Você está na página 1de 18

Anisotropy in the Hardness of Single Crystals

Author(s): C. A. Brookes, J. B. O'Neill and B. A. W. Redfern


Reviewed work(s):
Source: Proceedings of the Royal Society of London. Series A, Mathematical and Physical
Sciences, Vol. 322, No. 1548 (Mar. 23, 1971), pp. 73-88
Published by: The Royal Society
Stable URL: http://www.jstor.org/stable/77860 .
Accessed: 07/05/2012 02:14

Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .
http://www.jstor.org/page/info/about/policies/terms.jsp

JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of
content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms
of scholarship. For more information about JSTOR, please contact support@jstor.org.

The Royal Society is collaborating with JSTOR to digitize, preserve and extend access to Proceedings of the
Royal Society of London. Series A, Mathematical and Physical Sciences.

http://www.jstor.org
Proc. Roy. Soc. Lond. A. 322, 73-88 (1971)
Printed in Great Britain

Anisotropy in the hardness of single crystals


BY C. A. BROOKES,t J. B. O'NEILLI AND B. A. W. REDFERN?

tDepartment of Engineering Science, University of Exeter,


tMaterials and Chemistry Department, Rolls Royce and Associates, Derby,
and ?Jaterial Science Section, British Railwavas Board, Derby

(Communicated by D. Tabor, FR.S-Received 7 August 1970)

[Plate 1]

The results of this work, and those published by other researchers who have used Knoop
indentation measurements, confirm that the nature of anisotropy in hardness is essentially
determined by the crystal structure and the primary slip systems which accommodate
dislocation motion during indentation. Materials belonging to the same class of crystal
structure and having common slip systems possess similar anisotropic properties. The
varying extent of work-hardening or fracture, associated with indentations, does not
appear to influence the anisotropy-although twinning on the basal planes of hexagonal close-
packed metals may have a significant effect.
An analysis of the indentation process is presented which establishes a clear relationship
between the 'effective resolved shear stress' (r'), in the bulk of the crystal beneath the
indenter, and the observed hardness. Directions which correspond to the minimum values
of Tr, on specific crystallographic surfaces, are those of maximum hardness and conversely.
The analysis is shown to be equally applicable to a wide range of crystalline solids including
nonmetallic materials, of various crystal structure, and typical f.c.c., b.c.c. and c.p.h.
metals. Finally, anisotropy in hardness can be used to identify active slip systems in those
crystals where it is possible for dislocations to move on more than one system.

INTRODUCTION

Anisotropy in the hardness of crystalline minerals was firmly established during


the original development of the Knoop hardness test (Knoop, Peters & Emerson
1939; Winchell I945; Thibault & Nyquist I947). Great care was taken in that early
work to regularize the use of this type of indenter and to ensure comparable
results from various. sources. Subsequently, a number of measurements were
reported but no attempt was made to analyse the process of indentation, with
a view to understanding the nature of anisotropy, until the work of Daniels &
]Dunn (I949). They developed an explanation based on the distribution of 'effective
resolved shear stresses' due to tensile stresses whose axes were assumed to lie
parallel to the steepest slope of each facet on the indenter. The anisotropic be-
haviour of iron + 3 0 silicon crystals could then be explained fully but not that
of zinc-the only other crystal included in their work.
Later Feng & Elbaum (i958) proposed an alternative model, again based on the
effective resolved shear stresses, but they assumed that compressive stresses with
axes normal to the indenter facets controlled the deformation. Also, Garfinkle &
Garlick (I968) showed that, for lithium fluoride and a number of both face-centred
[ 73 ]
74 C. A. Brookes, J. B. O'Neill and B. A. W. Redfern
and body-centred cubic crystals, the hardness was dependent only on the crystal-
lographic orientation of the long axis of the inidenter and not on the plane of
indentation. However, the plane of indentation made a noticeable contribution
to hardness anisotropy in the case of an hexagonal close-packed crystal of titanium.
They concluded that the resolved shear stress analyses, whether based on tensile
or compressive forces, did not satisfactorily explain the observed anisotropy in
the hardness of crystals. Subsequently, Alexander & Carlson (I969) used the
results they had obtained on chromium and vanadium crystals to support Garfindle
& Garlick's observations on body-centred cubic metals.
Bowden & Brookes (I966) have utilized Knoop hardness measurements on
magnesium oxide and lithium fluoride crystal surfaces, to augment their studies
on the frictional anisotropy of nonmetallic crystals. Originally, we had intended to
use Knoop indentation hardness measurements in a similar way to extend that
earlier work. However, further consideration of our results on the anisotropy in
the hardness of non metallic crystals encouraged us to attempt an explanation of
this phenomenon in a much wider range of crystalline materials.

