Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

404 P. Alexander, A. Charlesby and M.

Ross
Bawn, C. E. H. 1948 Chemistry of high'polymers. London: Butterworth.
Baxendale, J. H., Bywater, S. & Evans, M. G. 1946 Polym. Sci. 1, 237.
Bischoff, J. & Desreux, V. 1952 Bull. Soc. Chim. Beige, 61, 10.
Bischoff, J. & Desreux, V. 1953 J. Polym. Sci. 10, 437.
Breger, I. A. 1948 J. Phys. Chem. 52, 551.
Charlesby, A. 1952 Proc. Roy. Soc. A, 215, 187.
Charlesby, A. 1953 Nature, Lond., 171, 167.
Charlesby, A. & Ross, M. 1953 Nature, Lond., 171, 1153.
Day, M. J. & Stein, G. 1951 Nature, Lond., 168, 644.
Eyring, H., Hirschfelder, J. O. & Taylor, H. S. 1936 J. Chem. Phys. 4, 479.
Grassie, N. & Melville, H. W. 1949 Proc. Roy. Soc. A, 199, 14, 24.
Lawton, E. J., Bueche, A. M. & Balwit, J. S. 1953 Nature, Lond., 172, 76.
Meyerhoff, G. 1953 Naturwissenschaften, 40, 106.
Meyerhoff, G. & Schulz, G. V. 1951 Makromol. Chem. 7, 294.
Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

Ross, M. & Charlesby, A. 1953 Atomics, 4, 189.


Schneider, E. E., Day, M. J. & Stein, G. 1951 Nature, Lond., 168, 645.
Wall, L. A. & Magat, M. 1953a J. Chim. Phys. 50, 308.
Wall, L. A. & Magat, M. 19536 Modem Plastics, p. I l l , July.
Winogradoff, N. N. 1950 Nature, Lond., 165, 721, 1237.

The form ation of cracks as a result of plastic flow

B y A. N . S troh
H. H. Wills Physical Laboratory, University of Bristol

{Communicated by N. F. Mott, F.R.S.—Received 21 November 1953—


Read 25 February 1954)

The stresses round a piled-up group of dislocations are investigated with reference to the
initiation of a crack. A crack should form when the group consists of about 1000 dislocations
piled up under a stress of the magnitude occurring in a cold-worked metal. The stress system
due to the crack is obtained. The crack will have a stable equilibrium length, but this length
is likely to be determined by the amount of plastic flow which takes place round the tip of the
crack.

1. I n t r o d u c t io n

In his theory of the brittle fracture of amorphous materials, Griffith (1921) has
assumed that there are present pre-existing cracks; when a stress is applied, these
produce high stress concentrations at their ends which enable them to extend.
This theory appears to give a satisfactory account of the fracture of substances
such as glass; but it seems unlikely that pre-existing cracks can, in general, play
any important part in the fracture of a metal. It is therefore necessary, in this case,
to start the theory one stage further back and consider how cracks may be produced
in the unbroken material. The aim of the present paper is to consider in greater
detail a suggestion, made by Mott (1953), that the stress concentration round the
piled-up group of dislocations at the end of a slip line is sufficient to initiate a crack.
We start by considering the conditions under which a crack will form. First, we
assume that, in considering the formation or spread of a crack, only the component
The formation of cracks 405
of the stress normal to the plane of the crack need be considered (the law of critical
normal stress). The validity of this assumption has recently been queried by
Deruyttere & Greenough (1953). However, it has the advantage of simplicity and
so will be adopted in the absence of any well-established alternative law.
Further, the stress must locally approach the true fracture stress of the material ,
which Orowan (1948) has estimated to be of order one-tenth of the elastic modulus.
It is, however, not sufficient merely to specify the stress, at a single point, for if
this were so a single dislocation could produce a crack. The necessary additional
condition may be obtained by requiring that the formation of the crack is to be
accompanied by a decrease in the energy of the system. We treat the problem
two-dimensionally, and use Griffith’s expression (1921) for the energy. If a tensile
Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

stress cr is applied to a body containing a crack of length r, the direction of the stress
being normal to the crack, the energy of the crack is
—£7r( 1 — v)cr2r2IG + 2
where G is the rigidity modulus, v Poisson’s ratio, and y the surface energy per unit
area. This expression is negative provided
or2r>16yG ln(l —v). (1)

