap Film Analogy

We now particularize this condition for a homogeneous and isotropic Cauchy surface
stress field, such that, in the chosen basis (with e3  g3 )

1 0 0 
T    0 1 0    1, P  
 0 0 0 

Even if the membrane presents non-zero deformations in the transversal direction,


that is  F33  0 , the virtual work developed by the stress field T is given by

 Wint   h  1 :  F dS   h   F11   F22  dA  0,  u


 

Comparing to the condition to minimum area we conclude that x  x*

That is, the geometry compatible to an auto-equilibrated, homogeneous and isotropic


Cauchy surface stress field also complies to the condition of minimal area!

17
26/10/2017

Consequence for Minimal Surfaces


N11 N 22
Considering the Membranes’ Equation   p
I  II

p  0
And taking into account that for minimum surfaces 
 N11  N 22  N 0

 1 1 
We have N0   0
  I  II 

Thus  I    II

Or, in words, minimal surfaces have zero mean curvature KM, with negative or zero
Gaussian curvature KG, that is, they are either anticlastic or flat!
• Soap films are minimal surfaces, when p=0.

• Strictly speaking, a soap bubble is not a minimal surface, although it is associated


to a similar minimization problem (‘minimize the area for a given volume’).

Ex. 8 - Hyperbolic Paraboloid


Show that a hyperbolic paraboloid (“hypar” )surface IS NOT a minimum surface.

z  axy

Hypars are usually mentioned as minimum surfaces in the literature about tension
structures, and in fact they are practically indistinct to the minimum surface with the
same boundary. The analytical expression of the true minimal surface was given by
Schwarz in 1890, and is quite more complicated than the hypar expression!

18
26/10/2017

Catenoid
The soap film analogy provides a shortcut to deduce the shape of the catenoid, that
is, the shape of a minimal surface connecting two circular rings of radius R, with
center on the save axis, and apart of each other by a distant h.

Symmetry requires that the solution is given by the rotation of a generatrix curve

y  yz

and that this curve is symmetric at mid-distance between the rings that is,
y  0   0

Catenoid
We consider the vertical equilibrium of a slice of the surface bounded by to vertical
radial planes, with and infinitesimal vertical strip spanned by horizontal angle d 

We denote y  0   y0 , so far an unknown constant value.

Using the soap analogy, we know that every cross section of the membrane is under a
uniform surface stress  0 t , where t is the membrane thickness.

Thus vertical equilibrium of a vertical slice from z=0 to z(y) requires:

 0t y0 d    0t yd   cos 

dz
And since cos    1   y
2

ds

y
We have y0 
1   y
2

19
26/10/2017

Catenoid
y 1
Rearranging: 
y y
2 2
0
y0

 y  1 z
Integrating along z:   2
dz   dz 
2 
y  y0  y y
C
 0 0

From standard calculus 


du
u2  a2

 ln u  u 2  a 2 
y  u  du  y dz
Thus, making
a  y0

Vertical equilibrium requires: 


ln y  y 2  y02   z
y0
C

Catenoid
y
Denoting: 
y0

We successively write:  
ln y0    2  1    yz  C
0


ln y0  ln    2  1   z
y0
C


ln    2  1   z
y0
 C  ln y0 
z
y0
 C*

Remembering the inverse hyperbolic relationship:

 
ln    2  1  arcosh  

 y  z
We have: arcosh     C*
 y0  y0

20
26/10/2017

Catenoid
y
Also denoting: 
y0

We successively write:  
ln y0    2  1    yz  C
0


ln    2  1   z
y0
 C  ln y0 
z
y0
 C*

Remembering the inverse hyperbolic relationship:

 
ln    2  1  arcosh  

 y  z
We have: arcosh     C*
 y0  y0

 z 
That is: y  y0 cosh   C * 
 y0 

Catenoid
Imposing the boundary condition

y  0   y0  y0 cosh  C *   cosh  C *   1  C *  0

 z 
y  y0 cosh  
 y0 

That is, the generatrix curve of the minimal surface connecting two circular rings is a
catenary, and the surface itself is the catenoid!

h h  R 
Defining the ratio  we have y    y0 cosh  R
R  
2  2 y0 

Which provides y0 for a given ratio 

It can be shown that the only surfaces of revolution which are also minimal surfaces are the
catenoids (Struik, 1950), as it is easy to conclude from the above development , since no
particular restrictions have been set for y  y  z 

21
26/10/2017

Goldschmidt limit
The maximum ring separation for a catenoid with equal lower and upper radiuses is
h  hlim  1.3254868 R

This is known as Goldschmidt limit (1831).


Carl Wolfgang Benjamin Goldschmidt, ‘Determinatio superficiei minimae rotatione curvae data
duo puncta jungentis circa datum axem ortae’, Gottingae, MDCCCXXXI.
1
ELEMENTS 1
AVG ELEMENT SOLUTION MAR 13 2011
12:54:40 MAR 13 2011

h  1.32 R h  1.33 R
STEP=1 12:54:22
SUB =1
TIME=1
RATIO (AVG)
DMX =.660E-04
SMN =1
SMX =1
MX

MN

h
Z
Z
X Y
X Y

2R
1
1

A family of conoids:

22
26/10/2017

Ex. 9- Catenoid
Analytically determine the area of a catenoid bounded by two parallel and coaxial
rings of radius R=1m, separated by a distance h=1m.Compare to numerical results
obtained by direct area minimization and DRM.

Consider different degrees of discretization and check convergence of the numerical


results to the analytical one.

Try to find the Goldschmidt limit, considering numerical models with separations just
below and above hlim=1.32R

Show that for distances h>1.056R, the area of the catenoid is actually greater than
the area of two independent circular discs bounded by the rings (the solution for the
minimum area problem with two separate discs is know as Goldschmidt discontinuous
solution).

23

You might also like