Comminution PHD
Comminution PHD
\   .
=   >
  [2.5] 
K1 is the particle size below which all particles will by-pass breakage and go directly 
to  the  product,  and  K2  is  the  size  above  which  all  particles  will  be  broken.    These  two 
parameters  are  expected  to  be  functions  of  the  crusher  sets,  K1  being  dependent  upon 
CSS  and  K2  upon  OSS.    As  an  exponent  K3  describes  the  shape  of  the  classification 
function for particles between the K1 and K2 sizes (Figure 2.10). 
 
  Research  has  shown  that  KI  and  K2  are,  in  addition  to  closed  and  open  side  set, 
functions  of  feed  size,  throughput,  crusher  throw,  and  plate  liner  characteristics  with 
exact relationships being found empirically using actual operating conditions/parameters 
(Napier-Munn  et  al.,  1999;  Whiten  1984;  Karra,  1982;  Andersen,  1988;  Andersen  and 
Napier-Munn, 1990).  The same work has shown that K3 remains fairly constant among 
  16
various  instances  with  a  value  around  2.3.    An  initial  estimate  of  the  classification 
function  sets  K1  equal  to  the  CSS,  K2  to  the  largest  particle  in  the  product  (should  be 
close to OSS), and K3 equal to 2.3.  
 
The  breakage  function  is  determined  from  single  particle  breakage  testing  (section 
2.4).    It  is  a  size  distribution  dependent  upon  the  nature  of  the  material  broken  and  to 
some extent the operating conditions of a crusher.  Within a jaw crusher, a large number 
of  single  particle  breakage  events  are  occurring  simultaneously,  with  a  portion  of  the 
products  of  breakage  undergoing  subsequent  crushing.    The  breakage  distribution 
mathematically  describes  this  process  as  it  considers  breakage  in  terms  of  a  series  of 
single particle fracture events (Kelly and Spottiswood, 1982).  A more detailed discussion 
of the breakage function can be found in section 2.4.3. 
 
K2 K1
0
P
r
o
b
a
b
i
l
t
y
 
o
f
 
b
r
e
a
k
a
g
e
,
 
C
(
x
)
Particle size, x
1.0
Figure 2.10 Whiten classification function
K3
 
Csoke (et al., 1996) developed an empirical method for determining the product size 
of  preliminary  crushers,  i.e.  jaw  crushers  and  gyratory  crushers.    The  product  size 
resulting from breakage of the material larger than the CSS of the crusher can be modeled 
using the following function: 
  (   )
max
m
r
P d
r
|   |
=
    |
\   .
  [2.6] 
with, 
  17
 
max
p
pMAX
d
r
Gap
d
r
Gap
=
=
 
where, m is exponent describing the product size distribution 
d
p
 is the particle size 
    d
pMAX
 is the largest size in the product (OSS) 
the Gap is the CSS of the crusher. 
Based on operating data found in the literature Equation 2.6 fits the product size of a jaw 
crusher with r
max
 equal to 1.223 and m of 0.842.   
 
  King (2001) outlined a model for the prediction of product size based on data from a 
crusher  manufacturer.    The  main  concept  is  that  the  size  distribution  of  the  product  is 
characterized by the particle size relative to the OSS and by the product type, P
t
, which is 
the fraction of the product smaller than the OSS.  P
t
 is related to the nature of the material 
and is distinguished by the work index and a qualitative description of the material.  The 
size distribution given by this model is determined from the following equations: 
 
(   )
(   )
1.5
0.85
1                  for  0.5
1                  for  0.5
u
r
K
r
KL
P d e r
P d e r
|   |
|   |
   |
   |
   |
\   .
\   .
|   |
|   |
   | 
   |
   |
\   .
\   .
=      >
=      <
  [2.7] 
with, 
 
1.5
0.67
1.18
0.5
1
ln
1
1
0.5 ln
1
1     
u
i
i
u
t
l
b
K
b
D
r
OSS
K
P
K
P
P e
|   |
|   |
   |
   |
   |
\   .
\   .
=
   ( |   |
=
    (    |
\   .    
   ( |   |
=
     (    |
\   .    
=   
 
Equation 2.7 should only be used when the size at which half of the feed material passes 
is greater then the OSS.   
  
Modeling of power draw 
The  most  popular  method  of  determining  the  power  draw  or  consumption  of  size 
reduction  equipment  is  Bonds  method.    Bonds  theory  states  that  the  work  input  is 
proportional to the new crack tip length produced in particle breakage and is equal to the 
  18
work  represented  by  the  product  minus  the  work  represented  by  the  feed  (Bond,  1952).  
Based on this theory the equation for determining the work input (kWh/metric ton) is: 
 
80 80
1 1
10 W WI
P F
|   |
=          
\   .
|
|
  [2.8] 
where, WI is the work index in kWh/metric ton (see section 2.1.2) 
    P
80
 is the size at which 80% of the product passes 
    F
80
 is the size at which 80% of the feed passes. 
Using Equation 2.6 the power draw of a size reduction machine can be found for a given 
feed throughput from: 
  P W T =      [2.9] 
where T is the throughput in metric tons per hour. 
 
  Although  Bonds  equation  applies  reasonably  well  to  grinding  conditions  (i.e.,  ball 
mills  and  rod  mills)  its  relevance  to  primary  crushing  equipment  processing  large  feed 
sizes  is  questionable.    Bonds  law  is  an  empirical  law  that  only  fits  experimental  data 
over  a  limited  range  of  variables  and  only  in  certain  cases  (Choi,  1982).    It  can  be 
corrected  for  other  operating  conditions  but  even  with  this correction  it  has  been  shown 
that  the  power  consumption  of  a  jaw  crusher  can  be  as  much  as  240%  greater  than  the 
expected  consumption  based  on  Bonds  equation  (Eloranta,  1997).    Rose  and  English 
have  suggested  a  relationship  for  power  consumption  based  on  Bonds  work  index  but 
take into account the material density and machine characteristics, i.e., gape, throw, CSS, 
etc. (Lowrison, 1974).   
 
A  more  recent  approach  to  predicting  power  consumption  has  focused  on  the  power 
required  by  laboratory  breakage  devices  to  achieve  the  same  size  reduction  as  seen  in 
actual  crushers  (Andersen  and  Napier-Munn,  1988;  Morrell  et  al.,  1992).    The  power 
required by the lab device, in this case the twin pendulum or the drop-weight (see section 
2.4.2), is related to the actual power drawn by a crusher using the following expression: 
 
c p
P AP P
n
=   +   [2.10] 
where, P
c
 is the power drawn by the crusher under load in kW 
    P
p
 is the pendulum power in kW 
    P
n
 is the power drawn by the crusher under no load in kW 
    A is scaling factor specific to a crusher (from regression analysis). 
The  process  for  determining  the  pendulum  power  is  based  on  the  classification  and 
breakage  functions  described  previously  and  will  not  be  covered  here.    It  can  be 
expressed as the total energy required to reduce the crusher feed size to the product size 
as if all the reduction took place in the laboratory device (Napier-Munn et al., 1999).  
 
Estimating capacity 
The  capacity  of  a  jaw  crusher  is  dependent  upon  the  machine  characteristics,  the  feed, 
and  the  nature  of  the  rock  material.    The  volumetric  capacity  of  a  jaw  crusher  can  be 
estimated from the following equations (Sastri, 1994): 
  19
 
(   )
  (   )
(   )
2
60 0.5        at low speeds
450
60 0.5                      at high speeds
DT
V Nw CSS T
G CSS T
g
V Nw CSS T
N
|   |
=   +
     |
   |
   +
\   .
|   |
=   +
     |
\   .
  [2.11] 
where, N is the speed of the crusher in rpm 
    w is the width of the jaws in m 
    CSS is the closed side set in m 
    T is the throw in m 
    D is the vertical depth between the jaws in m 
    G is the crusher gape in m 
g is gravitational acceleration in m/s
2
 . 
Equation  2.11  gives  volumetric  capacity  under  ideal  conditions.    At  low  crusher  speeds 
the  material  falls  down  the  chamber  due  to  gravity  flow,  and  the  distance  of  the  fall  is 
dependent on the geometry of the machine.  At higher speeds there is less time between 
strokes/revolutions  and  movement  of  material  down  the  chamber  is  constricted  and 
controlled by the speed of the machine (Hersam, 1923).  
 
  In order to determine the actual capacity of a jaw crusher, Equation 2.11 needs to be 
corrected  by  accounting  for  the  feed  size,  the  degree  of  compaction  resulting  from 
vibration,  and  the  nature  of  the  material.    Equation  2.12  gives  the  estimated  capacity 
using these parameters (Sastri, 1994): 
 
(   )
  (   )
(   )
1 2 3
1 2 3 2
60 0.5       at low speeds
450
60 0.5                     at high speeds
DT
V Nw CSS T K K K
G CSS T
g
V Nw CSS T K K K
N
|   |
=   +
     |
   |
   +
\   .
|   |
=   +
     |
\   .
  [2.12] 
with, 
 
2.5
1
6.5
2
0.85
1.92 10
avg
T
G
F
K
G
K
  
|   |
=   
   |
\   .
=   
  [2.13] 
where, F
avg
 is the average feed size in m 
    K
3
 is a parameter related to the nature of the material. 
The  parameter  K
3
  does  not  have  a  suggested  value  but  is  considered  to  increase  with 
increasing hardness or toughness.   
 
2.1.4 CONCLUSIONS 
 
  Jaw  crushers  are  large,  heavy-duty  machines  capable  of  crushing  large  quantities  of 
tough,  abrasive  materials.    They  are  typically  employed  as  primary  crushers  within 
aggregate  processing  plants.    Jaw  crushers  are  most  commonly  defined  by  gape  size, 
  20
open and closed side sets, and capacity.  There are a wide range of jaw crusher sizes for 
varying production and product size requirements. 
 
  The  plates  of  a  jaw  crusher  are  used  to  apply  compressive  forces  that  induce  tensile 
stresses within particles, causing fracture.  Particles are repeatedly nipped until they pass 
through  the  crushing  chamber.    The  breakage  process  occurring  between  the  jaw  plates 
acts  simultaneously  with  a  classification  process.    The  classification  process  defines 
whether or not a particle will be subjected to crushing and is dependent upon the settings 
of the crusher and particle size. 
 
The  selection  of  a  jaw  crusher  for  application  in  the  aggregate  industry  is  primarily 
based  on  technical  literature  provided  by  crusher  manufacturers,  experience,  and  cost.  
Charts  and  graphs  provide  data  on  electric  power  requirements,  crusher  size  (gape),  as 
well  as  expected  capacities  for  a  given  material  and  closed  side  set.    These  charts  and 
graphs  have  also  been  incorporated  into  computer  programs  to  aid  in  crusher  selection.  
In  order  to  account  for  material  variations  manufacturers  rely  on  a  suite  of  laboratory 
tests.    The  uniaxial  compression  and  Bond  impact  crushability  test  are  two  of  the  most 
common tests used to rank materials relative to their hardness/crushability.  Each of these 
tests  has  limitations  and  neither  adequately  describes  a  materials  resistance  to  fracture.  
Although  jaw  crushers  are  used  extensively,  the  lack  of  understanding  relative  to  their 
operational characteristics, as well as a reliance on an inappropriate, single set of material 
properties,  makes  selection  of  a  proper  machine  difficult.    Jaw  crusher  selection  is  also 
heavily influenced by the subjective judgment/experience of individuals, which can result 
in the conservative selection and operation of jaw crushers.  
 
  The  prediction  of  crusher  performance  is  typically  concerned  with  the  size 
distribution  of  the  product  exiting  the  crusher,  the  machines  power  draw,  and  the 
capacity.  The capacity of a jaw crusher is dependent upon the operating characteristics of 
the  machine  and  influenced  by  the  feed  size  and  the  nature  of  the  material.    It  can  be 
determined fairly easily but there has been difficulty in accounting for the variable nature 
of rock materials.  Bonds theory or equation has long been used to determine the power 
draw and motor requirements of size reduction machines.  However, its relevance to jaw 
crushers  handling  large  feed  sizes  has  been  questioned.    The  recent  trend  has  been  to 
calculate  power  draw  based  on  laboratory  crushing  tests.    The  modeling  of  crusher 
performance  has  focused  mainly  on  the  prediction  of  the  product  size  distribution.    The 
product  size  distribution  of  a  crusher  can  be  determined  using  the  Whiten  model  given 
the  feed  size,  the  classification  function,  and  the  breakage  function.    The  classification 
function  is  dependent  upon  the  operating  parameters  of  the  crusher,  namely  the  closed 
side and open side set.  The breakage function considers breakage in terms of a series of 
single  particle  fracture  events  and  is  dependent  upon  the  nature  of  the  material.    It  is 
determined from a series of laboratory single particle breakage tests. 
 