T able 1
surface 1 0 -11
energy, y rigidity, 1086
(dynes/cm) (dynes/cm2) (cm) a — y /b G

copper 1430 110 2-5 0-059


silver 1140 7-7 2-9 0-051
gold 1510 8-0 2-8 0-067

It is convenient to write y = abG, where 6 is the Burgers vector and a is a dimen­


sionless quantity. Values of the surface energy, according to Fisher & Dunn (1952)
of some face-centred cubic metals, are given in table 1. From the table we see that
for these metals a ~ 006. Equation (1) may now be written
ari >4[al7r(l —v)]i biG. (2)

2. The initiation of a crack


Koehler (1952) has pointed out that there are large tensile stresses near a piled-up
group of dislocations, and has estimated the stress at one or two points near the
group. In this section we calculate the stresses by a somewhat different method.
We then use these results in conjunction with the considerations of the previous
section to obtain the condition that a crack should be formed.
We suppose the piled-up group consists of n positive-edge dislocations; the
leading dislocation is locked in position, and the remainder, which are free to move
in the slip plane, are held in equilibrium by an applied stress —<r0. This array has
been studied by Eshelby, Frank & Nabarro (1951). We take an x-axis in the slip
plane, y-axis normal to the slip plane, with the origin on the locked dislocation,
and choose G6/47i(l —v) <x0 as the unit of length. The equilibrium positions of the
406 A. N. Stroh
free dislocations are then given by the zeros of the derivative of the nth
Laguerre polynomial, and the stresses due to the 1) free dislocation by

Pxx+Pyy=
(3)
Pxy Pxx Pyy)^ (2) + i (2),

where G(z)= 2cr0ln^f^(z) and 2 = x + iy .We use


imaginary parts of a complex quantity.
For n large, we may use the asymptotic expansion of £"n{z) due to Perron (1921;
cf. also Szego 1939):
££ n
' )z( = eiz ( —z)~*
Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

Here ( —z)~l and ( —2)* are to be taken real and positive when 2 is real and negative.
Equation (4) holds for points not on the positive real axis such that
4<
n
/l 2| j <^4w.

Now the distance between the two loading dislocations is of order 4/n, and the length
of slip plane occupied by the whole piled up group is L 0 = 4 so that (5) states that
the distance from the locked dislocation must be large compared with the distance
between the two leading dislocations, and small compared with L 0.
From equations (3) and (4), we find that the stresses due to the 1) free
dislocations are given by

l(Pxx +Pyy)l(0r = - 3 sin d\2r + 2


i(Pxx~Pyy)l°ro ~ —3 sin 0 (l +cos 20)\2r + (2 s in
lyxPv0 = 1 + 3(cos 0 —sin 0 sin 2 —(n/r)* (2 cos \0 —sin 0 sin §0)
where \z \ — r and arg z = n + O, terms of order n~^ have been neglected. In addition,
we have the stresses due to the locked dislocation and the applied shear stress.
These are ., , w _ . n,
t ( P x x + P y y ) l ° ro = 2sm 0/r,

U
Pxx-Pvi,)I<To = 4cos2

Pxyl = — 2 cos 0 cos 2 1.

From (6) and (7) the total stresses round the piled-up groups are

UPxx+Pyy)lv0 *= Sin 0/2 r+ 2(n/r)i sin ,


l(P xx-P yy)lv0 = sin 0 cos2 0\r + ( )*(2si
Pxyl Vo = c~os0 cos20/2r —

Now since r^> 1,tn


/ he terms in 1/r are small compared with those in (
neglecting the former the stresses round the piled-up group become

\{Pxx +Pyy)lv0 = 2(n/r)i sin ,


UPxx~Pyy)lv0 = *)r/n
(2( sin + sin cos §0), . (8 )

xy/v o=
P 2 cos |0 - s i n 0 sin§0)._
The formation of cracks 407
The normal stress er acting on the plane OP (figure 1) making an angle with the
slip plane is
\(Vxx + P vy)- UPxx-Pyy) cos 2 0 sin 20
— f (LJr)* (Tqin
s 0 cos \0. (9)
On differentiating equation (9) with respect to 0 we find that is a maximum when
cos 0= | so that 0 — 7b|°. The maximum value of a is