EXPERIMENTAL
The hardness measurements were made with a Leitz micro-hardness machine
using a Knoop indenter and following the general principles established by
Knoop et al. (I939), Winchell (I945) and Thibault & Nyquist (I947). The Knoop
indenter is a pyramidal diamond where the included conical angles subtended by
the longer and shorter edges, are 172? 30' and 130? respectively. It develops an
indentation having the shape of a parallelogram in which the longer diagonal is
about seven times the length of the shorter diagonal. The long diagonal of the
indenter was alined along a low index direction, on the plane of the crystal, and
measurements were made for incremental rotations of ten degrees on the specimen.
Normal loads between 300 and 1000 g were used to minimize the Rebinder effect
known to influence this type of measurement at lower loads (Westwood, Goldheim
& Lye I967, i968; Hanneman & Westbrook I968). Some crystal surfaces, corre-
sponding to the (001) planes of magnesium oxide and lithium fluoride and the
(111) planes of calcium fluoride, were freshly cleaved before indentation. Other
low index planes, of these crystals and aluminium oxide, were prepared by con-
ventional metallographic polishing techniques and chemical polishing. These
various methods of preparing the surface planes, under the given experimental
conditions, did not influence the measurements for the crystals used in this work.
However, the loading cycle was strictly controlled to ensure a constant time of
12 s under full load and thereby avoid creep effects observed by Mulhearn & Tabor
(I960); Westbrook & Jorgensen (I965), and Atkins, Silverio & Tabor (i966). This
aspect was of particular significance with crystals of calcium fluoride and lithium
fluoride (O'Neill 1970).
Anisotropy in the hardness of single crystals 75

HARDNESS RESULTS

Magnesium oxide and lithium fluoride


Anisotropy in the hardness of these crystals, both having rocksalt type lattice
structures, was comparable in nature. The periodicity of the curves, shown in
figures 1 and 2, is similar on both the (001) and the (110) planes, although the
hardness values of these two materials differed substantially. Minimum hardness
on the cube plane was found in the <100> and maximum hardness coincided with

Elool [1
E1 o
E010 10
Eii 10Eoo 1000_ Eool] 1 0 1 iol
I I I~~~~~~~~~~~~~~~~~~~~to
800 - (a) (

}600 tq00 y o/i\ SJ i (a) R~~~~~~~~~~


F
10 11 > | X ? RO(a

I I I I 4000OX
'~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~I
I
I I I , I_I I I
00 600 1200 1800 00 600 1200 1800

'A [ioo] E
[11 [0Eo10 [110i] [100LJoolIl Lull ETi
oll t11? looij
P-~10
- I II I I I I 0 1
400 1 0- I_OI2 \ I
nesium oxide d (b) lithiu 1(bd 'Id 0
I I 010 1 0 Q d
0 I~

0 on\I 110
FIUR 1 FIGUR 2
ta,ls. For example,onthe(001planeofmanesiumoxd100 t an

00 600 1200 1800 00 600 1200 1800


azimuthal angle
FIGURE I FIGURE 2

]FIGURE 1. Similarities in the riature of anisotropy ini hardness of (001) planes of (a) mag-
nesium oxide and (b) lithium fluioride.
]FIGURE 2. Comparable anisotropy in the h-ardness for (110) planes of magnesium oxide and
lithium fluoride.

tihe <110> directions. Likewise, on the (110) planes of both crystals the minimum
liardness was in KOOlI> directions and the maximum in KIllI>; the KIIlO>directions
bad hardnesses of intermediate value between those of KOO I> and <I fIi>.
The nature of anisotropy is the same for both magnesium oxide and lithium
fluoride, but the degree of that anisotropy is markedly different for the two crys-
tals. Fo-r example, on the (001) plane of magnesium oxide the hard and soft
dtirections differ by a factor of 2-i.e. 400 kg/MM2 and 800 kg/MM2 in <iOO>and
76 C. A. Brookes, J. B. O'Neill and B. A. W. Redfern
(110> directions respectively (1 kg = 9.8 N). The difference in hardness for these
two directions on a similar plane of lithium fluoride is only 1000.

Calcium fluoride
Anisotropy on the (001) plane of the fluorite structure reflected its fourfold
symmetry as in the case of the rocksalt structure crystals. However, figure 3

liool 7101 EL,0 [ool (a) o


0o E ono0[1 [1 oiolj
1 11201
g XI I I I I I
0 10 (a)a (
180 o 60 - i I SO I I
5 I l c l
0 01tI 40 I\ ' 0' 1
170 ~~~~~~~~ 0f

?10 0 11261E)ololE12o 100 20]1010[o-]Ch20]

150 ? o20 k 1200 0 2


0 Ioo ang
I I
3.
160GURE0 118 300 1 )p 8ea
6 170 [1120] [o] 12] [o] [n ]

0 170 051 01 I 0
IQI / I II I 01I
165b)6 01) 18ne 10 1 0Y
10
(b) 0\J][11
CDs(11
011
lavaepae
12 ,Ol[11 l100 1? oI

0~~~~~~~
16 0 1 I I I 0 I
msim hardness(182
I g/mm2) in the <1400 Ii T 0
1.60-1 0 60 I
10~ 180 010 24130

azimuthal angle
FiGuRE 3 FiGuRE 4
FIGureu 3. Anisotropy in the hardness of calcium fluoride crystals on (a) the (001) plane and
(b) the (111) cleavage plane.
FiGURE 4. Anisotropy in the hardness of alUMiDiuimoxide crystals on (a) the (0001) plane and
(b) the (1100) plane.

shows that the minimum hardness (158 kg/MM2) was now in the (110> and the
maximum hardness (182 kg/mM2) in the <100> directions. This is the converse of
the results obtained in the crystals of magniesium oxide a-nd lithium fluoride. The
(111) cleavage plane showed threefold symmetry in hardness results. The (iTO>
were the hardest directions (165 kg/mm2) with the <112> some 5 o less in magni-
tude (157 kg/mm2).
Brookes et al. Proc. Roy. Soc. Lond. A, volume 322, plate 1