  There is significant room for improvement when it comes to selecting and optimizing 
jaw  crushers.    The  enhancement  of  crusher  selection  and  the  prediction/optimization  of 
  21
its  performance  are  dependent  upon  the  operating  parameters  of  the  machine  and  the 
physical  properties  of  the  material  being  broken.    Early  attempts  at  establishing 
relationships  between  these  components  focused  on  the  energy  consumed  by  a  crushing 
machine and the resultant size reduction.  Energy-size relationships themselves require an 
awareness  of  the  mechanisms  of  fracture,  the  fundamentals  of  which  are  covered  in  the 
following section.   
 
 
2.2 PHYSICS OF PARTICLE FRACTURE 
 
The breakage of particles in any comminution or crushing machine is a difficult process 
to understand.  Comprehending the sub-processes that occur prior to material failure can 
provide a basis for evaluating the limitations of crushing machines or set the direction for 
potential  improvements  (Schoenert,  1972).    Knowledge  of  the  mechanisms  of  fracture 
provides  information  about  the  final  stages  of  strain  within  a  particle,  namely  the 
formation and propagation of cracks (Tkacova, 1989).  This provides a basis for how to 
properly characterize materials, i.e., which mechanical properties are representative of the 
fracture process.   
 
The  mechanisms  of  fracture  also  affect  the  distribution  of  resultant  particle  sizes.  
Different  methods  of  breakage  produce  different  product  sizes  based  on  the  amount  of 
energy  input  and  process  of  loading.    A  perception  of  the  expected  distribution  of 
resultant particle sizes can be ascertained simply based on fracture physics. 
 
2.2.1 PARTICLE BEHAVIOR  
 
Rocks  are  brittle  materials  and  the  amount  of  strain  within  a  particle  under  load  is 
proportional  to  the  applied  stress.    Recalling  that  jaw  crushers  induce  tensile  stresses 
within particles, it is the tensile strength of rock materials that must be exceeded in order 
for the material to fracture.  Theoretically the tensile strength of an ideal brittle material 
should be approximately: 
 
10
t
E
   =   [2.14] 
 
where, E is Youngs modulus. 
In  reality  the  actual  tensile  strength  of  a  rock  is  well  below  the  estimate  given  by 
Equation  2.12,  usually  at  least  100  times  lower.    This  decrease  in  strength  is  due  to  the 
presence of pre-existing flaws or cracks within the rock.  
 
  Pre-existing  flaws  and  cracks  within  rocks  act  as  stress  concentrators.    Inglis  (1913) 
provided  a  solution  for  the  case  of  an  elliptical  hole  in  a  stressed  plate  (Fig  2.11).    He 
found that the stress concentration near the tip of the ellipse was proportional to the size 
of the ellipse and the ellipse radius.  Inglis stress concentration factor can be written as: 
  22
 
 
2b 
2a
Figure 2.11 Elliptical hole in an infinite plate
  1 2
a
k
|   |
=   +
\   .
|
|
  [2.15] 
where, a is the half-length of the ellipse 
     is the tip radius of the ellipse. 
Griffith (1921) extended Inglis work by hypothesizing that pre-existing cracks in brittle 
materials  would  act  like  Inglis  ellipse  causing  stresses  to  concentrate  at  the  crack  tips.  
Griffith used Inglis stress analysis to establish a relationship between fracture stress and 
crack  size  using  an  energy  balance  approach.    Griffith  established  that  the  fracture  of 
brittle materials is due to the existence of pre-existing flaws and is an energy-controlled 
process. 
 
2.2.2 GRIFFITHS THEORY 
 
Griffiths proposed that the failure of a brittle solid is caused by the extension of inherent 
cracks resulting in a new crack surface that absorbs energy supplied by the work done by 
an  external  force  or  by  the  release  of  stored  strain  energy  from  within  the  solid  (failure 
occurs  when  the  energy  supplied  by  an  external  force  or  by  the  release  of  stored  strain 
energy  is  greater  than  the  energy  of  the  new  crack  surface).    This  implies  that  the  more 
energy a material absorbs, the more resistant it is to crack extension.  
 
  Two important criterions must be met in order for brittle failure to occur according to 
Grifffiths theory (I. Chem. E. Report, 1974). 
1.  There must be a mechanism for which crack propagation can occur.  At some 
point  in  the  material,  the  local  stress  must  be  high  enough  to  overcome  the 
  23
molecular  cohesive  strength  of  the  material.    This  is  accomplished  through 
stress concentration around Griffith cracks or inherent flaws. 
2.  The process must be energetically feasible.  Enough potential energy must be 
released  in  order  to  overcome  the  materials  resistance  to  crack  propagation.  
This  can  be  accomplished  by  increasing  the  work  done  by  external  forces 
acting on the material.  
 
  As  noted  previously  Griffith  used  Inglis  stress  analysis  in  his  energy  balance 
approach.  Thus the total energy for infinite plate (Fig 2.11) with an elliptical crack can 
be written as: 
 
t c
U U U W U
s
=   +      +   [2.16] 
where, U
t
 is the initial elastic strain energy of an uncracked plate 
    U
c
 is the elastic energy release caused by the introduction of a crack 
    W is the work done by external forces 
    U
s
 is the change in the elastic surface energy due to the new crack surfaces. 
Using elastic theory expressions, the four energy components can be determined, yielding 
the following equation (Whittaker et al., 1992): 
 
2 2 2
4
2 ' ' 2
s
A a A
U
E E
      
a =         +   [2.17] 
where, A is the infinite area of the plate 
    E is the effective Youngs modulus (E for plane stress; E/(1-v
2
) for plane strain) 
     is the strain 
    
s
 is the specific surface energy. 
Griffith  defined  the  fourth  energy  component,  where  
s
  is  a  constant  material  property 
representing the energy required to create a unit area of new crack surface. 
 
  According to Griffith the crack will propagate when an increase in its length does not 
change  the  net  energy  of  the  plate,  when  dU/da  equals  zero.    Differentiating  Equation 
2.14 with respect to a results, and setting the result equal to zero, gives: 
 
2
2
4
'
s
a
E
 0    =   [2.18] 
Equation  2.18  can  be  arranged  to  determine  the  fracture  stress  required  to  cause  crack 
initiation and more importantly the strain energy release rate, G: 
 
2
'
a
G
E
=   [2.19] 
G characterizes the energy per unit area required to extend the crack and is expected to be 
a fundamental physical property controlling the crack (Dowling, 1999).  
 
  Modifications of Griffiths theory have since followed, mainly in order to compensate 
for plastic deformation near the crack tip.  In fact, due mainly to plastic deformation, the 
energy  required  for  crack  extension  has  been  found  to  be  10  times  higher  than  that 
predicted  by  Griffith.    Plastic  deformation  acts  to  relax  strain  energy  since  energy  is 
  24
consumed  in  deforming  the  material  ahead  of  the  crack  tip.    The  high  stress  that  would 
ideally exist near the crack tip is effectively spread over a larger region or redistributed, 
resulting in a lower stress near the crack tip that may be resisted by the material.  Orowan 
(1949)  and  Irwin  (1948)  took  this  into  account  and  included  the  work  of  plastic 
deformations in calculations of the energy balance of fracture.  
 
  Despite  the  definition  of  a  characteristic  material  property  that  describes  crack 
propagation,  the  importance  of  Griffiths  work,  in  particular  reference  to  particle 
breakage, really lies in the connection made between the stress required for fracture and 
the presence and size of cracks or flaws.  Any criterion used to define or distinguish the 
crushability  of  rock  material  needs  to  account  for  the  presence  of  flaws  and  their 
influence on the work required to initiate fracture. 
 
Influence of particle size 
Griffiths theory indicates that fracture is dependent upon the presence of inherent flaws.  
Rocks, due to their geologic nature and inhomogeneity, contain a large number of flaws 
and  cracks,  both  on  the  macro  and  microscopic  level.    Thus  in  addition  to  the  mere 
presence of flaws controlling fracture, the distribution of flaws also affects the fracture of 
particles.  In order to describe, and even compensate for, the distribution of flaws within 
brittle materials statistical approaches are generally employed.   
   
  The  most  well  known  statistical  description  of  flaw  distribution  and  its  effect  on 
particle fracture is Weibulls (1939) model.  Weibulls weakest link theory assumes that 
the fracture of a particle is dependent only on the local strength of its weakest flaw from 
which  the  most  severe  crack  will  propagate  and  is  independent  of  all  other  flaws.  
Weibulls model shows that the presence of this weakest flaw becomes less probable with 
decreasing  particle  size.    This  is  in  agreement  with  experimental  results  that  indicate 
strength increases with decreasing particle size.  As particle size decreases critical flaws 
are essentially used up in order of their significance.   
 
Particle size effect can also be explained using energy considerations.  Recalling that 
failure  of  a  brittle  solid  occurs  when  the  energy  supplied  by  an  external  force  or  by  the 
release of stored strain energy is greater than the energy of the new crack surface, smaller 
particles,  which  have  less  capacity  for  storing  elastic  energy  (U
t
  is  proportional  to 
volume),  will  require  more  work  by  external  forces,  or  higher  stresses,  in  order  to 
fracture.    Smaller  particles  also  exhibit  a  more  plastic  response  than  larger  ones.    The 
high  stress  levels  in  small  particles  result  in  irreversible  deformations  that  modify  the 
stress  distribution  within  the  particle  and  result  in  coarser  fragments  upon  fracture 
(Schoenert, 1972).   
 
  The effects of size on particle fracture call in to question the applicability of strength 
parameters to crushing resistance.  A jaw crusher sees a wide  range of particle sizes.  It 
also  must  re-crush  smaller  particles  produced  from  the  crushing  of  larger  ones.    The 
smaller  particles  will  require  higher  stresses  as  the  flaw  distribution  changes  with  a 
  25
decrease  in  particle  size.    Since  tensile  strength  mostly  controls  fracture  initiation  of 
particles  in  a  jaw  crusher,  characterization  of  a  rock  material  using  a  laboratory  based 
tensile  strength  test  would  require  either  a  large  number  of  samples  of  varying  size  or 
statistical rock mechanics in order to analyze size effects.  A more appropriate method of 
characterizing  rock  materials  would  be  one  that  compensates  for  size  effects.    For 
example rearranging Equation 2.18 into: 
  2 '
s
a E       =      [2.20] 
indicates that, according to Griffiths criterion, fracture will occur when  a     reaches a 
constant  critical  value  determined  by  the  characteristic  material  properties  E,  ,  and  
s
.  
Higher  stresses  are  normalized  by  the  smaller  flaw  size,  a,  around  which  they  are 
concentrating.  When considering fracture physics as it pertains to crushers, resistance to 
fracture should account for the stresses required for breakage, the presence of flaws, and 
be independent of particle size. 
 
2.2.3 FRACTURED SIZE DISTRIBUTION 
 
The fracture process due to the point contact loading that occurs between the plates of a 
jaw crusher and a particle is illustrated in Figure  2.12.    The  induced  tensile  stress  along 
the  axis  of  the  particle  results  in  fracture  by  cleavage.    The  areas  directly  below  the 
loading  contacts  fail  in  compression  producing  abrasion  fracture.    Abrasion  can  be 
thought  of  as  type  of  shatter  fracture  (shatter  is  the  third  type  of  fracture    common  in 
impact breakers).  Shatter occurs when there is excess energy input and a large number of 
flaws  are  stressed  almost  simultaneously  or  before  the  weakest  flaw  is  unloaded.  
Abrasion is localized shatter.  There is not enough energy to shatter the entire particle but 
the impact of the jaw plates on the particle does produce attrition at the load point. 
Figure 2.12 Fracture caused by compression crushing 
(After Kelly and Spottiswood, 1982)
 
  Coarse progeny particles are a result of cleavage fracture and finer sized progeny are 
a  result  of  abrasion/shatter.    A  comparative  display  of  the  distribution  of  progeny  sizes 
  26
after  particle  failure  is  represented  in  Figure  2.13.    Since  two  mechanisms  of  fracture 
occur  during  point  contact  loading,  resultant  size  distributions  from  jaw  crushers  are 
expected to be bimodal.  Lynch (1977) attributed the two components as being a result of 
localized fracture against the crusher face (fine size distribution) and main fracture of the 
particle (coarse distribution).   
Figure 2.13 Size distributions occurring due to mechanisms of fracture 
(From Bowers et al., 1991)
 
The  distribution  of  particle  sizes  after  fracture  is  dependent  on  the  fracture 
mechanisms occurring as a result of particle loading.  The mechanisms occurring during  
fracture are controlled by the intensity of energy applied to the particle.  The physics of 
fracture  and  quantification  of  the  energy  required  for  fracture  have  been  explained  by 
Griffiths  theory.    Since  the  same  mechanisms  control  product  size,  there  is  expected  to 
be  a  relationship  between  the  distribution  of  progeny  sizes  and  a  particles  resistance  to 
fracture (characterized by some form of Griffiths criteria).   
 