O'max. = (2/V3) (LJr)i (TQ. (10)


Comparing (2) and (10) we find the condition for a crack of length r to form with
decrease of energy dU
Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

in ordinary units L0 = Gbnjn(l- v) <r0,


so that (11) may be written ncr0 > 12 ( 12)

&___
X

F igure 1

Since the length r does not occur in (12) it follows that, if this condition is satisfied
for one length of crack, it will be satisfied for all lengths, and the formation of a
crack will be accompanied by a decrease of energy at all stages. It seems reasonable
then to take (12) as the condition of initiation of a crack.
If a = 0-06, (12) takes the simple form
nor0> 0-7G. (12a)
Now in a work-hardened metal, say copper, er0 may be of order 109dynes/cm2,
while G is about 1012dynes/cm2; (12a) then indicates that a crack should be formed
near a piled-up group of about 1000 dislocations. It is not unreasonable to expect
that such piled-up groups will occur; for with a Burgers vector of 2 x 10~8cm, the
motion of 1000 dislocations in a single slip plane will give a slip line of height
2000A, and steps of this height have been observed by Brown (1951) on aluminium.
We have assumed that the crack will be formed in a position such that the
tensile stress normal to it is a maximum, this was found to be at an angle of 70|°
to the slip plane for an isotropic material. Actually in a crystal the crack will be
formed along a cleavage plane, which, in general, will not coincide with the position
of maximum stress. In a face-centred cubic metal we expect that there will usually
be a suitable cleavage plane near the position of maximum stress, so that deviation
of the crack from the position determined by purely elastic considerations will be
408 A. N. Stroh
fairly small; the high symmetry of these crystals suggests that there are a number of
crystallographic planes on which cleavage can occur. In hexagonal metals, on the
other hand, there is a single well-defined cleavage plane, the basal plane. In this
case we do not expect fracture to occur when (12) is fulfilled unless the cleavage
plane is in a suitable orientation. There is, of course, not necessarily any relation
between the orientations of the slip and cleavage planes, for if dislocations pile up
on a grain boundary it is possible for the slip to have taken place on one grain, and
the fracture to occur in the other.
Instead of producing fracture, a piled-up group on a grain boundary may cause
slip to spread into the next grain by the generation of dislocations in the perfect
lattice in the manner described by Frank (1950). It is not clear when the stresses
Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

round a pile-up will be relieved by slip and when by fracture. Possibly it depends
on which crystallographic plane is most suitably orientated.

3. The stress system due to the crack: crack normal to slip


In this section and the next we calculate the stresses due to the crack. We suppose
that once the crack has formed the n dislocations of the piled-up groups will enter it,
so that the extra material at the one end will wedge the crack open giving it roughly
the shape shown in figure 2.

F igure 2.

We shall consider first a crack normal to the slip plane and later extend our
results to cracks meeting the slip plane obliquely. We take the origin at the mid­
point of the length of the crack, a;-axis normal to, and y-axis in the direction of,
the crack. As unit of length we take half the length of the crack. At large distances
from the crack the stress field will be that due to a dislocation with Burgers
vector nb at the point A (0 ,1). This stress system gives a normal stress over the
surface of the crack of amount p = D j ( l - y ) (13)
The formation of cracks 409
where D = nbG/2n(l —v). We must therefore add to the stresses due to the dis­
location a system of stresses which vanish at infinity, and which will annul the
stress given by (13) at the surface of the crack. We shall follow closely the treatment
of Sneddon & Elliott (1947), making a slight generalization which is required because
of a lack of symmetry in the present problem. As the system is symmetrical about
a: = 0w e need consider only the semi-infinite region >0; then we require a solution
of the elastic equations for which the stresses vanish at infinity and which satisfies
the following boundary conditions at x = 0:

(i) P x y ~ 0>
(ii) Vxx = ~T>l{l~y), i f —l e y < 1,
Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

(iii) u = 0 , if \y \> 1 .
Now consider the stresses
Tco
Vxx= J { M cos p y + fr(p) sin 1+ px) e~f>x dp, (14)