*I : I

.1~~~~~~~~~~~~I

FIGURE 5. Typical indentations in (100> and (110> directions on a (001) plane of magnesium
oxide. (Unetched; magn. x 150.)
FIGURE 6. The distribution of dislocation etch pits around indentations on the (001) plane
of magnesium oxide in the principal directions. (Etched; magn. x 150.)
FIGURE 9. An indentation, on the (001) plane of magnesium oxide, with two of the indenter
facets alinod parallel to some {1 10) planes normal to the surface. Notc that there are no
dislocation etch pits ogi those particular slip planes adjacent to these two facets but the
other two facets, which are 16' off (110>, have produced dislocations on those planes.
(Etched; magn. x 150.)
(Facing P. 77)
Anisotropy in the hardness of 8ingle crystal8 77

Aluminium oxide
Anisotropy in the hardness of aluminium oxide crystals is shown in figure 4.
The prismatic (1100) plane had a maximum hardness (1800 kg/mm2) in the [0001]
direction and a minimum hardness (1380 kg/mm2) in the <1120> directions. The
anisotropy was not so pronounced on the basal (0001) plane but the <1120> were
approximately 500 harder than the [1100] direction.

DEFORMATION OF THE CRYSTALS

Both magnesium oxide and lithium fluoride have {110} <110> slip systems.
Conseq-uently, the nature of deformation in the region of indentations, for a given
plane and direction, was similar. The increased tendency towards fracture asso-
ciated with indentations on magnesium oxide in <100> directions, on both (001)
and (110) planes, was the only significant difference in the two materials. Figure 5,
plate 1, shows typical indentations in the two principal directions on an unetched
(001) surface of magnesium oxide. Dislocation etch pits in characteristic rosette
patterns are revealed by etching and are shown in figure 6, plate 1, for indentations
in <100> and <110> directions on this surface. The action of the etchant has
exaggerated the degree of cracking associated with indentations in <100> directions.
However, the presence of such cracks did not interfere with the reproducibility
of the hardness measurements.
Evans, Roy & Pratt (I966) have studied the deformation of calcium fluoride.
They showed that this material was normally brittle, under conditions of uniaxial
stress, at room temperature. Bulk plastic flow did not normally occur below
experimental temperatures of 60 'C when slip was activated on the {001} <110>
primary slip systems. Slip was also observed on {1 10} <110>, as secondary slip
systems, at temperatures above 200 'C. In our experiments the indentations were
of a ductile nature, on both the (001) and (111) planes, for all crystallographic
directions. Observations on etch pits, around the indentations, confirmed that
dislocations had moved on the {100} <011> slip systems.
Aluminium oxide is also normally brittle at room temperature. Primary slip
has been observed on {0001} <1120> slip systems above 1800 'C (Conrad I965).
Limited dislocation mobility has also been reported in several friction experiments,
nominally at room temperature, by Steijn (I96I), Reisz & Weber (I964), Eiss &
Fabiniak (I966) and Buckley (I966). We have observed perfectly formed ductile
indentations in [0001] directions on the prismatic (i100) plane. However, con-
choidal fracture occurred around the indentation, often as much as 5 s after the
removal of the load, when measurements were made within 30? of the <T120>direc-
tion on this plane. Subsurface fractures were always produced after indentation
on the basal plane. These fractures appeared to lie on basal planes approxi-
mately 9 lm below the indented surface. Again, the hardness measurements
were not significantly influenced by the formation of these cracks. Dislocation
78 C. A. Brookes, J. B. O'Neill and B. A. W. Redfern
etch pit studies, on the basal planes of aluminium oxide, revealed 'grown in'
dislocations but no evidence of glissile dislocations.

DISCUSSION
The results presented in this paper, together with those previously published
by Daniels & Dunn (I949), Partridge & Roberts (I963), Garfinkle & Garlick (I968),
Petty (1962), and others, confirm that anisotropy in hardrness is an intrinsic
property of all crystalline solids. Clearly, the use of the Knoop indenter facilitates
the study of this phenomenon. However, indications of anisotropy in hardness
have also been observed around indentations produced by a Vickers pyramidal
indenter on planes of metal crystals by Partridge & Roberts (I963) and Petty
(i962). They explained the irregular shape of the resultant indentations on the
basis of topographical changes determined by the crystallographic nature of plastic
flow and twinning.
Emergence of dislocations on the surfaces of the crystals used in this work
formed slip steps which were of insufficient size to markedly change the surface
topography. Nevertheless, the interrelationship between operative slip systems
and the measured anisotropy in hardness suggests that a possible explanation
should be based on the behaviour of dislocations and bulk plastic flow. For
example, the same type of anisotropy is observed for crystals which have slip
systems in common-such as magnesium oxide and lithium fluoride. On the other
hand, the nature of anisotropy is reversed for calcium fluoride where slip occurs
on a different system. Therefore, in the following discussion we shall be concerned
with those earlier explanations of anisotropy which were based on the distribution
of shear stresses in the bulk of the crystals. We shall also restrict this discussion
to the nature of anisotropy and not the magnitude of hardness.
Let us first consider the deformation of a cylindrical crystal under a unidirec-
tional tensile stress. Slip on the primary slip system, in the early stages of plastic
deformation, causes the crystal to become elliptical in cross-section. This is due to
the rotation of the slip planes about an axis lying in those planes and normal to
the slip direction. The resolved shear stresses, tending to plastically deform the
crystal, are then governed by the well known Schmid & Boas (I950) equation:
T (F/A) COSAcos 0, (1)
where T is the resolved shear stress, F = the applied force, A the cross-sectional
area of specimen, A the angle between the stress axis and slip direction, and b the
angle between the stress axis and the normal to the slip plane.
This equation can form the basis for an analysis of anisotropy in the hardness of
crystals when two further points have been established. First, the angle between
the axis of the stress responsible for deforming the crystal and the surface of
indentation must be determined. Values of A and 0, for a given orientation, can
then be calculated. Secondly, we must appreciate that now the available slip
Anisotropy in the hardness of single crystals 79
planes, unlike those submitted to a unidirectional stress, are not free to rotate
about a number of axes governed only by the slip directions in those planes. The
indenter and the hinterland of material, which is only elastically deformed, will
both influence the choice of slip system. Material between these two regions is
displaced, during the indentation process, from within the bulk on to the surface
of the specimen. Consequently, a slip system which allows rotation about an axis
parallel to an adjacent indenter facet will be more favourably oriented for slip
than one whose axis of rotation is normal to that facet.

slip ~~~~~~~~~s.d.