  27
2.2.4 CONCLUSIONS 
 
A  thorough  understanding  of  fracture  physics  and  the  mechanisms  that  drive  fracture 
initiation is necessary in the evaluation and improvement of jaw crushing operations.  It 
is  also  important  to  be  aware  of  fracture  mechanisms  when  using  material  properties  or 
indexes to describe the particle breakage process or determine crusher performance. 
 
  Brittle  materials  fail  at  stress  levels  well  below  what  is  predicted  based  on  stress-
strain  behavior  (Youngs  modulus).    The  presence  of  cracks  accounts  for  the  markedly 
lower strength.  Rocks are brittle materials and by their nature also contain inherent flaws 
(i.e., grain boundaries, voids/pores, etc.) on both small and large scales.  These cracks act 
as stress concentrators, with the increase in stress at these locations being proportional to 
the length of the crack and the crack tip radius.  
 
  Griffith  used  Inglis  stress  analysis  of  an  elliptical  crack  in  an  infinite  plate  to 
determine  a  relationship  between  crack  size  and  fracture  stress.    Griffith  hypothesized 
that  fracture  occurs  when  the  energy  supplied  by  an  external  force,  or  by  the  release  of 
stored strain energy, is greater than the energy of the new crack surface.  Using an energy 
balance  approach  Griffith  was  able  to  define  the  strain  energy  release  rate  G,  a  material 
property  that  characterizes  the  energy  per  unit  area  required  to  a  crack.    Griffiths  work 
also defined a constant critical value of  a      that when met will result in fracture.  The 
importance in terms of crushing operations is that fracture is dependent on applied loads 
and  the  presence  and  size  of  flaws,  and  both  need  to  be  accounted  for  when 
characterizing a rock materials resistance to fracture.  
 
  It  has  been  shown  that,  for  a  given  material,  as  particle  size  decreases  strength 
increases.  This is due to the distribution of flaws within the material.  Weibulls weakest 
link  theory  states  that  the  strength  of  a  particle  is  dependent  upon  its  most  critical  flaw.  
Fracture initiates from that flaw independently of all other flaws within the particle.  As 
the size of the particle becomes smaller the existence of such a critical flaw becomes less 
probable,  thus  the  increase  in  strength  with  decreasing  particle  size.    Defining  a 
materials resistance to fracture should account for the effect of size or be independent of 
it. 
 
  Since  the  mechanisms  of  fracture  also  control  the  distribution  of  progeny  particle 
sizes  and  specific  fracture  mechanisms  produce  specific  fragment  sizes,  particle  physics 
is the foundation for the first models/equations used to describe size reduction processes.  
Energy-size  reduction  relationships  are  related  to  Griffiths  energy  criterion  and  the 
presence of Griffith cracks. The energy criterion states that enough potential energy must 
be  released  in  order  to  overcome  a  materials  resistance  to  crack  propagation,  requiring 
an increase in the work done by external forces acting on the material.  This is the amount 
of energy input into reducing the size of a particle.  The amount of size reduction, or the 
size distribution resulting from fracture, is dependent upon the presence and distribution 
  28
of  Griffith  cracks.    Energy-size  reduction  relationships  are  a  natural  progression  of  the 
physics of particle fracture and form the principles of comminution. 
 
 
2.3 FRACTURE ENERGY AND SIZE REDUCTION 
 
Relationships  between  energy  and  size  reduction  are  concerned  with  the  amount  of 
energy input by a comminution machine and the degree of size reduction attained.  It was 
pointed  out  in  section  2.2.2  that  different  sized  particles  require  different  amounts  of 
energy in order to be fractured.   The effect of particle size on energy input was used to 
develop the first relationships between energy and size reduction.   
 
It  is  now  generally  accepted  that  established  relationships  between  energy  and  size 
reduction can be expressed using the following singular equation (Charles, 1957): 
 
d
d
n
d
E K
d
=    [2.21] 
where, E is the specific energy necessary to supply the new surface energy 
    K is a constant 
    d is the particle size 
n is a value describing different size ranges. 
Equation  2.21  is  the  general  form  energy-size  relationships  proposed  by  various 
researchers.  The principles behind their suggestions and their specific equations follow. 
 
2.3.1 LAWS OF COMMINUTION 
 
The work of Von Rittinger, Kick, and Bond constitute what are commonly referred to as 
the  laws  of  comminution.    Von  Rittinger  (1867)  stated  that  the  breakage  energy  is 
proportional  to  the  area  of  the  new  surfaces  produced  and  that  the  energy  requirement 
remains constant for a unit of surface energy produced.  Von Rittingers law is: 
  (   )
2 1
E K A A =      [2.22] 
where, A
1
 is the specific surface area of the initial particle 
    A
2
 is the specific surface area of the final particle. 
Since  the  specific  area  is  inversely  proportional  to  the  diameter,  Equation  2.22  can  be 
rewritten as: 
 
2 1
1 1
E K
d d
|   |
=   
\   .
|
  [2.23] 
where, d
1
 is the initial particle diameter (mean size for a distribution of particles) 
    d
2
 is the final particle diameter (mean size for a distribution of particles). 
Von  Rittingers  theory  is  similar  to  Griffiths  since  it  relates  the  energy  input  into 
creating  new  surfaces  (cracks)  to  the  surface  area  of  the  cracks.    However,  it  only 
accounts  for  the  energy  required  to  pull  apart  molecular  bonds,  and  as  Griffith  showed 
there are many other energy aspects to overcome. 
  29
 
  Kick  (1883)  postulated  that  the  energy  required  for  breakage  is  proportional  to  the 
size  or  volume  of  the  particle  and  that  the  energy  requirement  remains  constant  for 
equivalent geometrical changes.  Kicks equation is: 
 
1
2
ln
d
E K
d
=   [2.24] 
The problem with Kicks law is that it assumes the energy required to achieve a certain 
degree of size reduction will remain constant for equivalent changes in particle volume.  
The  effect  of  size  on  the  amount  of  energy  required for fracture, as discussed in section 
2.2.2,  is  not  considered  by  Kick.    But  smaller  particles  will  require  more  energy  to 
fracture and the energy requirement will increase as particle volume decreases. 
 
Bond stated that the energy input is proportional to the new crack tip length produced 
and is equal to the work represented by the final particle size minus the work represented 
by  the  initial  particle  size.    His  law  of  comminution  is  in  fact  an  empirically  derived 
relationship  based  on  a  series  of  grinding  tests.    A  specific  form  of  Bonds  equation  is 
given  on  page  19  in  Equation  2.8.    A  general  form  in  relation  to  Von  Rittingers  and 
Kicks is: 
 
2 1
1 1
E K
d d
|   |
=    
\   .
|
|
a. 
b. 
c. 
  [2.25] 
In  some  respects  Bonds  law  resolves  Kick  and  Von  Rittingers  theories.    Bond  argued 
that  crushing  is  concerned  with  both  surface  area  (Von  Rittinger)  and  particle  volume 
(Kick).  The volume of the particle is proportional to the amount of stress absorbed and 
stresses concentrate on the surface, which eventually propagate into cracks.  Bonds work 
attempted  to  follow  Griffiths  theory  since  the  premise  is  that  when  local  deformation 
exceeds the critical strain a crack tip forms, the surrounding stress energy flows to it, and 
fracture  follows  (Pryor,  1974).    However,  Austin  and  Brame  (1983)  found  that  Bonds 
work has no factually based relation to Griffiths theory.  Bond did introduce the effect of 
material  properties  though,  noting  that  different  materials  will  have  different  resistances 
to  a  reduction  in  their  size  (Equation  2.6  included  the  work  index,  WI,  of  the  material 
being crushed).      
 
  The three laws of comminution have been shown to be applicable over different size 
ranges (Figure 2.14).  For particle sizes encountered during crushing Kicks law is most 
applicable.  Von Rittingers law fits fine size reduction fairly well and Bonds empirical 
law  of  course  corresponds  to  the  grinding  range.    Despite  the  inability  of  each  law  to 
fully  characterize  an  energy-size  reduction  relationship  for  the  entire  spectrum  of  size 
ranges, they do recognize some important aspects (Manca et al., 1983): 
The relative importance of the energy expended 
The importance of reduction ratio 
The importance of the feed and product size distribution 
  30
d. 
e. 
Figure 2.14 Size ranges applicable to Von Rittenger, Kick, and Bond equations
(After Hukki, 1961)
The  effects  of  different  conditions  of  stress  application  for  different  size 
reduction devices 
The influence of the characteristics of the material. 
 
Advancement  of  the  three  laws  has  included  relating  energy  requirements  to  non-
uniform  particle  size  distributions  as  well  accounting  for  changes  in  particle  strength.  
Holmes  (1957)  modified  Kick  law  in  order  to  account  for  elastic  deformation  and  the 
effect of reduced particle size on strength.  For a given reduction ratio and product size, 
Holmes  considers  the  exponent  n  in  Equation  2.21  a  measure  that  expresses  the  degree 
of variation in particle strength with variations in size, and is specific to the material and  
the manner of stress application.  Von Rittinger, Kick, and Bond assumed that the shape 
of  the  size  distribution  remains  fairly  constant,  which  is  not  the  case,  particularly  for 
crushers (Napier-Munn et al., 1999).  Charles (1957) included the size modulus from the 
Schuhmann distribution function in his equation for energy.  His equation relating energy 
requirement to product size distribution reduction is: 
  31
 
1
1
1
1 1
n
n
K
E
n n d
|
=   
      +
\   .
|
|
  [2.26] 
where,  is the Schumann size modulus 
     is a constant in Schumanns distribution function 
    d is the particle size. 
Equation 2.26 describes the relationship between energy and size reduction and includes 
all the parameters upon which energy input is dependent on, but it only applies to simple 
comminution  systems  and  falls  short  of  being  applicable  over  a  wide  range  of  sizes.  
Although  Charles  equation  itself  is  not  universally  applicable,  it  also  offers  the 
theoretical proof that a general law valid for all comminution processes cannot exist and 
that  every  rock  fragmentation  process  requires  the  derivation  of  a  specific  analytical 
expression (Manca et al., 1983). 
 
2.3.2 CONCLUSIONS 
 
A number of researchers have investigated the relationship between energy input and size 
reduction in an attempt to establish an equation or model capable of predicting the energy 
requirements of comminution machines (for a given degree of size reduction).  The laws 
of Von Rittinger, Kick, and Bond do not describe the energy-size reduction relationship 
for all size ranges but each has been found to pertain to some particle size class.   
 
The  three  laws  are  related  to  Griffiths  theory  and  seem  to  be  founded  in  fracture 
physics.  However, Von Rittingers law only considers surface area, thus its application is 
to fine sizes where volume effects are minimal.  Kicks law considers particle volume but 
disregards  the  effect  of  flaws  or  cracks  on  particle  strength,  the  distribution  of  which 
tends  to  decrease  with  size,  making  smaller  particles  more  difficult  to  fracture.    Kicks 
law is applicable to coarse particle sizes.  Bond considered both surface area and volume 
in his theory and actually argued that it was related to Griffiths theory.  That argument 
has been shown to have no basis, but since Bonds equation was empirically derived from 
grinding tests his model fits particle sizes typically found in that application.   
 
Bond  introduced  the  concept  of  work  index  to  compensate  for  the  reality  that 
different  materials  exhibit  different  resistances  to  a  reduction  in  their  size.    Additional 
work  demonstrates  that  energy-size  relationships  need  to  account  for  non-uniform  size 
distributions  as  well  as  variability  in  particle  size,  strength,  and  load  application.    Yet 
inclusion  of  these  factors  still  does  not  provide  a  singular  relationship  capable  of 
predicting energy requirements for any range of size reduction.  In reality a common law 
valid  for  all  size  reduction  processes  probably  does  not  exist  and  empirically  derived 
models need to be fully developed for particular breakage processes.   
 