C00
Vyy = Jo {<f>(p)cospy(18)
+ i/r(p)8

I' 00
Pxy = XJ
0 silW ~ f(P) cos dp, (16)

where (j>(p) and tyip) are functions to be determined. It is easily verified that these
stresses satisfy the equilibrium equations

tyxx dpxy ^Pxy ^Pyy ^


- + - =0 and 0 -1- - = 0 , (17)
ox oyox
and also the compatibility equation
/ 02 02 \
(18)
\0 ^ + 0y2/ 0-
Hence they are the solution of some elastic problem. They lead to the displacements
1 f°°
Gu = —~ {<fi(p)cospy+ i/r(p) sin py} (2 —2v+px)e~ ?x dp Ip, (19)
*J 0
i r°°
Gv = 7: {<J>(p) sin py —i/r(p) cos py} {1 —2v —px)e~^xdplp. (20)
0
It is evident that the stress vanishes when x tends to infinity, and also the boun­
dary condition (i) is satisfied. Accordingly we have only to find <p(p) and fr(p) so
that conditions (ii) and (iii) are satisfied. Putting — 0 in (14) and (19) we obtain
the following equations for <j){p) and ijr(p):
T°° 'i
I {<j>{p) cos py + ft(p) sin py} dp = - D / { l - y ) (-lcycl),
(21)
r °° 1
{(f>{p) cospy + )p(stf in py} dp/p = 0 (12/1> 1)-
0
410 A. N. Stroh
On replacing y yb — y nd
a combining the resulting equa
the two pairs of equations
P 00

I <fi(p) cospy dp = - D / { l - (0 < 2/ < 1),


( 22 )
f* 00
I <f>(p)cospydplp = 0 («/>!)>

C 00

and I i!r{p) sin py dp = - D y l { l - y 2) (0 <«/< 1),


(23)
l*00
I f(p )s m p y d p ! p = 0 0/ > l ) -
Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

If the sine and cosine are expressed in terms of Bessel functions by the relations
J^(x) — (2jnx)isin x and J^(x) = (2lnx)icos x, th
cases of a pair of dual integral equations studied by Busbridge (1938). Her general
solutions leads, after some calculation, to
<f>(p)=
and rjr(p) = D co sp —DJ0(.)p(25)
Alternatively, it may be verified by direct substitution that these expressions are
solutions of (22) and (23).
In evaluating the stresses, it is convenient to introduce the function

O(z) = - J o { f {p) ~ i e~pz dp (^(z) > 0) (26)

1)
(27)
2 —i + (z2+ 1)*'
Then, from (14), (15), (16) and (26) we have, on taking 2 = x + iy,

MPxx+Pyy)

Pxyd~ ^i(Pxx y)P ^ 37, ^l(z).

These stresses must be added to the stresses due to a (giant) dislocation at the point
z —i. Now the first term of (27) leads to a system of stresses which exactly cancels
the stress field of the dislocation. Accordingly, the total stress field due to the crack
will be derived from the second term of (27).
Taking , , , ., 1 , •1
Iz I = r , | 2 - 1 | = rv \z + i I = r2,
and arg 2 = y, arg (2 - i) = y x, arg y2,
(see figure 2), the stresses due to the crack become

Pxx = - ^ ( r 1r2)~i s m l { x 1 + X2 ) + I)r2(r1r2)-^cosx sin ( y ~ f y i - f y 2), (28)


Pyy = - D{rxrz)-^ sin \( Xl + y 2) - Dr2(rxr2)-icosy
Pxy = D r \ r xr2)~*cosy cos (y - %Xi ~ fXa)- (30)
The formation of cracks 411
These stresses reduce to those due to a dislocation, at distances large compared with
the length of the crack; while near the ends of the crack the stresses vary inversely
as the square root of the distance from the tip of the crack.
The shape of the crack may be found from equations (18) and (19) by putting
x = 0 and using (23) and (24). The displacements over the surface of the crack are

uQ= Jw6^1 + ^ sin _1y j , = 0. (31)

As might have been expected (31) gives a maximum width of the crack at y 1
of nb, the total strength of the dislocations which have entered into the crack. The
Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

crack tapers down to zero width at y =—1.