/ a.r.
FIGuRE 7. A schematic illustration to show the angles 0, A, f and y used in equation (3).
FF, Tensile axis; HH, axis parallel to indenter facet ABC; N, slip plane normal;
s.d., slip direction; a.r., axis of rotation in slip plane. Note that the angles between
FF and HH, s.d. and N and s.d. and a.r. are all constant and equal to 900. The angles
?4,A, ?I and y are determinedby indenter: specimen orientation and here ABC is the
relevant facet.

Daniels & Dunn (I949) suggested that the process of indentation might be
considered in terms of cylindrical elements of material, in the bulk of the crystal,
subjected to tensile forces parallel to the lille of steepest slope on the indenters'
-individual facets. Further, that the constraint opposing the rotation of slip planes,
during indentation, is governed by the cosine of the angle (3/) between the face of
an adjacent facet and the axis of rotation for a given slip system. The schematic
illustration in figure 7 shows the angles A, 0 and 3f for one cylindrical element
adjoining a given indenter facet. Initial slip on the primary slip plane, subjected
to the maximum resolved shear stress, causes that plane to rotate about an axis
a.r. Daniels & Dunn (I949) concluded that the constraint is minimal when the
axis of rotation is parallel to the indenter facet-i.e. along the direction H given
in figure 7. In this case, the angle ?fr= 0 and the constraint term is unity. They
further suggested that the constraint is a maximum when f = 90 and that
80 C. A. Brookes, J. B. O'Neill and B. A. W. Redfern
there is no slip because rotation of the slip planes cannot occur. Thus, their effective
resolved shear stress (Te) equation was developed:
Te = (F/IA) cOS coscoscos
0 . (2)
Clearly, values of F and A cannot be unambiguously defined in a hardness test.
Nevertheless, the product of the cosine factors may be used to demonstrate the
relative magnitude of resolved shear stresses for indentations in various crystallo-
graphic directions on a given crystal plane. Those directions having the lowest
effecti-ve resolved shear stress should be the hardest.
The application of equation (2) to (001), (110) and (I 1) surfaces of iron + 30
silicon showed that, on the basis of primary {112} K11l> slip systems, a cormpara-
tively small effective resolved shear stress was developed when the long axis of the
indenter was alined in directions of high hardness. Conversely, comparatively
large effective resolved shear stresses were produced in directions of low hard-
ness. A similar analysis explained their results on the (1J450)plane of zinc but, the
basal plane gave apparently anomalous results.
A modification of the explanation suggested by Daniels & Dunn was subse-
quently attempted by Feng & Elbaum (1958) and Garfinkle & Garlick (I968).
Instead of considering a tensile force parallel to the indenter facets as the effective
deformation force, these authors preferred to use a compressive force normal to
those facets. However, their analyses gave little agreement with the experi-
mentally established anisotropy in the hardness of crystals (Garfinlkle & Garlick
I968).
One other point must be clarified before we discuss our analysis of anisotropy in
the hardness of single crystals. Garfinkle & Garlick (I968) showed that Daniels
& Dunn (I949) omitted one of the {112} <11K>slip systems in the determination
of resolved shear stresses for iron + 3 00 silicon. When all the possible slip systems
of this type were included then they claimed that, for indentations on the (001)
plane, the minimum effective resolved shear stress occurred approximately 300
from <110>. This should then have been the direction of maximum hardness and
would have been contrary to both the observed hardness maximum and the
lowest resolved shear stress at 450 to <110> directions established by Daniels &
Dunn (1949). However, Garfinkle & Garlick (I968) had considered the resolved
shear stresses for only one facet of the indenter. All four facets must be considered,
to obtain a truly representative picture, and the mean value computed using only
the maximum value of resolved shear stress developed by each facet. Figure 8
shows the complete family of resolved shear stresses for indentations on a (001)
plane assuming {112} <II1> slip systems. The curve representing the mean of the
maximum values is also shown and confirms that the lowest resolved shear
stresses are obtained at 450 from <110>, i.e. <100> directions, and is consistent with
the observed hardness maximum for this crystal.
Further consideration of figure 7, however, together with the dislocation distri-
bution associated with indentations on the (001) plane of magnesium oxide shiown
Anisotropy in the hardness of single crystals 81
in figure 6, indicates that the constraint term of Daniels & Dunn (I949) is incom-
pletely defined. The resolved shear stress on the two {110} planes at 90? to the
indented (001) plane, i.e. (110) and (11 0), should be zero on the basis of equation (2).
This is because the axis of rotation is normal to the axis HH lying in the surface
of the indenter facets. Thus, cos Vt = 0 on these particular slip planes for all
orientations of the indenter on this (001) surface. However, the presence of some
dislocations on the (110) and (110) planes, as shown in figure 6, confirms that
there is a finite shear stress on all the {110} slip planes.