The  complexity  of  the  energy-size  reduction  relationship  has  lead  investigators  to 
seek an alternative method of assessing the energy requirements and product distribution 
of size reduction devices.  In the past 40 years a method has been developed in order to 
  32
investigate  the  science  of  size  reduction.    The  method  is  based  on  the  fracture  physics 
discussed in section 2.2 and focuses on: fracture mechanisms, crushing under controlled 
conditions,  packed  bed  crushing,  measurement  of  energy  needed  for  fracture,  fragment 
size  distribution,  new  surface  area  produced,  mathematical  simulation  of  crushing 
processes,  and  practical  applications  (Rumpf,  1966).    The  approach  used  to  investigate 
these issues is referred to as single particle breakage analysis. 
 
 
2.4 SINGLE PARTICLE BREAKAGE ANALYSIS 
 
The  breakage  process  occurring  in  a  crusher,  or  any  comminution  device,  is  a  series  of 
single particle crushing events.  A particle is loaded and fractures only under the stresses 
applied  to  it.    The  product  coming  out  of  the  crushing  device  is  just  a  collection  of 
progeny  particles  from  numerous  single  particle  events.    An  understanding  of  single 
particle  fracture,  which  is  the  fundamental  process  of  comminution,  is  of  great  benefit.  
Additionally  it  provides  a  means  to  separate  a  materials  fracture  behavior  from  the 
operating conditions of comminution machines.  The three laws discussed previously are 
limited  to  certain  methods  of  comminution  and  include  characteristics  of  those  methods 
(Krogh,  1980).    Characterization  of  a  materials  fracture  behavior  (i.e,  by  determining 
values like fracture energy, fragment size) should be independent of the eventual method 
of comminution used.  
 
  Although  the  concept  of  applying  fracture  physics,  through  single  particle  breakage, 
to  comminution  processes  was  pioneered  by  Rumpf  beginning  in  the  1960s,  the  first 
single  particle  tests  where  conducted  in  the  1930s  (Carey  and  Bosanquet,  1933).    The 
purpose of these tests was to determine the energy necessary to reduce coal and gypsum 
particles (ranging from 2 to 50 mm) to a certain size.  But most of the work at this time 
focused  on  the  validation  or  refutation  of  Kicks  and  Von  Rittengers  theories.    Using 
compression  devices  (to  calculate  input  energy)  and  gas  adsorption  techniques  (to 
determine  surface  area),  the  relationships  suggested  by  Kick  and  Von  Rittenger  were 
tested.    But  since  Rumpfs  proposal  that  the  fundamental  areas  of  comminution  to  be 
studied  should  include  fracture  physics  and  the  material  characteristics  related  to 
breakage,  a  great  deal  of  progress  has  been  made  in  employing  results  from  single 
particle  breakage  analysis  to  improvements  in  actual  crushing  and  comminution 
applications.  The following outlines the concepts behind single particle breakage, the test 
methods  employed,  and  the  related  data  used  to  characterize  a  materials  response  to 
fracture.     
 
2.4.1 CONCEPTS OF SINGLE PARTICLE BREAKAGE 
 
Single particle breakage (SPB) is the study of fracture physics (section 2.2).  The state of 
stress arising in a particle loaded by a comminution machine is dependent on the size and 
distribution  of  cracks  or  flaws  within  the  material  and  the  deformation  behavior  of  the 
material.    The  flaw  size  and  distribution  are  dependent  on  particle  size  and  its 
  33
homogeneity.  And the type of breakage device affects the type, direction, and number of 
contact  forces  acting  on  the  particle,  and  the  velocity  at  which  they  are  applied.    Figure 
2.15 shows the number of events occurring during the breakage of a particle and their 
interdependence. 
Figure 2.15 Phenomena that effect single particle fracture 
(After Schoenert, 1987) 
  The presence and effect of flaws on particle fracture has been discussed in detail, as 
has  the  effect  of  particle  size  (section  2.2.2).    It  is  important  to  note  again  though  that 
rock  materials  are  inherently  flawed  and  only  at  very  small  particle  sizes  will  the 
presence of flaws be less critical.  At those fine sizes the behavior of the particle becomes 
inelastic.    In  terms  of  single  particle  breakage,  since  the  determination  of  progeny  size 
distribution is of great importance, the importance of flaws is two-fold.  In addition to its 
influence on fracture, the crack pattern arising out of flaw propagation will determine the 
size  and  the  shape  of  the  progeny  fragments  and  the  new  surfaces  created  (Schoenert, 
1979). 
 
  One  aspect  of  the  behavior  of  a  stressed  particle  that  has  not  previously  been 
discussed is the effect of deformation rate.  Although single particle breakage isolates the 
material  behavior  from  comminution  machine  operating  parameters,  there  is  some 
differentiation between slow compression and impact compression.  The deformation rate 
in a crusher is between 0.1 and 10 m/s, as compared to an impact mill that might deliver a 
rate between 20 and 200 m/s.  Single particle breakage analysis applicable to jaw crusher 
processes is generally considered slow compression (but only in a relative sense, a more 
appropriate  classification  might  be  slow  impact).    In  slow  compression  elastic  waves 
  34
arising  from  load  contact  do  not  affect  the  breakage  of  the  particle,  nor  does  the 
deformation rate, as long as the particle behaves mainly elastically (Schoenert, 1991). 
 
  The stress field generated in single particle crushing is dependent upon the direction 
and  number  of  forces  applied.    The  stress  field  arising  in  a  particle  under  load  has  been 
studied  mainly  for  the  case  of  an  elastic  sphere  since  the  stress-strain  behavior  of 
irregularly shaped particles cannot be calculated.  When an elastic particle is loaded by a 
contact  force,  the  area  underneath  the  contact  develops  a  cone-shaped  zone  of  stress 
concentration  in  which  particle  degradation  is  severe  compared  to  the  space  outside  the 
zone  (Prasher,  1987).    Fine  fragments  originate  mainly  from  these  cone-shaped  zones, 
with  coarse  fragments  coming  from  the  zones  away  from  the  contact  area.    This 
corresponds to the situation of a particle under localized compression (Figure 2.3) and the 
resultant fracture pattern (Figure 2.12).  When multiple forces act on a particle the result 
is  more  fracture  since  cracks  grow  under  the  contact  areas  and  propagate  towards  other 
contacts.  Since jaw crushers perform arrested crushing particle stressing is through two 
contacts only, between the jaw plate and particle (see section 2.1.1).   
 
Figure  2.12  does  not  take  into  account  what  occurs  after  initial  breakage.    Since  the 
throw  (crushing  stroke)  of  a  jaw  crusher  is  fixed,  particles  are  often  loaded  beyond  the 
point  of  first  fracture.    It  is  important  in  single  particle  breakage  to  distinguish  between 
the first stage of primary fracture and the second stage in which a group of fragments are 
compressed;  with  fragmentation  continuing  until  the  reaction  stress  of  the  fragments  is 
equal to the applied stress (Schoenert, 1987).  The amount of fragment particles crushed 
in  the  second  stage  is  dependent  on  the  material  hardness  (harder  materials  will  have 
fragments that leave the crushing chamber at higher velocities) and the loading velocity.  
 
2.4.2 TEST METHODS 
 
There  are  basically  three  ways  to  reduce  the  size  of  particle,  by  slow  compression, 
impact,  and  abrasion/shear.    Since  size  reduction  processes  are  carried  out  in  various 
ways  there  are  different  methods  of  conducting  single  particle  breakage  (SPB)  analysis, 
including:  impact  by  falling  media,  dropping  the  particle  on  a  surface,  and  slow 
compression  (Krough,  1980).    Single  particle  breakage  analyses  utilizing  slow 
compression  and  impact  crushing  are  most  applicable  to  the  conditions  of  jaw  crushing 
and are the focus here.  The distinction between slow compression and impact is the rate 
of load application.  Slow compression tests occur over a time scale of approximately 1 to 
10 seconds, while impact tests can take as little as 110
-5
 seconds to complete (Peters Rit 
et al., 1983).   
 
Slow compression is simple and has the advantage of being easy to execute (Mehrim 
and Khalaf, 1980).  More recently, single particle tests used to determine the behavior of 
rock  have  evolved  to  include  pendulum  and  drop  weight  tests  (Bearman  et  al.,  1997).  
The  Comminution  Center  at  the  University  of  Utah  has  made  variations  to  these  tests, 
developing the Ultrafast Load Cell Apparatus.  The Julius Kruttschnitt Mineral Research 
  35
Center  (JKMRC)  has  also  specialized  in  the  use  of  similar  methods  (Narayanan  and 
Whiten, 1988).   
 
Slow compression 
Slow compression tests are essentially compressive strength tests but with slightly faster 
loading  rates  (i.e.,  uniaxial  compressive  strength  test  might  be  run  at  0.001mm/s  while 
slow compression for single particle breakage might be 0.1 mm/s).  A particle is placed 
between two surfaces, an axial force is applied, and the particle is loaded until failure.  A 
typical set-up employs a materials testing system to measure the force and displacement 
(using a load cell and linear voltage displacement transducer). 
 
Slow  compression  was  developed  as  a  method  of  single  particle  breakage  analysis 
because it breaks particles in the most efficient manner possible.  No more energy than is 
required  to  fracture  the  particle  is  applied.    Bergstrom  conducted  experiments  in  which 
95%  of  the  applied  energy  was  determined  to  be  strain  energy  being  absorbed  by  the 
particle (Bergstrom et al., 1962; Bergstrom and Sollenberger, 1962).   
 
Although  some  researchers  have  conducted  slow  compression  SPB  tests  on 
cylindrical  core  specimens,  that  particle  shape  is  not  representative  of  what  is  seen  in 
commercial  applications,  and  due  to  the  large  loading  surface  area  of  a  cylinder  the 
induced stress field is not that same as seen under point contact loading.  Most research 
using slow compression has been conducted using spheres or irregularly shaped particles 
(Hanish  and  Schubert,  1986;  Yashima  and  et  al.,  1979;  Arbiter  et  al.,  1969).    In  those 
cases the particle size ranged from 2.5 mm up to 30 mm.  
 
Data  collection  is  straightforward.    The  energy  required  for  fracture  is  determined 
from  load-displacement  data  collected  during  testing.    The  fragmented  particle 
distribution  is  determined  from  sieve  analysis  of  the  tested  particles.    For  irregular 
particles  it  is  difficult,  if  not  impossible,  to  determine  the  particle  strength.    For  spheres 
elastic  theory  can  be  used  to  relate  the  applied  load  to  the  stress  field  induced  in  the 
particle. 
 
  Slow  compression  tests  have  given  way  to  impact  tests  such  as  the  drop  weight, 
pendulum,  and  falling  media  because  of  a  need  to  achieve  higher  deformation  rates  and 
because  breakage  devices  that  supply  excess  energy  are  more  analogous  to  commercial 
applications.  A great deal of research has focused on single particle breakage relative to 
ball  mills  and  grinding  mills,  thus  the  development  of  various  impact  crushing  tests.  
Despite  the  misnomer  as  slow  compression  machines,  jaw  crushers  apply  loads  over  a 
time  period  of  between  0.2  to  0.3  seconds  (based  on  typical  operating  rpms).    The 
corresponding rate of deformation (dependent also on the stroke) is much faster relative 
to  previous  slow  compression  SPB  research,  and  can  more  readily  be  simulated  using 
impact type single particle breakage tests. 
 
  36
Ultrafast load cell 
The  Ultrafast  Load  Cell  (UFLC)  was  developed  in  order  to  mimic  the  loading  and 
fracture  of  particles  in  a  tumbling  mill,  which  occurs  within  a  very  short  period  of  time 
and from a wide range of drop heights (Hofler, 1990).  The device was developed at the 
Comminution Center at the University of Utah (Weichert and Herbst, 1986).   
 
  The UFLC has a sampling rate of 2 s enabling it to pick up data from impacts that 
typically  occur  over  a  period  less  than  1500  s.    The  UFLC  uses  the  propagation  of 
elastic  waves  to  measure  the  force  acting  on  a  particle  under  impact  loading.    A  drop 
weight  impacts  a  particle  resting  on  top  of  a  steel  rod  resulting  in  compressive  strain 
wave that propagates through the rod (Figure 2.16).  Strain gages on the rod are used to 
transmit a signal to a digital oscilloscope.  A computer is then used to calculate the force 
versus time trace.   
Figure 2.16 Ultrafast load cell configuration
(From Herbst and Lo, 1992) 
 
The  energy  consumption  is  determined  in  the  same  manner  as  a  slow  compression 
test,  by  integrating  the  force-displacement  plot.    The  fragmented  particles  are  collected 
and sieved in order to determine the progeny size distribution.  The particle strength can 
be  determined  from  the  maximum  force  applied  (assuming  the  particle  is  essentially  a 
sphere). 
 