4. Crack oblique to the slip plane


As the results of § 2 indicate that the crack will not in fact be normal to the slip
plane, it seems desirable to extend the results of the previous section to cracks
making an arbitrary angle 6 with the slip plane.
We again take x-and y-axes normal and parallel to the crack. Then t
Burgers vector nb of all dislocations which have entered the crack may be resolved
into components nb sin 6 normal, and —nb cos 6 parallel, to the crack. The stresses
associated with the normal components will be those considered in the previous
section; they are given by (28), (29) and (30) with D replaced by D sin#.
The component —nb cos 0 parallel to the crack produces a shearing stress over the
surface of the crack given by
Vxv = - D 00801(1- y ) ,
and we seek a system of stresses which will cancel this. If we consider the semi­
infinite region x>0, the boundary conditions on 0 are

(i) Pxx= 0,
(ii) VxV = D o o a 6 l ( l - y ) , if - 1 < y < 1,
(iii) v = 0, if \y>| 1.
Now the stresses
P xx = *J o { f l i p ) c° s py~i>i(p)sin (32)
f*O
O
P w = l{flip) cospy - ^ p ) sin py} e~Px { 2 - p x ) dp, (33)
f* 00

Pxy = - Jo s)pilf{ inp y + 0i cospy) (J ~ p x) dP (

satisfy equations (16) and (17), and vanish as x tends to infinity. Also boundary
condition (i) is satisfied. The corresponding displacements are
1 f°°
Gu = - \ { — f l i p )cos yoy -l- ^
i(p)inpy}
s (1 —

1 f 00
Gv =- { f M sin py + 0 1(p) cos (2 - 2
*J o

Vol. 223. A. 27
412 A. N. Stroh
Putting x = 0 in (34) and (36) we see that boundary conditions (ii) and (iii) a
satisfied if
\

(~l<y<l),
sinp y + $ i (p ) ) dP = —D cos 0/(1 - y )

(37)
{friip) smpy 4 $Mp) cospy}dp/p = 0 (|y|>i).

Comparing equations (37) and (21) we see that


!>
f<p)
( = 0(p) cos 0, ^i(P) = cos 0,

^(p) and ^(p) being given by (24) and (25).


Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

Then the stresses (32), (33) and (34) are equivalent to

UPxx+Pyy) = - cos

y ra + ii(fe -3 > « » ) = »(£i + * ^ Q<z)) co se>

where O(z) is given by equation (26). After adding the stresses due to the dislocation
— nbcos 0, the stresses become
p xx= - D r \ r xra)-» cos 0 cos (x - §Xi~§Xa) cos
p yv= - 2 D{rxr2)"i cos 0 cos |( y x 4- y 2) + Dr
2)~* cos
^ = £ > ( W 4c«s 0sin \ { x x 4- x%) +
(38)
The total stress system due to the crack may now be obtained by adding the stresses
(28) to (30) (multiplied by sin0) to (38). We obtain

Pxx = ~ D (ri **2)~4sin 6 sin ${Xi4- Xa) ~


Pyv = ~ D (ri ra)~*(2 cos 6 cos Mxx + Xa) + sin 0 sin £(;& +
4- Dr2{rxr2)~* co
Pxy = J>{rxr2)~4 cos 0 sin £(*14-y 2) + Dr2{rxr ^ cosx s

5. T he energy and equilibrium length of the crack


The contribution of the stress field to the energy is most easily found by making
a cut along the positive y-axis in the unstrained material, and giving the material
on one side of the cut a displacement nb = {nb sin 0, —nb cos 0) relative to that on
the other. Then the elastic energy of the crack is
r-R rR
W' = -£ n fe sin 0 2 /| {pxx)x=0dy 4 \nb cos 6L'\ ^ {pxv)x^ d y ,

where L' is the length of the crack in the z direction. As in the case of a dislocation,
it is necessary to introduce a finite cut off distance we assume that R is much
greater than the length of the crack (i.e. R^> 1). Now along the positive y-axis
the stresses (39) and (41) become
Pxx —— I )sin 0(y2—1)~* and p xy D co
formation of cracks 413
Hence the energy is W' = \n bD U In
or if the length of the crack is c, the energy is
„ r, n2b2G L '. 4
W = ---------- rln — , (42)
c
and is independent of the angle the crack makes with the slip plane.
To (42) we must add the surface energy of amount giving the total energy
n2b2GL' 4
W = ——----- -In— + 2
4n(l —v ) c
Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