0.4 meanmr.

~0.3 -

0 g

0.2 -'

0? 400 80? 1200


azimuthal angle
FIGURE 8. The complete family of resolved shear stress curves, calculated from equation (2)
(Da,niels & Dunn 1949), for iron + 3 ?/ silicon crystals. Zero azimutha,l angle corresponds
to a <110> direction, on the (001) plane, and the minirnum effective resolved shear stress
(-,) to a <lO0> direction. The original mean resolved shear stress curve, due to Daniels
& Dunn (I 949), iS shown as the broken line c3ontinuation of our mean r, curve.

The above e-videnceestablishes that the maximum constraint is not defined by


3f =90' alone. Nevertheless, we can show that the rotation of slip planes does
determine the constraint term and hence the anisotropy in hardness. The only
other crystallographicrelation which could influence slip-plane rotation, anldnot
previously considered,is that between the slip direction in a given plane and the
adjacent facet. When this is taken into account, the maximum constraint is
obtained only where the slip direction and HH are coincident and #f is auto-
matically 90?. If the angle between the slip direction and HH, designated as y,
exceeds 0? then the slip plane can rotate even though Vfis 90?. On the other hand,
the minimum constraint is obtained when the axis of rotation and HH are coinci-
dent-when y must always be 90?. Thus the modifying function which reduces

6 VoI. 322. A.
82 0. A. Brookes, J. B. O'Neill and B. A. W. Redfern
the effective resolved shear stress, due to rotational constraint, is I(cos 3f+sin y).
The complete form of the effective resolved shear stress equation now becomes
r = I(FI/A) cos A cos h(cos f + siny). (3)
One significant aspect of this modified equation is that, whatever the crystallo-
graphic nature of the indentation, there will always be a finite resolved shear stress
0.4
0.3 [leo] [iTYo] [oio] Mi0] [0 03 [too] 1[TO] LoT6J 1
iTo Ffoo30
I I I I~~~~~I(a
I I (a)~~~~~~~~~~~~~~~~~~ a

Nz 0280? 160? 0.3 - 1\ Z1 X

0.2URE
1V FI E80 1600
~0.1 08110 0.3
to0.4
( (b)
io II 0 [l 00 1i Co
0.2systems
-7 (Top)
or (b {110 (lT>s (b)
Iol 11 [1 ICl [0I .
I I I ~~~~~
~~0.2 I
1
~0.3 I j I

Iof L 'J Eli l I 0

0O I _.
s21n s_Is
10 - 0__ 1o l a I i iI ui o (2)
0 ~~~800 1600 00 80" 1600
azimuthal angle
FIGURE10 FIGURE 11
10. The mean effectiveresolvedshear stress (r') curves, calculatedfrom equation (3),
FiGURet
for rocksalt
modified type
constrintfco str-uctureswith
a {I I10}(110TO>
hce by slip systems. The
indetin surfaces corresponld
crystalsufaceof
[to] (00o) [agn0]
to (a) the (001) and (b)the (110) planes.
FiGuRE 11. -r'curves for the (001) plaae of fluorspar crystals for slip either on (a) {001} (110>
systems (Top) or (b) {110} (110o>systems.

on slip systems in crystals. This overcomes the limitation implicit in equation (2)
and explains the dislocation behaviour apparent in figure 6. The validity of the
modified constraint factor was checked by indenting the (001) surface of magne-
sium oxide with the indenter alined so that HH for two facets was parallel to a plane
of the type (110O).The effective resolved shear stresses on (11I0)planes adj acent to
these facets should then be zero since i/r = 900 and y 0O. Figure 9, plate 1,
confirms this because, while there are prominent bands of dislocation due to the
two facets 160 off the [1101 direction, there are no dislocations on the (11-0)planes
adjacent to the two facets parallel to those planes.
Anisotropy in the hardness of single crystals 83
The validity of equation (3) can be substantiated by its application to a wide
range of crystalline materials. First, we shall consider the results obtained in this
work. The mean effective resolved shear stress curves for rocksalt type crystal
structures, having {110} <110> slip systems, are shown in figure IOa, b for the (001)
and the (110) planes of indentation respectively. Comparison of these curves with

0.4

P04 Ao; j 0
0 [1120] Loi[ Ln1o1 11o001

0.2 -

0.2 -
(0.4- o (b)

id 0. l 3l
O - 1 ! 1 I *_ i LL_ I I _I I I
0? 400 800 1200 1600(> 40? 0 120?
0. 1

0.2 I~~~~~~~~~~~~~~.

C or I {1210}
I (b)
00
I aI slip
syI41010>I I systems.~
I I it ~ ~ ~
{11}I ~<10>sli_sytem._Tes
~ ~ ~~~(b 800 1200
400 800 1200 16O ( 400
pea in Ltheo[11fietoInte azimuthal
10 angle
ln,aepe ictd ot onl1npos0] o
FIGURE12 FiGTUREP13
FiGURE 12. -r curves for the (iT1o) plane of hexagon al crystals on the basis either (a) of
(0001) (1120> or (b) of {1210} (10_10) slip systems.
FiGURE rT13. curves for the basal plane of hexagonal crystals during slip on (a) (0001) (1120>
or (b)
{1b10} (1010) slip systems.