  37
The UFLC has been employed successfully in SPB testing.  Bourgeois (et al., 1992) 
broke  100  particles  in  the  3.35  to  4  mm  range  in  order  to  determine  the  mass  specific 
breakage energy, breakage function, and particle strength.  But these tests are relevant to 
ball  mill  conditions  and  there  has  been  no  application  of  the  UFLC  to  slower  velocity 
impacts seen in crushers. 
 
Hopkinson pressure bar 
The  Hopkinson  Pressure  Bar  (HPB)  was  originally  employed  to  simulate  the  energy 
levels commonly encountered in cone crushers (Napier-Munn et al., 1999).  The purpose 
of  the  HPB  is  to  enable  the  resolution  of  the  force  experienced  by  a  rock  as  it  is 
dynamically  loaded  in  compression  (Briggs  and  Bearman,  1995).    It  is  also  used  to 
measure the energy required to initiate failure in a particle. 
 
  The  HPB  is  similar  to  the  UFLC  but  is  aligned  horizontally.    Figure  2.17  shows  the 
test  set-up.    A  horizontally  aligned  steel  bar  with  a  rock  sample  attached  on  its  end  is 
impacted by a smaller bar.  A spring mounted on the small bar is compressed to a known 
distance allowing for control of the impact velocity.  An optical sensor is used to measure 
the speed of the impact bar at the point of impact.  As with the UFLC strain gage bridges 
are used to resolve the force versus time trace.  
Figure 2.17 Hopkinson pressure bar experimental set-up 
(From Briggs and Bearman, 1995) 
 
Since  both  bars  are  fitted  with  strain  gages,  the  lost  strain  energy  can  be  subtracted 
from  the  input  energy,  allowing  for  the  calculation  of  actual  fracture  energy.  
Additionally  the  particle  strength  can  be  determined  from  the  time  at  which  a  sudden 
drop  in  force  is  observed  on  the  force  versus  time  plot.    Currently  the  HPB  is  used  to 
determine the energy required for breakage and although it has not been used specifically 
for single particle breakage analysis it may aid in crusher design (Napier-Munn, 1999). 
 
  Other investigators have employed HPB type methods.  Shockey (et al., 1974) used a 
gas  gun  to  develop  an  approach  for  predicting  the  fragment  size  distribution  in  rock 
resulting  from  dynamic  loads.      Santurbano  and  Fairhurst  (1991)  employed  a  similar 
  38
device  in  order  to  develop  a  mechanistic  explanation  of  fragmentation  by  impact.    The 
rock  gun  uses  pneumatic  air  instead  of  a  spring  to  accelerate  a  rock  particle  down  a 
hollow cylinder from which it discharges and impacts a Hopkinson bar.  The velocity of 
the  particle  is  determined  using  paired  photocells  and  reflectors.    The  study  by 
Santurbano  and  Fairhurst  considered  the  energy  required  for  fracture  and  observations 
were  made  about  the  resultant  fracture  patterns  but  no  determination  of  fragment  size 
distribution was made.   
 
Drop weight tests 
Drop weight tests are the simplest of the single particle breakage tests.  Various types of 
drop weight devices have been employed due to their simple configuration (Fairs, 1954; 
Arbiter  et  al.,  1969;  Jomoto  and  Majima,  1972;  Jowett  and  Van  Der  Waedern,  1982; 
Pauw and Mare, 1988).   
 
A  weight  is  dropped  from  a  known  height  onto  a  particle  resting  on  a  hard  surface.  
The  kinetic  energy  is  determined  from  the  mass  of  the  drop  weight,  the  height  from 
which it was dropped, and gravity.  The potential energy of the weight is transmitted to 
the particle, and if high enough the particle fractures.  By adjusting the drop height and/or 
the  mass  of  the  drop  weight  the  input  energy  level  can  be  altered.    The  JKRMC  uses 
combinations  of  drop  height  and  drop  weight  mass  that  represent  an  energy  range  of 
0.001 to 50 kWh/t.   Figure 2.18 depicts the JKRMC drop weight test set-up.  
Figure 2.18 JKRMC drop weight test set-up 
(From Napier-Munn, 1999) 
  39
Since  the  drop  weight  comes  to  rest  on  top  of  crushed  fragments  after  breakage,  the 
actual applied energy is the initial potential energy of the drop weight minus the potential 
energy  it  retains  at  the  offset  height.    The  progeny  fragments  are  retained  for 
determination of cumulative size distribution.     
 
Pendulum tests 
Pendulum  tests  employ  twin  pendulums.  The  input  pendulum  is  released  from  a  certain 
height and impacts a particle attached to the rebound pendulum.  Figure 2.19 illustrates a 
twin pendulum set-up.  The specific fracture energy is the calculated by determining the 
difference  in  the  energy  of  the  input  pendulum  before  impact  and  the  energies  of  the 
rebound and input pendulum after impact (Narayanan, 1985). 
Figure 2.19 Twin pendulum test set-up
Impact Pendulum
Rebound Pendulum
Rock Particle
 
Bonds  crushability  test  for  determining  the  work  index  is  a  twin  pendulum  test 
(section  2.1.2).    Gaudin  and  Hukki  (1946)  extended  the  use  of  the  pendulum  test  to 
determine  the  size  distribution  of  fragmented  particles.    At  the  JKRMC  two  twin 
pendulum  devices  were  developed  in  order  to  determine  the  energy  available  for 
breakage  and  the  resultant  size  distribution  for  a  wide  range  of  input  energy  levels 
(Narayanan, 1985).  A twin pendulum device was also developed by Allis-Chalmers, and 
found  to  accurately  predict  rock  crushability  and  crusher  product  size  distribution 
(Moore,  1982).  However,  Napier-Munn  (et  al.,  1999)  has  found  that  twin  pendulum 
devices are restricted in their energy and particle size range, and that operating the device 
is time consuming. 
 
2.4.3 DATA COLLECTION 
 
The  analysis  of  single  particle  breakage  is  concerned  with  characterizing  the  crushing 
characteristics  of  a  material  and  relating  those  characteristics  to  the  energy  required  to 
  40
fracture  the  material.    A  materials  crushing  characteristics  are  measured  using  some 
parameter  of  material  strength  or  resistance  to  fracture  and  the  fragmented  size 
distribution,  which  are  in  turn  used  to  develop  energy-based  relationships.    Schoenert 
(1991) has recommended that the fracture properties of a material should be characterized 
in  terms  of  three  fundamental  properties:  particle  strength,  mass  specific  breakage 
energy, and breakage fragment size distribution.  Ultimately these characteristic and their 
inter-relationships can be used, in conjunction with comminution machine parameters, to 
predict  and  optimize  the  performance  of  size  reduction  processes.    The  following 
discussion focuses on some of the parameters obtained through SPB used to describe the 
three fundamental fracture properties.  
 
Product size distribution 
Fracture  of  a  single  particle  results  in  a  collection  of  daughter  fragments.    A  number  of 
SPB tests are conducted on particles of the same material in order to group the fragments 
together  and  determine  their  size  distribution.    Figure  2.13  showed  the  distribution  of 
fragments that results from each of the three fracture mechanisms: cleavage, shatter, and 
abrasion (section 2.2.3).  Based on those distributions early researchers attempted to use 
mathematical  expressions  to  describe  size  distributions.    Gilvarry  (1961)  derived  a 
Poisson  distribution  based  on  the  Griffith  flaws  associated  with  fracture  of  a  particle.  
The Gilvarry equation for product size resulting from single particle fracture is: 
 
2 3
1
x x x
k j i
y e
     (
|   | |   |   |   |
   (    +   +
   |    |      |
   ( \   .   \   . \   .
   
=      [2.27] 
where, k is the average spacing of activated edge flaws 
    j
2
 is the average amount of surface containing one activated surface flaw 
    i
3
 is the average volume corresponding to an activated volume flaw. 
When  surface  and  volume  flaws  are  ignored  the  Gilvarry  equation  reduces  to  the  well-
known  Rosin-Rammler  distribution,  and  for  very  small  values  of  particle  size,  x,  it 
simplifies to the Schumann equation (Gilvarry and Bergstrom, 1962).   
 
Other  researchers  have  proposed  similar  mathematical  expressions.    Gaudin  and 
Meloy (1962) derived a size distribution expression for single particle fracture using the 
binomial  distribution.    Klimpel  and  Austin  (1965)  combined  the  approaches  of  Gilvary 
and Gaudin and Meloy and developed a general expression for product size distribution.  
Despite the attempts to describe the size distribution of a single particle fracture event on 
theoretical  bounds,  none  of  these  models  reproduce  the  size  distribution  and  most 
researchers now agree that empirically derived functions are adequate (King, 2001). 
 
Empirically  derived  expressions  for  product  size  resulting  from  SPB  are  called 
breakage  functions.    Kelsall  and  Reid  (1965)  used  a  breakage  function  that  produced  a 
nearly straight line on a log-log scale and that had a distribution modulus near 1.  Herbst 
and Fuersternau (1972) generated a distribution that deviated from a straight line but did 
find  that  fine  particle  sizes  were  characterized  by  the  Schumann  distribution  and  a 
distribution modulus of 0.72.  Based on these, and similar findings, the breakage function 
  41
has  been  shown  to  be  dependent  only  on  the  ratio  of  the  progeny  fragment  sizes  to  the 
initial particle size: 
  (   )
1
1
,
n
i
i
d
B d d
d
|   |
   |
\   .
  [2.28] 
where, B is the breakage function 
    d
i
 is the particle size 
    d
1
 is the initial particle size 
     n is the distribution modulus. 
The concept behind the breakage function is that the progeny population is made up of a 
mixture of separate populations (the populations occurring as a result of cleavage, shatter, 
and  abrasion)  and  that  it  is  independent  of  initial  particle  size.    Different  values  of  n 
account  for  the  larger  fragments  produced  by  induced  tensile  stresses,  the  smaller 
fragments produced by compressive stresses directly under the area of loading, and/or the 
fragments produced by abrasion. 
Figure 2.20 A typical breakage function and corresponding parameters
0.01 0.1 1
0.01
0.1
1
n
2
 = slope
n
1
 = slope
1-K
K
 
B'
d/d
1
 
A typical product size distribution resulting from SPB is shown in Figure 2.20.  The 
breakage  function  modeled  to  fit  the  distribution  in  Figure  2.20  has  two  components,  a 
fine  distribution  resulting  from  localized  fracture  of  the  particle  against  the  loading 
surface  and  a  coarse  distribution  resulting  from  main  fracture  of  the  particle  (Lynch, 
  42
1977).  For  a  2 size  fraction  the  breakage  function  of  Figure  2.20  can  be  fit  with  the 
following expression: 
  (   )   (   )
1 2
1
1 1
, 1
n n
i
i
d
B d d K K
d d
|   |   |   |
=   +   
   |      |
\   .   \   .
i
d
  [2.29] 
where K is the fraction of progeny fragments that contribute to the finer fraction.  Typical 
measured values of n
1
 lie between 0.6 and 1.3 while values of n
2
 have been shown to be 
in the range of 2.5 to 5 (Arbiter et al., 1969; Kelly and Spottiswood, 1990). 
 
If  a  measured  breakage  function  becomes  bimodal,  which  sometimes  occurs  with 
large  particles  that  have  a  tendency  to  chip  rather  than  shatter,  a  third  distribution 
modulus  can  be  used  to  properly  model  the  breakage  function.    Additionally,  if  size 
distributions  are  determined  for  various  parent  particles  of  the  same  material  and  the 
breakage  functions  do  not  normalize  (i.e.,  show  a  dependency  on  initial  particle  size), 
then the fraction K must be determined as a function of the parent size. 
 
  More recently the JKRMC has developed the one-parameter family of curves method 
to  represent  SPB  breakage  data  (Narayanan  and  Whiten,  1988).    The  parameter  is  the 
percent of progeny particles passing a size one-tenth of the initial original particle size, or 
t
10
.    t
10
  is  employed  as  a  characteristic  size  reduction  parameter  and  is  determined 
primarily  by  the  energy  absorbed  during  a  SPB  test.    For  crushing  applications  t
10
  is 
usually in the range of 10% to 20% and for tumbling mills t
10
 ranges between 20% and 
50% (Napier-Munn, et a., 1999).  Figure 2.21 shows a plot of the one-parameter family of 
curves for a given material.  Each t
n
 value (% passing 1/n of the original particle size) is 
0 10 20 30 40 5
0
20
40
60
80
100
t
n
 
 
%
 
P
a
s
s
i
n
g
t
10
  %
 t
2
 t
4
 t
10
 t
25
 t
50
 t
75
Figure 2.21 One-parameter family of curves 
(After Narayanan, 1985)
0
  43
uniquely related to t
10
.   Each vertical line represents a complete size distribution and if a 
value of t
10
 is given the full product size distribution can be reconstructed.  An alternative 
to  the  one-parameter  family  of  curves  is  to  use  standard  truncated  distribution  functions 
(i.e., Rosin-Rammler, logistic, log-normal) that are dependent on t
10
 and one or two other 
parameters (King, 2001).   
 