This is a minimum when c = w26/87t(1 —v) a, (43)


and this will then be the equilibrium length of the crack. The corresponding value
of the energy is , 3ifare(I- v ) a R
cr> 4n(l —v) 1 n2b
The fact that this is a minimum of the energy has the important result that the
equilibrium of the crack (unlike that of the Griffith crack) will be stable.
The energy of the original piled-up group of dislocations (Stroh 1953) was
n2b2GL' e*(l
ei ~ 4n(l - v )
Hence the formation of the crack results in a decrease of energy of amount

AW W -W = in - W<T° (44)
gr- ” cr- 4n(l-

The requirement that AIT> 0 leads, to a close approximation, to the same


condition as (12) for the formation of a crack,
If v= ^ and cl = 0*06, the length of the crack, from equation (43), becomes
c~ w 26; (43 a)
with n = 1000, b ~ 10~8cm, this gives a crack of length 10~2cm. This result seems
rather large, as a crack of this size should grow by the ordinary Griffith mechanism,
producing immediate fracture of the material. However, if any plastic flow takes
place round the tip of the crack, the stresses here will be greatly reduced and further
growth of the crack prevented. Hence in a ductile material we expect cracks con­
siderably smaller than equation (43) indicates.
On the other hand, should all the dislocation sources be locked so that they cannot
respond to a stress in the short time it takes a crack to form, the growth of the crack
will be unimpeded and the crack will spread through the material; we suggest this
is the condition for notch brittleness. In iron the dislocations will be locked by a
Cottrell mechanism and will only move with difficulty at low temperatures so that
the material will then be brittle; at higher temperatures thermal fluctuations should
be sufficient to release the dislocations under the stresses prevailing and ductile
fracture will occur. These considerations, however, at present, have not been worked
out in detail.
414 A. N. Stroh
It is a pleasure to thank Professor N. F. Mott, F.R.S., for suggesting the problem,
and for a number of helpful discussions.

R eferences
Brown, A. F. 1951 J. Inst. Metals, 80, 115.
Busbridge, I. W. 1938 Proc. Lond. Math. Soc. 44, 115.
Deruyttere, A. E. & Greenough, G. B. 1953 Nature, Lond., 172, 170.
Eshelby, J. D., Frank, F. C. & Nabarro, F. R. N. 1951 Phil. Mag. 42, 351.
Fisher, J. C. & Dunn, G. G. 1952 Imperfections in nearly perfect crystals, p. 317. New York:
Wiley.
Frank, F. C. 1950 Symposium on the plastic deformation of crystalline solids, p. 89. Pittsburg:
Carnegie Institute of Technology and Office of Naval Research.
Griffith, A. A. 1921 Phil.Trans. A, 221, 163.
Downloaded from https://royalsocietypublishing.org/ on 02 June 2024

Koehler, J. S. 1952 Phys. Rev. 85, 480.


Mott, N. F. 1953 Proc. Roy. Soc. A, 220, 1.
Orowan, E. 1948 Rep. Progr. Phys. 12, 185.
Perron, O. 1921 J . reine angew. Math. 151, 63.
Sneddon, I. N. & Elliott, H. A. 1947 Q. Appl. Math. 4, 262.
Stroh, A. N. 1953 Proc. Roy. Soc. A, 218, 391.
Szego, G. 1939 Orthogonal polynomials, p. 193. American Mathematical Society Colloquium
Publications.

Some observations on the behaviour of liquids


between rotating rollers
B y W. H. B a n k s a n d C. C. M il l
ThePrinting, Packaging and Allied Trades Research Association,
Leatherhead, Surrey

(Communicated by Sir Geoffrey Taylor, F.R.S.—Received 10 December 1953)

[Plate 13]

Direct observations of cavitation in liquid films separating two rotating rollers are reported.
The speed of rotation and separation of the rollers at which cavitation sets in is consistent
with hydrodynamical theory. The onset of cavitation is sharp at a rarefaction of about
2 atm for castor and paraffin oils.

I n t r o d u c t io n

The behaviour of liquids in the region of closest approach (the nip) of two rotating
rollers is of importance in many industrial processes and is complicated by the fact
that the cylinders are rarely completely immersed in liquid, but there is just suffi­
cient liquid to form a very thin film around them.
In attempting to study these systems it was considered expedient in the first
instant to examine the ideal case of complete immersion, in order to avoid difficult
boundary conditions. The experiments were designed to provide direct evidence
of cavitation in the region of the nip and to test whether the limiting conditions

You might also like