-the results presented in figures 1 and 2 confirm that hardness maxima correspond
-with effective resolved shear stress minima. Details, such as the small hardness
peak in the [I11] direction on the (110) plane, are predicted not only in position
'but also in comparative magnitude. Figure 11 concerns indentations on a (001)
-plane of calcium fluoride and shows one resolved shear stress curve for slip on
{001I}<110I > slip systems and another for slip on {II0} <1I10>slip systems. These
curves indicate converse anisotropy for one set of slip systems compared with the
other. The hardness results shown in figure 3 a, together with dislocation etch pits
observed on {100} planes, confirm that the {001} <110> slip systems control the
6-:2
84 C. A. Brookes, J. B. O'Neill and B. A. W. Redfern
indentation process in this crystal at room temperature. A similar effect is apparent
in the effective resolved shear stress curves for the (1100) plane of aluminium oxide
and is shown in figure 12. A comparison of these curves with the hardness results,
given in figure 4, indicates that slip on {0001} <1120> systems controls the hardness.
Figure 13 shows that the degree of ainisotropy on the basal plane, indicated by the
effective resolved shear stress curves for both possible slip systems, is predictably
small. The evidence for slip in the basal planes is augmented if we consider the
relevant values of T' and hardness on the two planes. For example, the hardness

TABLE 1
Knoop hardness number

(001) (110) (111)


slip
crystal structure (100> (110> (001> (111> (110> (110> (112> systemt references
MnS rocksalt 122 142 119 142 142 140 140 {110}(1I0> Riewald&VanVlaek(i969)
MnS rocksalt 164 183 162 194 196 162 160 {110} (110> Moore & Van Vlack (I968)
LiF rocksalt 87 93 87 97 93 93 -- {110} <110> Garfinkle & Garlick (I968)
LiF rocksalt 96 103 98 120 116 - {110}( 110> this work
NaCl rocksalt 18 20 - { - {110 (10> O'Neill (I 970)
MgO rocksalt 400 800 - - - - {110} (110> Bowden & Brookes (1966)
MgO rocksalt 400 780 420 930 810 - - {110} (1I0> this work
Muo rocksalt 252 285 252 287 287 - {110} <110> Moore & Van Vlack (I968)
CaF2 fluorspar 178 157 - - - 164 158 {100} (011> this work
Al f.c.c. 23 18 23 17 18 18 - {111}(1I0> Garfinkle & Garlick (I968)
Al f.c.c. 18 14 22 15 16 17 18 {111} (110> Petty (i962)
Ni f.c.c. 105 72 115 84 84 93 - {111}( 110> Vecchia & Nicodemi (i965)
MnSe f.c.c. 65 43 73 54 48 57 57 {111}( 110> Riewald&VanVlack(i969)
Cu f.c.c. 47 34 - - - - - {111}<110> O'NeillQ(970)
C diamond 9600 6900 - - - {111} 110> Brookes (970)
cuLbic
W b.c.c. 445 375 440 360 360 390 408 {110}, {112},Rieck, Vaessen & Vogel
{123}, (111> (I968)
W b.c.e. 409 337 409 343 337 337 - {123}, <111> Garfinkle & Garlick (i968)
Nb b.c.e. 81 59 81 63 59 59 - {123}, <111> Garfilnkle & Garlick (I968)
Fe(Si) b.c.c. 229 183 240 196 203 198 214 {123}, (111> Daniels & Dunn (1949)
V b.c.c. 103 79 97 - 78 79 92 {123}, (111> Alexander & Carlson (i969)
Cr b.c.c. 139 108 159 - 115 108 123 {123}, (111> Alexander & Carlson (I969)
t Slip systems thought to control indentation process.

onthe (1100) plane is 1800kg/mm2 whenT,' 0.10 and 1400kg/mM2 when r 0. 35.
The hardness on the basal plane is 1300 to 1400 kg/mm2 and is more consistent with
the value of r' = 0. 34 to 0.38, due to slip on basal planes, than 4r = 0.12 to 0.15
as given by slip on the {1210} planes. The absence of dislocations, around indenta-
tions on the basal plane and discussed earlier in this paper, would also be consistent
with slip on the {0001} <1120> systems. Any dislocations activated in these sys-
tems by the indentation process would mo-ve parallel to the indented surface and
would therefore not emerge or intersect that surface. This might also account for
the subsurface cracks, observed on the basal plane, after the mechanism proposed
Ani8otropy in the hardne88 of 8ingle cry8tal8 85
by Bullough (I964) and Gilman (I958). That mechanism relies essentially upon the
creation of a crack on the slip plane due to a 'pile-up' of edge dislocations on that
plane. It has been observed in similar situations, with magnesium oxide crystals,
by Brookes & O'Neill (I968). An extensive survey of the literature was made to
collate all the available data on anisotropy in hardness testing by the Knoop
indenter. These are summarized in tables 1 and 2 and cover a wide range of
crystalline materials-both metallic and non metallic-with hardness values
varying from 13 kg/mm2 to approximately 104 kg/mm2.
Manganese sulphide, lithium fluoride, sodium chloride and magnesium oxide
all have the rocksalt type crystal structure and {1 0} K<1O>slip systems. Therefore,
figure 10 gives the relevant resolved shear stress curves, for (001) or (110) planes,
and the hardness values given in table 1 are consistent with these curves.