Fracture energy 
The  calculation  of  fracture  energy  is  dependent  upon  the  method  of  testing  but  in  each 
case  it  is  usually  determined  as  a  mass  specific  energy  (i.e.,  kWh/t).    There  are  few 
types of specific energy that should be differentiated.  The specific energy required to 
initiate fracture is called the specific breakage energy.  The specific breakage energy is a 
material property since it represents the energy stored in the particle due to elastic strain, 
and  it  represents  the  minimum  amount  of  energy  required  to  cause  fracture.  
Measurement  of  the  specific  breakage  energy  requires  the  use  of  sophisticated  devices 
like the Ultrafast Load Cell.  The specific comminution energy is the energy required to 
produce a certain amount of size reduction or the characteristic size reduction parameter 
t
10
.    The  specific  comminution  energy  is  determined  in  drop  weight  or  pendulum  tests 
where there is no way to discern the time of first fracture.  Specific comminution energy 
is  more  applicable  to  actual  operating  conditions  since  size  reduction  machines 
commonly supply excess amounts of energy.  
 
  SPB testing has indicated that the breakage function is related to the specific fracture 
energy and the specific comminution energy.  In general more energy results in a larger 
portion  of  daughter  fragments  in  the  fine  size  range.    This  is  particularly  true  with 
specific comminution energy since the excessive amount of energy input into breakage is 
dissipated  through  secondary  fracture.    Figure  2.22  illustrates  the  relationship  between 
specific comminution energy and product size (recalling that larger values of t
10
 indicate 
a finer size distribution).  From Figure 2.22 it can be also be seen that t
10
 can be related to 
specific comminution energy using the following function: 
 
(   )
10 10max
1
c
bE
t t e
=      [2.30] 
where, t
10max
 is a material specific limiting value of t
10
 
    E
c
 is the specific comminution energy 
b is a material specific parameter. 
Mehrim  and  Khalaf  (1980)  fit  their  test  results  with  a  similar  exponential  relationship 
between  size  distribution  and  energy  input.    They  also  found  that  the  relationship  was 
linear  at  lower  levels  of  input  energy  and  deviated  from  linear  at  high  levels  of  input 
energy  (similar to what is occurring in Figure 2.23).   
 
Bergrstrom  (et  al.,  1962;  1965)  found  that  the  specific  energy  is  inversely 
proportional  to  the  size  modulus  of  the  product  distribution  and  confirmed  that  the 
relationship  holds  for  impact  crushing  and  slow  compression,  as  well  as  for  irregularly 
shaped  particles  and  spheres.    Since  smaller  size  moduli  describe  finer  particle 
distributions this inverse relationship indicates that more applied energy produces smaller 
  44
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00
0
5
10
15
20
25
30
35
40
45
50
t
10
 = 49.1(1-e
-0.87E
c
)
t
10
(%)
Specific Comminution Energy (kWh/t)
Figure 2.22 Relationship between product size and specific comminution energy 
(After Napier-Munn, 1999) 
fragments  and  also  proves,  once  again,  that  smaller  particles  require  more  energy  for 
fracture  than  larger  ones.    Another  important  aspect  of  this  relationship  is  that  although 
increased  energy  results  in  a  finer  product  size  distribution,  only  the  proportion  of 
material in the fine size range is changing (i.e., K from Equation 2.29) not the distribution 
modulus  (n
1,2
  remain  unchanged).    Hanisch  and  Schubert  (1986)  confirmed  this  using 
slow compression tests, stating that the distribution moduli are essentially not influenced 
by  stress  conditions  or  comminution  energy,  and  that  only  the  mass  fractions  of  the 
product depend strongly on the comminution energy.   
 
Particle strength 
In  studies  of  single  particle  fracture  particles  break  in  tension  under  compressive  loads.  
This  is  in  agreement  with  the  breakage  processes  occurring  in  most  comminution 
machines,  particularly  crushers.    The  stress  distribution  determined  by  Oka  and  Majima 
(1970)  (Figure  2.3)  for  localized  compressive  loading  is  essentially  the  distribution 
determined by Hondros (1959) for the case of a disk subject to short strip loadings.  The 
maximum  tensile  strength  can  easily  be  determined  using  these  distributions  if  the 
applied load is known as well as the specimen dimensions. 
  
  For  disk  specimens  the  tensile  strength  can  be  determined  from  the  following 
equation: 
  45
 
2
t
P
Dt
=   [2.31] 
where, P is the compressive load, at failure, applied over a small, localized area 
    D is the disk diameter 
    t is the disk thickness. 
It  is  more  common  in  SPB  testing  to  test  irregularly  sized  particles.    In  this  case  the 
tensile  stresses  cannot  be  determined  from  elastic  theory.    It  is  assumed  that  the  contact 
area  is  a  certain  fraction  of  the  particle  cross-section  and  that  the  stress  resulting  in 
fracture is simply the compressive force acting on that fractional area.  Equation 2.29 is 
then used to determine particle strength. 
 
2
4
t
P
D
=   [2.32] 
where, D is the particle diameter.  Hiramatsu and Oka (1966) used photoelastic methods 
to show that a sphere, prism, and cube all have a similar stress states when subjected to a 
pair of concentrated compressive forces.  They obtained the following expression for the 
tensile strength of an elastic sphere: 
 
2
2.8
t
p
P
D
=   [2.33] 
In this case D
p
 is the distance between the loading contacts but in practice it is assumed to 
be  equal  to  the  particle  diameter.    Jomoto  and  Majima  (1972)  and  Oka  and  Majima 
(1970) have suggested similar equations for determining the tensile strength of irregular 
particles. 
 
  Experimental results indicate that particle strength is related to both energy input and 
the  breakage  distribution.    Yashima  (et  al.,  1979)  found  that  the  reciprocal  of  the  size 
modulus of the fragmented product size was proportional to the compressive strength of 
spheres  by  a  power  of  1.7  (for  slow  compression  testing).    Jomoto  and  Majima  (1972) 
suggested,  based  on  drop  weight  tests,  that  the  square  of  the  tensile  strength  is 
proportional to energy input and it could be used as a useful criterion under conditions of 
impact  crushing.    Shockey  (et  al.,  1974)  measured  both  dynamic  tensile  strength  and 
fracture toughness of Arkansas novaculite and found that making quantitative predictions 
of  rock  fragmentation  based  on  the  knowledge  of  a  few  determinable  rock  strength 
properties is feasible.   
 
2.4.4 CONCLUSIONS 
 
Single  particle  breakage  analysis  provides  a  method  for  studying  the  most  fundamental 
process  occurring  in  a  size  reduction  machine.    It  is  a  physical  representation  of  the 
fracture  physics  associated  with  the  breakage  of  a  particle  under  loading  conditions 
typically seen in comminution machines.  There are a multitude of factors that affect even 
the breakage of one single particle and they are inter-dependent.  The loading condition is 
dependent  on  the  machine  type,  which  controls  the  number  and  direction  of  contact 
forces,  the  deformation  velocity,  and  the  stress  state  of  the  particle.    The  stress  field 
  46
induced in the particle also depends on its shape, the presence and distribution of flaws, 
its  homogeneity,  and  the  particles  stress-strain  behavior.    An  understanding  of  how  to 
properly  represent  single  particle  breakage  at  the  experimental  level  can  be  acquired  if 
proper consideration is given to these factors.  
 
  Several methods exist for the testing of single particles.  Slow compression tests were 
the  first  tests  to  be  used  and  have  now  given  way  to  tests  capable  of  applying  large 
amounts  of  energy  at  high  impact  velocities  (or  deformation  rates).    The  Ultrafast  Load 
Cell,  Hopkinson  Pressure  Bar,  pendulum  tests,  and  drop  weight  tests  are  the  most 
common.  The drop weight tests are the most simplistic but do allow for a wide range of 
input  energies  to  be  applied  representing  basically  all  comminution  applications.    The 
Ultrafast  Load  Cell  has  the  ability  to  collect  data  very  rapidly  and  allows  for  the 
differentiation between the energy required to initiate fracture and the total energy input. 
 
  The purpose of SPB analysis is to characterize a materials crushability and relate it 
to  the  energy  required  to  fracture  the  material.    A  materials  crushability  is  described 
using  the  breakage  function,  an  empirically  derived  function  that  expresses  the 
distribution of progeny particles that result form particle fracture.  The breakage function 
is  considered  to  be  independent  of  initial  particle  size  and  is  a  mixture  of  separate  size 
populations.    When  its  discrete  values  are  put  in  matrix  form  it  can  be  used  to  help 
predict/optimize  the  product  size  emerging  from  an  actual  crushing  machine.    The 
JKRMC has popularized the use of t
10
, the characteristic size parameter.  For a given t
10
 
the one-parameter family of curves can be used to develop the entire breakage function of 
a  material.    t
10
  can  also  be  used  with  standard  truncated  size  distributions  in  order  to 
develop a full product size distribution. 
 
  t
10 
is  determined  primarily  by  the  energy  absorbed  during  a  single  particle  breakage 
test, as are the proportions of  coarse and fine material.  The specific energy is inversely 
proportional  to  the  size  modulus  of  the  product  distribution  but  has  no  effect  on  the 
distribution modulus.  Thus for a given material an increase in energy intensity changes 
only  the  proportion  of  fine  and  coarse  fragments  in  the  progeny  distribution.    An 
exponential  relationship  has  been  suggested  for  determining  the  size  distribution,  or  t
10
, 
resulting from a certain level of specific comminution energy. 
 
  Schoenert  has  proposed  that  the  fracture  properties  of  a  material  should  be 
characterized  in  terms  of  three  fundamental  properties:  particle  strength,  mass  specific 
breakage  energy,  and  breakage  fragment  size  distribution.    Despite  the  realization  that 
some  measure  of  a  materials  resistance  to  fracture  should  be  intimately  related  to  a 
particles  behavior  in  SPB,  there  have  been  only  minor  attempts  to  include  particle 
strength along with other SPB data.  Most of these attempts have focused on the tensile 
strength since tensile stresses control breakage in comminution processes.  Even though 
there is evidence that the nature of a material is related to the energy required to fracture 
it  and  the  resultant  breakage  distribution,  it  is  generally  thought  that  standard  rock 
mechanics  tests  or  properties  do  not  provide  useful  information  for  the  design  and 
  47
optimizations of size reduction processes.  But the role of standard tests and properties in 
understanding and optimizing comminution is likely to increase in the future if a link can 
be  made  between  a  materials  crushability  as  described  by  SPB  analysis  and  its 
mechanical response to fracture.   
 
  The  breakage  function  determined  from  SPB  testing  is  dependent  upon  the  energy 
input  and  the  nature  of  the  material  broken.    Since  the  energy  input  of  a  comminution 
device  is  related  to  the  method  and  rate  of  load  application,  the  breakage  function 
parameters  are  most  likely  dependent  upon  some  operational  characteristics  (such  as 
closed side set and speed or frequency), as well as the nature of the material broken.  A 
rapid  method  of  establishing  the  breakage  function  for  a  given  crushing  process  and 
material  may  be  possible  if  SPB  is  used  to  determine  the  breakage  function  and  the 
energy input for a certain set of operational settings, and those measures are related to an 
inherent property that describes a materials resistance to fracture. 
 
 
2.5 ROCK FRACTURE MECHANICS AND FRACTURE TOUGHNESS 
 
Modifications  to  Griffiths  theory  have  led  to  the  development  of  the  field  of  fracture 
mechanics.    Fracture  mechanics  deals  with  facture  initiation  and  crack  propagation,  and 
provides  quantitative  methods  for  characterizing  the  behavior  of  an  intact  material  as  it 
fractures  due  to  crack  growth.    The  extension  of  fracture  mechanics  to  rock  is 
understandable  since  rock  masses  contain  cracks  and  discontinuities.    States  of  stress 
around  these  flaws  cannot  be  predicted  using  macroscopic  failure  criteria  (i.e.  Mohr-
Coulomb,  ultimate  strength  theories).    In  order  to  deal  with  crack  propagation, 
particularly  in  terms  of  intentional  fracturing  as  in  size  reduction  processes,  rock 
fracture mechanics must be used.   
 
Although fracture mechanics has an undeniable place in rock mechanics applications, 
it  was  not  developed  for  geomaterials.    It  should  be  recognized  that  differences  exist 
between  fracture  mechanics  for  man-made  materials  (metals)  and  rock  fracture 
mechanics,  particularly  in  basic  material  response  and  engineering  application.  
Whittaker  (et  al.,  1992)  gave  a  comprehensive  list  and  explanation  of  these  differences, 
which can be summarized as: 
1.  Stress  state    Many  rocks  structures  are  subjected  to  compressive  stresses  as 
opposed  to  tensile  stresses.    However,  in  comminution  and  crushing  the 
induced  stress  state  is  tensile  (from  point-load  compression)  and  thus  tensile 
fracture is seen in rock. 
2.  Rock  fracture    Rock  materials  usually  fracture  in  a  brittle  or  quasi-brittle 
manner and usually do not exhibit plastic flow. 
3.  Fracture  process  zone  (FPZ)    Non-elastic  behavior  ahead  of  a  crack  tip  in 
rock  takes  the  form  of  micro-cracking  as  opposed  to  excessive  shear  stresses 
and the resultant plastic process zone seen in metals.  If the size of the FPZ is 
small then linear elastic fracture mechanics applies. 
  48
4.  Crack  surface    Crack  surfaces  in  rock  can  be  non-planar  with  friction  and 
inter-locking  occurring,  but  linear  elastic  fracture  mechanics  assumes  that  no 
forces are transmitted across the surface of a smooth planar crack 
5.  Crack  propagation    In  rocks  there  is  a  tendency  for  crack  propagation  to 
wander  along  grain  boundaries  or  planes  of  weakness.    The  area  of  newly 
created surface is then larger then the assumed fracture area. 
6.  Rock fracture mechanics applications  In rock mechanics, as in (man-made) 
materials  engineering,  the  prevention  of  failure  by  fracture  growth  is  a 
concern.  But the optimizing the generation and propagation of cracks is also a 
concern  as  in  size  reduction  processes.  Thus  the  application  dictates  how 
material parameters should be determined and used. 
7.  Influence  of  scale    Due  to  the  complicated  geologic  nature  of  rock  masses, 
the  characterization  of  a  rock  mass  is  high.    For  the  prevention  of  crack 
growth  and  failure,  parameters  measured  experimentally  are  of  secondary 
importance but for rock fragmentation applications, experimentally measured 
properties are of primary importance. 
8.  Heterogeneity    Changes  in  local  structure  and  strength  ahead  of  a  crack  tip 
affects the continuity of crack growth. 
9.  Presence of discontinuities  Pre-existing discontinuities affect the local stress 
states and crack propagation. 
10. Anisotropy  Rocks can be anisotropic affecting measured fracture parameters 
as a function of crack orientation.   
Recognition  of  these  variations  has  led  to  more  practical  and  developed  concepts  of 
fracture mechanics as it applies to rock behavior, with principles of linear elastic fracture 
mechanics being extended even to rocks that behave non-linearly and much of the focus 
centering on the measurement of fracture toughness. 
 
The  most  fundamental  aspect  of  rock  fracture  mechanics  is  the  establishment  of  a 
relationship  between  rock  fracture  strength  and  the  geometry  of  the  flaws  that  result  in 
fracture.    Through  this  relationship  an  intrinsic  material  property  that  describes  a 
materials  resistance  to  crack  propagation  can  be  measured.    This  property  is  called 
fracture  toughness.    The  application  of  fracture  toughness  in  size  reduction  processes  is 
clear.    Fracture  toughness  represents  a  critical  level  above  which  crack  extension  and 
fracture occurs.  When individual rock particles are subjected to the applied forces of size 
reduction,  it  is  most  likely  that  the  intrinsic  tensile  property  measured  as  the  fracture 
toughness will control breakage (Bearman, 1998).  Since the amount of energy input into 
a size reduction process and the amount of size reduction achieved (i.e., the fractured size 
distribution) are related to the type of loading and the crack pattern in the material, there 
should be a relationship between these parameters and fracture toughness.   
 
2.5.1 LINEAR ELASTIC FRACTURE MECHANICS 
 
In  section  2.2.2  Equation  2.20  was  presented  as  a  size  independent  expression 
characterizing  the  resistance  of  a  material  to  fracture  based  on  the  stresses  required  for 
  49
breakage and the presence of flaws.  It showed that fracture initiation in a brittle solid is 
controlled  by  the  product  of  a  far-applied  stress  and  the  square  root  of  the  flaw  length 
which  reaches  a  critical  value  determined  by  the  characteristic  material  properties  E,  , 
and 
s. 
 This critical value is called the critical stress intensity factor and is denoted K
c
.   
 
  Irwin (1957) used a stress intensity approach to relate the critical strain energy release 
rate  G
c
  to  the  critical  stress  intensity  factor  K
c
.    Rather  than  follow  Griffiths  global 
approach,  Irwin  considered  the  crack  tip  region,  which  is  small  compared  to  the  rest  of 
the  body  (or  plate,  in  reference  to  section  2.2)  but  large  enough  with  respect  to  atomic 
dimensions  such  that  linear  elastic  theory  applies  (Knott,  1972).    Irwin  determined  the 
work  required  to  close  up  a  small  portion  of  a  crack  by  superimposing  tensile  forces 
along the crack surfaces and hypothesized that this work is equal to the energy released 
when the crack extends.  Thus the work required to close a unit length of the crack is the 
strain  energy  release  rate  and,  based  on  the  stresses  and  displacements  occurring  as  a 
result of the tensile forces, is equal to: 
 
(   )
2 2
1 K
G
E
 
=   [2.34] 
Since  crack  propagation  occurs  when  G  reaches  a  critical  value,  the  critical  value  of 
stress intensity can be defined as: 
 
(   )
2
1
c
c
G E
K
  [2.35] 
By  demonstrating  the  equivalence  of  K  and  G,  Irwin  provided  the  basis  for  the 
development  of  Linear  Elastic  Fracture  Mechanics  (LEFM).    In  LEFM  the  crack  tip 
stresses, strains, and displacements can be characterized by K as long as inelastic yielding 
ahead  of  the  crack  tip  is  small.    The  advantage  of  LEFM  is  that  it  provides  a  universal 
approach for determining a materials resistance to fracture, as defined by K
c
.  As long as 
an  explicit  function  for  the  stress  intensity  near  a  crack  tip  is  known  for  a  given  crack 
geometry and loading configuration, K
c
 can be measured experimentally. 
 
2.5.2 STRESS INTENSITY FACTOR  
 
The stress intensity factor K, alluded to in the previous section, characterizes the severity 
of  the  crack  condition  as  affected  by  crack  dimension,  stress,  and  geometry  (Dowling, 
1999).    Determining  K  is  based  on  a  linear-elastic  approach  (hence  LEFM),  which 
assumes the material in which the crack is located is isotropic and behaves according to 
Hookes Law.   
 
Different  loading  configurations  at  a  crack  tip  lead  to  different  modes  of  crack  tip 
displacement.  The different types of crack deformation are generalized using three basic 
modes  (Figure  2.23).    Mode  I  is  the  opening  mode  due  to  tension,  where  the  crack 
surfaces  move  directly  apart;  Mode  II  is  the  sliding  mode  due  to  shearing,  where  the 
crack  surfaces  move  over  one  another  in  a  direction  perpendicular  to  the  crack  front; 
  50
Mode III is the tearing mode also due to shearing, where the crack surfaces sliding over 
one another but in a direction parallel to the crack front.  The three basic modes can also 
occur  in  combination  as  mixed-mode  loading  with  the  superposition  of  the  modes 
sufficient  to  describe  most  general  three-dimensional  cases  of  local  crack  tip  stress  and 
deformation fields (Tada et al., 2000).  Mode I is the most commonly encountered mode 
in engineering applications and is also the easiest to analyze, produce experimentally on 
laboratory specimens, and apply (Schmidt and Rossmanith, 1983). 
Figure 2.23 The three basic modes of crack surface displacement 
(After Tada et al., 2000) 
 
  Using  theory  of  elasticity,  namely  the  stress  analysis  methods  of  Muskhelishvili 
(1963)  and  Westergaard  (1939),  the  crack  tip  stress  and  displacement  fields  (and  hence 
K)  for  each  mode  of  loading  can  be  determined  (for  a  complete  derivation  see  Pook, 
2000).    With  Figure  2.24  representing  the  coordinate  system  measured  from  the  leading 
edge  of  a  crack,  the  Mode  I  stress  components  are  given  according  to  the  following 
equations: 
 
(   )
3
cos 1 sin sin
2 2 2
3
cos 1 sin sin
2 2 2
3
cos sin cos
2 2 2 2
, for plane strain
0, for plane stress
0
I
x
I
y
I
xy
z x y
z
xz yz
K
r
K
r
K
r
2
2
      
   
   (
=   
   (
   
   (
=   +
   (
   
   (
=
     (
   
=   +
=
=   =
  [2.36] 
where K
I
 is the stress intensity factor for Mode I.  The displacements at the crack tip can 
be found by substituting Equations of 2.36 into Hookes Law.  
  51
Figure 2.24 Coordinate system for a crack tip
(After Tada et al., 2000)
It  can  be  seen  from  Equations  2.36  that  at  the  crack  tip  (as  r  approaches  zero)  the 
stresses  approach  infinity,  as  has  already  been  indicated  by  Inglis  solution  for  stresses 
around an elliptical hole in a stressed plate (section 2.2.1).  Since no value of stress at the 
crack tip can be given, and all non-zero stresses of Equation 2.36 are proportional to K
I,
 
with  the  remaining  factors  varying  only  with  r  and  ,  the  stress  field  near  the  crack  tip 
can be determined by giving the value of K
I
, which has a formal definition of (Dowling, 
1999; Pook, 2000): 
 
, 0
lim 2
I r y
K
  
r    
=   [2.37] 
It was noted earlier that K
I
 is affected by the crack size, stress, and geometry.  In order to 
account for different geometries Equation 2.37 can be rewritten as: 
  2
I
K F a     =   [2.38] 
where, F is a dimensionless constant dependent on the geometric configuration 
     is the stress averaged over the gross area 
    a is the half-crack length. 
F  can  generally  be  described  as  a  function  of  loading  geometry  and  a w  where  w  is 
defined  as  the  maximum  possible  crack  length.    When  F  is  determined  for  a  given 
geometry the critical value of stress intensity, or fracture toughness, can be determined as 
long as inelastic yielding ahead of the crack tip is small and the conditions for LEFM are 
met.    Equations  and  values  of  F  for  a  wide  range  of  crack,  specimen,  and  loading 
geometries  are  determined  using  analytical,  numerical,  and  experimental  methods  and 
have been compiled in various handbooks (see Tada et al., 2000; Murakami, 1987; Rooke 
and Cartwright, 1976; Sih, 1973).   
 
  52
2.5.3 MODE I FRACTURE TOUGHNESS TESTING 
 
The  critical  value  of  stress  intensity  factor,  K
Ic
,  is  determined  by  testing  a  prepared 
specimen  that  has  a  crack  in  it.    There  are  no  standardized  test  methods  for  the 
measurement  of  Mode  I  rock  fracture  toughness.    The  earliest  applications  of  rock 
fracture  roughness  testing  employed  the  ASTM  standard  method  (ASTM-E399)  for 
metals.    Although  ASTM-E399  seemed  to  be  effective  for  rock  testing,  the  general 
consensus for rock testing has become that an ideal method would yield a representative 
fracture  toughness  value  and  yet  be  simple,  requiring  neither  pre-cracking,  nor  crack 
length  and  displacement  measurements,  nor  sophisticated  evaluation  techniques 
(Ouchterlony,  1989).    In  reference  to  these  requirements  ASTM-E399  is  cumbersome, 
and most testing methods for rock now employ core-based specimens.  The International 
Society  of  Rock  Mechanics  (ISRM)  has  suggested  two  methods  be  established  as 
standardized tests in order to obtain accurate, compatible, and reproducible K
Ic
 values for 
rocks (ISRM, 1988). 
 
  Before  covering  the  actual  test  methods  used  to  determine  fracture  toughness, 
consideration  needs  to  be  given  to  some  factors  that  influence  the  applicability  of 
measured values and the testing procedures. 
 