TABLE 2
Knoop hardnessnumber

(0001) (0110) (1120)


- h C- _ ( A slip
crystalstructure<(100><1120><0001><1120><0001><011O> systemt references
Ti h.c.p. 117 117 56 131 - - {1010}<1120> Feng & Elbaum (I958)
90 99 38 117 36 111 {10O} <(1120>Garfinkle& Garlick(i968)
115 115 55 132 - - {1010}<1120> Schwartz,Nash&Zeman
(I96I)
Zr h.c.p. 271 264 115 198 112 198 {1010}<1120> Rittenhouse & Picklesimer
(I964)
SiC hex. 2917 2954 2027 2700 2321 2755 {10I0} <1120> Shaffer (I964)
Mg h.c.p. 28 24 36 15 - - {0001} <1120> Partridge & Roberts (I963)
28 23 32 13 - - {0001} <1120> Schwartz, Nash & Zeman
(I96I)
Zn h.c.p. 18 15 44 24 - {0001} <1120> Daniels & Dunn (I949)
Co h.c.p. -- - 250 180 - - {0001} <1120> Morral (1958)
Mo2C hex. 1580 1540 - {0001} <1120> Vahldiek & Mersol (i968)
-
A1203 hex. 1300 1400 1800 1400 {0001}<1120> this work
- -

t Slip systems thought to control indentation process.

Figure 14 shows that minima and maxima on the resolved shear stress curves
for the (001) plane of face-centred cubic metals also match the hard and soft
directions on that plane. All face-centred cubic metals have {111} <Kio>primary
slip systems and those shown in table 1 have maximum and minimum hardness
in <100> and <110> directions, respectively, on both (001) and (110) planes.
Very recent preliminary work, prompted by the analysis presented in this paper,
shows that even the behaviour of diamond is consistent with cubic crystals which
slip on {111} <110> systems (Brookes 1970).
Slip may occur in body-centred cubic metals on {10}, {1 12} or {123} planes but is
always in <111> directions. Daniels & Dunn (I949) have shown that each of these
slip systems would predict the same anisotropy in hardness. The effective resolved
86 C. A. Brookes, J. B. O'Neill and B. A. W. Redfern
shear stress curve for indentations on the cube plane, assuming slip on {1To}<KIl>
systems, is shown in figure 15. Again this curve fits the observed anisotropy in
hardness for these crystals, with the minimum hardness in <110> directions and
the maximum hardness in <100> directions.
In hexagonal crystal structures there are commonly two sets of slip systemn,the
{0001} <1120> and the {1IO0o}<1210>, although slip and twinning have also been
observed on other systems in the metals. Correlations of anisotropy in hardniess,
for a number of close-packed hexagonal metals, have been investigated by Par-
tridge & Roberts (I963) and their results are included in table 2. However, they
did not analyse the anisotropy in terms of resolved shear stresses. In most cases
the observed anisotropy on the basal planes of hexagonal crystals is very small.
This is reasonably consistent with our results and the effective resolved shear
- 0.4 04
m o01 11[no] [lo0] [Eo1] Loi]

Eo30 [410] [700] [Elio] [olo] I

E0t3X 0.3- 1 / 1

?>0.21-2 : W I 0.2 - _
O
O~~~~~~~~~~~~~(
0r 80 160 0 80
azimuthal angle
FIGUnE 14 FIGURE15
FIGURE14. fe curves for (001) surfaces of f.c.c. metals, and diamond cubic crystals, with
{111} <1I0) slip systems.
FIGURE15. Te curves for (001) surfaces of b.c.c. metals with {110} <111) slip systems.

stress curves shown in figure 13. It is also possible that twinning interferes with
the application of equation (3). Partridge & Roberts (I963) observed that twinning
is a significant mode of deformation during indentations on the basal planes, of
close-packed hexagonal metals, but not on the prism plaanes. It is on these prism
planes, where slip predominates, that the most marked degree of anisotropy in
hardness is obtained. Furthermore, hexagonal crystals fall into two distinct
categories which appear to be governed by the predominant slip systems. Table 2
verifies that those crystals with (0001) <1120> slip systems are hardest in [0001]
directions and softest in <1120> directions, on the (0110) planes, whilst the con-
verse is true for crystals which slip on {1010} <1120> systems. The nature of this
anisotropy is strikingly predictable on the basis of the relevant shear stress curves
presented in figure 12.
Authors of recent papers on hardness mechanisms have favoured explanations
based on compressive stresses (Marsh I964; Hirst & Howse I969). If we assume
Anisotropy in the hardness of single crystals 87
that the effective resolved shear stresses are determined by compressive forces,
on axes normal to the Knoop indenter facets, the resultant r. curves from equa-
tion (3) give almost as good an indication of anisotropy in hardness as those based
on tensile forces. It can be seen from figure 7 that, for a given orientation of the
indenter, the constraint term is the same for both cases. However, there are
differences in the Schmid Boas term, and therefore in the calculated values of
r' which are sometimes significant. In figure 16 we show the reciprocal of the

tension

7e

~0.3 I
~3 Ii
-ooiU [fi~ilol10
;~~~~~~~~~~~~~~~~~~~i Eoo < I I I

I I ~ 0.2- II
*'~~~I j
either tensile orI/ azm eal
t anl
compressionl

C~~~~~~~~~ 0.1

Cy referenceto the reciprocalof themeasuredLoh thie [ureifbae] o oilfr


o
00 ~8O0 1600 O0 8O" 160"
azimuthal angle
FiauR.E16. Effective resolved shear stress curves, for a (110) magnesium oxide surface, using
either tensile or compressive forces in conijunction with equation (3). It is clear that,
by reference to the reciprocal of the measured hardness, the curve based on tensile forces
is more consistent with the observed anisotropy.

measured hardness, for a (110) magnesium oxide plane, together with a resolved
shear stress curve derived for compression and one for tension. Clearly, the curve
based on tensile forces is more consistent with the observed anisotropy in hardness.
Finally, we should summarize some of the more important points that have
emerged from this study. The nature of anisotropy in the hardness of crystals
can be directly related to the effective resolved shear stresses, by using equation (3),
for a wide range of crystalline solids. This relation is equally applicable to the
results obtained in this work and those of earlier researchers. The extent of work-
hardening or fracture, associated with the indentation process, does not appear
to influence t:he anisotropy although twinning on the basal planes of hexagonal
close packed metals may have a significant effect. These observations indicate
that the mechanism controlling anisotropy, but not necessarily the absolute
value of hard-ness, is due to the behaviour of dislocations during primary slip in
the crystal beneath the indenter.
88 C. A. Brookes, J. B. O'Neill and B. A. W. Redfern
The authors would like to thank Professor D. L. Smare and the University of
Bradford for a maintenance grant to one of us (J. B. O'N.), the S.R. C. and
U.K.A.E.A. for their support. They are also grateful to Dr D. Tabor, F.R.S.,
Professor H. G. Edmunds and Dr H. K. Zienkiewicz for helpful discussion and
Miss Jill Kent for technical assistance.

REFERENCES
Alexander, D. G. & Carlson, 0. N. I969 Trans. TMS-AIMfE 245, 2592.
Atkins, A. G., Silverio, A. & Tabor, D. I966 J. Inst. Metals. 94, 369.
Bowden, F. P. & Brookes, C. A. I966 Proc. Roy. Soc. Lond. A 295, 244.
Brookes, C. A. I970 Nature, Lond., 228, 660.
Brookes, C. A. & O'Neill, J. B. I968 Proc. Symp. Anisotropy in Refractory Compound Single
Crystals, vol. 2 (eds. F. W. Vahldiek and S. A. Mersol), p. 291. New York: PlenumnPress.
Buckley, D. H. I966 NASA. TMX-52211.
Bullough, R. I964 Phil. Mag. 9, 917.
Conrad, H. I965 J. Am?,.Ceram. Soc. 48, no. 4, 195.
Daniels, F. W. & Dunn, C. G. I949 Trans. Am. Soc. M/+etals41, 419.
Eiss, W. & Fabiniak, J. I966 Am. Ceram. Soc. Meeting, Philadelphia, paper 7-C-65.
Evans, T., Roy, C. & Pratt, P. I966 Proc. Br. Ceram. Soc. 6, 173.
Feng, C. & Elbaum C. I958 Trans. TMS-AIME 212, 47.
Garfinkle, M. & Garlick, R. G. I968 Trans. TMS-AIMIE 242, 809.
Gilmian, J. J. I958 Trans. TMS-AIME 212, 310.
Hanneman, R. E. & Westbrook, J. H. I968 Phil. May. 18.
Hirst, W. & Howse, M. G. J. W. I969 Proc. Roy. Soc. Lond. A 311, 429.
Knoop, F., Peters, C. G. & Emerson, W. B. I939 J. Res. Nat. Bur. Stdc.23, 39.
Marsh, D. M. I964 Proc. Roy. Soc. Lond. A 279, 420.
Moore, J. W. & Van Vlack, L. H. I968 Proc. Symp. Anisotropy in Refractory Compound
Single Crystals, vol. 1 (eds. F. W. Vahldiek and S. A. Mersol), p. 220. New York: Plenum
Press.
Morral, F. R. I958 J. Metals 10, 662.
Mulhearu, T. 0. & Tabor, D. I960 J. Inst. Metall. 89, 7.
O'Neill, J. B. I970 Ph.D. Dissertation, University of Bradford.
Partridge, P. G. & Roberts, E. I963 J. Inst. Metall. 92, 50.
Petty, E. R. I962 J. Inst. Metall. 91, 54.
Reick, G. O., Vaessen, G. H. G. & Vogel, D. L. I968 Trans. TMS-AIME 242, 575.
Riesz, C. H. & Weber, H. S. i964 Wear 7, 67.
Riewald, E. & Van Vlack, L. H. I969 J. Am. Ceram. Soc. 52, 370.
Rittenhouse, N. & Picklesimer, 0. i964 O.R.N.L.-P.-1165.
Schmid, E. & Boas, W. 1950 Plasticity of crystals, p. 105. London: F. A. Hughes and Co.
Schwartz, M., Nash, S. K. & Zeman, R. I96I Trans. TMS-AIME 221, 554.
Shaffer, R. i964 J. Am. Ceram. Soc. 47, 466.
Steijn,R. P. I96I J. Appl. Phys. 32, 1951.
Thibault, N. W. & Nyquist, H. L. I947 Trans. Am. Soc. Metals 38, 271.
VahJdiek, F. W. & Mersol, S. A. 1968 Proc. Symp. Anisotropy in Refractory Compound
Single Crystals, vol. 1, 2 (eds. F. W. Vahldiek and S. A. Mersol). New York: Plenum
Press.
Vecchia, A. L. & Nicodemi, W. i965 Metallurgia Italiana 9, 321.
Westbrook, J. H. & Jorgensen, P. J. I965 Trans. TMS-AIME 233, 425.
Westwood, A. R. C., Goldheim, D. L. & Lye, R. G. I967 Phil. M3ag.16, 505.
Westwood, A. R. C., Goldheim, D. L. & Lye, R. G. I968 Phil. Mag. 17.
Winchell, H. 1945 Am. Mineral. 30, 583.

Você também pode gostar