Fracture process zone 
The  application  of  linear  elasticity  has  been  shown  to  be  valid  even  in  cases  of  plastic 
yielding ahead of the crack tip, as long as the amount of yielding is small relative to the 
geometry  of  the  crack  and  other  characteristic  dimensions  (specimen  thickness,  length).  
When the non-elastic field is large and dominates the crack behavior before failure, then 
LEFM is not applicable and other methods of fracture analysis must be employed.  Based 
on  the  size  of  the  non-elastic  field,  specimen  and  crack  size  requirements  have  been 
developed  to  ensure  that  measured  values  of  fracture  toughness  are  not  underestimated 
(as is the case when LEFM is applied to non-linear conditions). 
 
  The  fracture  process  zone  (FPZ)  in  rocks  is  the  region  ahead  of  the  crack  tip  that 
experiences non-linear behavior.  The FPZ is formed by the initiation and propagation of 
micro-cracks in the vicinity of the crack tip (Figure 2.25).  Models used to determine the 
size  and  shape  of  the  FPZ  are  based  on  the  models  used  to  describe  the  plastic  zone  in 
metals.  Schmidt (1980) used a maximum normal stress criterion to describe the shape of 
the FPZ and found that the size of the FPZ can be given as: 
  (   )
2 2
2
1
cos 1 sin
2 2
I
t
K
r
2
   
   
|   |   |
=
     |   
\   . \   .
  |
+
  |
  [2.39] 
where, 
t
 is the tensile strength of the rock material. 
    r,  are defined by the coordinate system in Figure 2.26. 
An illustration of the FPZ is shown in Figure 2.26.  The characteristic size of the FPZ is 
defined when  is equal to zero and is identical to the plastic zone in metals under plane 
stress conditions.   The maximum size of the FPZ is defined when  is equal to 60. 
  53
Figure 2.25 Development of the FPZ 
(After Hoagland et al., 1973 and Whittaker et al., 1992) 
Based  on  the  previous  analysis  the  size  and  shape  of  the  fracture  process  zone  is 
independent  of  either  plane  stress  or  plane  strain  conditions.    Nolen-Hoeksema  and 
Gordon  (1987)  proved  this  experimentally  using  dolomite,  showing  that  the  FPZ  is  the 
same  for  a  crack  on  a  free  surface  (plane  stress)  and  for  a  crack  located  within  the  rock 
(plane  strain).    In  comparison  to  the  plastic  zone  of  metals,  the  FPZ,  and  fracture 
toughness  accordingly,  is  not  affected  by  specimen  thickness  (but  it  cannot  be 
subjectively  small).    It  is  however  dependent  upon  the  crack  length  and  ligament  width 
(uncracked length).  Lim (et al., 1994) compiled a list of minimum crack lengths required 
  54
Figure 2.26 Size and shape of the FPZ
(After Schmidt, 1980)
for  valid  fracture  toughness  testing  of  different  rocks,  citing  that  minimum  crack  length 
depends  on  both  material  type  and  testing  technique.    However  it  is  generally  accepted 
that  Equation  2.39  along  with  aspects  from  the  plastic  zone  of  metals  can  be  used  to 
define the specimen dimensions required to give representative fracture toughness values 
(Schmidt, 1980; Barton, 1983).  The requirements are: 
 
2
2
2.5
27
32
Ic
t
Ic
t
a
K
w a
K
t
   
|   | 
` |
  )   \   .
|   |
     |
\   .
  [2.40] 
where, w-a is the uncracked length 
    t is the specimen thickness. 
The  requirements  of  Equation  2.40  are  conservative,  and  other  factors  have  been 
proposed,  but  in  general  if  the  requirements  given  are  not  met  the  measured  fracture 
toughness  value  (i.e.,  the  apparent  fracture  toughness)  is  considered  invalid,  or  outside 
the bounds of LEFM. 
 
  Micro-cracking  ahead  of  the  crack  tip,  and  subsequent  FPZ  development,  is 
influenced  by  the  microstructure  or  grain  size  of  the  rock.    Minimum  ratios  of  the 
smallest  specimen  dimension  to  the  grain  size  have  been  suggested  in  order  to  make 
certain  representative  values  of  K
Ic
  are  measured.    A  minimum  ratio  of  10:1  is 
  55
recommended  by  the  ISRM  for  the  two  suggested  standard  methods  of  testing  (ISRM, 
1988).   
 
Crack requirements 
In  order  to  determine  fracture  toughness  a  specimen  needs  to  be  cracked.    The  crack 
should be representative of a natural crack, one that is sharp and free from the effects of 
residual  stresses,  specimen  boundaries,  and  pre-fabricated  notches  (Whittaker  et  al., 
1992).    In  metals,  specimens  are  usually  pre-cracked  by  fatigue  in  order  to  obtain  a 
naturally sharp crack.  In rocks fatigue pre-cracking is difficult since the loads required to 
produce  fatigue  crack  growth  are  usually  too  high  (they  approach  catastrophic  failure 
levels) and care is required in order to prevent further propagation of the pre-crack after it 
has been initiated.   
 
Other  methods  are  required  to  pre-crack  rock  materials,  the  simplest  being  to  notch 
the rock with a thin saw-cut.  Despite the ease of notching there is evidence that notched 
specimens  tend  to  underestimate  fracture  toughness.    Sun  and  Ouchterlony  (1986) 
showed  that  apparent  fracture  toughness  values  measured  using  notched  specimens  are 
invariably  lower  than  those  measured  from  pre-cracked  specimens  (for  the  same  rock 
type),  concluding  that  the  use  of  notch  length  in  the  calculation  of  K
Ic
  ignores  micro-
crack  growth  prior  to  crack  extension.    Fenghui  (2000)  has  proposed  a  model,  based  on 
the size of the FPZ and notch radius, that can be used to determine the fracture toughness 
based  on  the  measured  fracture  toughness  of  a  notched  specimen  as  long  as  the  notch 
radius is not greater than the average grain size of the rock. 
 
  The  most  widely  accepted  method  of  pre-cracking  rocks  employs  a  chevron-notch.  
The  chevron-notch  is  a  V-shaped  notch  that  allows  the  length  of  the  crack  front  to 
increase as the crack propagates.  Further propagation of the crack requires an increase in 
the  load,  which  is  the  condition  of  stable  propagation,  and  inherent  monotonic  pre-
cracking is produced during testing (Sun and Ouchterlony, 1986).  A sharp natural crack 
is automatically formed and the resistance to propagation becomes fully developed after 
initial  crack  growth.    But  the  chevron-notch  is  not  particularly  easy  to  produce  in 
comparison to standard or straight-through saw-cut notches, and the decision to use one 
or the other is based on the level of testing required. 
 
Testing level 
The  ISRM  (1988)  has  defined  two  testing  levels  in  order  to  let  researchers  decide  what 
combination of screening and accuracy is best for the given application of the measured 
fracture  toughness  values.    Level  I  testing  can  be  performed  using  portable  equipment 
and  requires  only  the  registration  of  maximum  load.    The  associated  value  of  fracture 
toughness has the nature of an index property more so than of a material property.  Level 
I  testing  is  more  appropriate  for  screening  purposes  or  for  the  rapid  estimation  and 
comparison of fracture toughness values. 
 
  56
  Level II testing requires load and displacement measurements, and thus is laboratory 
based  and  intrinsically  more  complicated  to  perform.    It  is  recommended  for  the 
determination of accurate, compatible, and reproducible fracture toughness values. 
 
Loading rate 
Various loading rates have been prescribed for rock fracture testing, ranging from 0.01 to 
0.03 MPa m s (Barton,  1983).    For  the  ISRM  recommended  tests,  the  loading  rate  is  not 
supposed  to  be  greater  than  0.25 MPa m s or  such  that  failure  occurs  within  10  seconds 
(ISRM, 1988).  Although rock strength properties exhibit loading rate dependency, there 
is no consensus on whether or not K
Ic
 is affected by an increase in loading rate.  Based on 
the FPZ it is logical to expect that dynamic loading will result in underestimated values of 
fracture  toughness  since  fracture  will  occur  before  the  FPZ  is  fully  developed.    But  by 
testing  in  accordance  with  the  recommendations  of  the  ISRM  it  has  been  shown  that 
loading rate has a negligible effect on measured fracture toughness (Khan and Al-Shayea, 
2000).  
 
Calculating fracture toughness 
A  materials  resistance  to  crack  propagation  can  be  quantified  using  fracture  toughness, 
K
Ic
, the critical stress intensity factor.  When the stress intensity factor, K
I
, in the region 
of a crack tip exceeds K
Ic
, fracture initiates and propagates until the stress intensity factor 
decreases  below  K
Ic
.    From  Equation  2.38,  fracture  toughness  is  related  to  the  applied 
stress and crack length, and the function F, which is a dimensionless function dependent 
on the crack and specimen geometry, and the loading configuration.  When working with 
applied  loads  and  planar  geometries,  as  is  this  case  in  rock  testing,  the  critical  stress 
intensity  factor  is  determined  from  the  critical  applied  load,  P
Q
,  and  the  corresponding 
crack length at failure, a
c
.  The measured fracture toughness value is termed the apparent 
fracture  toughness,  K
Q
.    After  K
Q
  has  been  measured  it  is  checked  against  the  specimen 
size requirements of Equation 2.40 in order to verify its validity. 
 
  There are various methods for the determination of P
Q
 and a
c
, and they are dependent 
upon the test method and the test results (for a comprehensive review see Whittaker et al., 
1992).  For example, with notched specimens the critical crack length at failure is equal 
to  the  original  notch  length  and  the  critical  applied  load  is  equal  to  the  maximum  load 
applied.    Frequently  some  type  of  load-displacement  curve  is  used  to  determine  P
Q
  and 
a
c
.    These  load-displacement  curves  are  usually  in  the  form  of  applied  load,  P,  versus 
either  Load  Point  Displacement  (LPD)  or,  if  possible,  P  versus  Crack  Mouth  Opening 
Displacement  (CMOD).    Again  various  methods  can  be  employed  using  P  vs. 
LPD/CMOD curves (i.e., 5% Secant Approach, Compliance Calibration).  An additional 
function  of  P  vs.  LPD/CMOD  curves  is  that  they  can  be  used  to  observe  and  even 
account  for  non-linear  behavior  at  the  crack  tip,  allowing  for  the  extension  of  LEFM 
analysis to rocks that exhibit some non-elasticity. 
 
  57
  After a series of K
Q
 measurements have been validated using Equation 2.40, K
Ic
 can 
be  determined  by  extrapolating  the  K
Q
  versus  crack  length  curve  until  K
Q
  becomes  a 
constant.    Since  K
Ic
  is  independent  of  specimen  dimensions  it  is  a  limiting  value  of  K
Q
 
(which has been shown to increase with crack length). 
 
Test methods 
As  mentioned  previously  there  are  two  test  methods  suggested  by  the  International 
Society for Rock Mechanics for the determination of rock fracture toughness.  They are 
the Chevron Edge Notched Round Bar in Bending (CB) and the Chevron Notched Short 
Rod (SR).  The motivation for developing the suggested methods was to provide testing 
methods  that  consistently  yield  accurate  and  precise  K
Ic
  values  (ISRM,  1988).    Each 
method is also core based, the only viable specimen alternative as rock is often available 
in the form of core pieces (Ouchterlony and Sun, 1983). 
 
  Another commonly employed test method is the Semi-Circular Bend (SCB) test.  The 
SCB  can  be  prepared  from  rock  cores  and  is  especially  adaptable  to  small,  compact 
samples  that  require  duplicate  samples  to  test  parameters  that  may  affect  K
Ic
  such  as 
loading rate, specimen thickness, and crack  length (Karfakis et al., 1986; Chong, 1980).  
Despite  the  suggested  standard  methods  of  the  ISRM  and  the  popularity  of  the  SCB, 
various  other  methods,  each  with  its  own  advantages,  are  still  used  to  determine  rock 
fracture  toughness.    Whittaker  (et  al.,  1992)  provides  an  exhaustive  review  of  these 
alternative methods. 
 
Chevron edge notched round bar in bending 
Figure  2.27  shows  the  specimen  configuration  and  test  set-up  for  the  CB  test.    The  CB 
test employs a bend specimen test with a chevron-notch cut perpendicular to the core axis 
(Ouchterlony,  1980).    The  specimen  rests  on  roller  supports  and  a  compressive  load  is 
applied  causing  crack  growth  and  transverse  splitting  of  the  specimen.    For  a  complete 
review  and  background  of  the  test  see  Sun  and  Ouchterlony  (1986),  ISRM  (1988), 
Ouchterlony (1989), Ouchterlony (1989a), and Whittaker (et al., 1992). 
 
  The  core  based  specimen,  chevron-notch,  stable  crack  growth,  ability  to  account  for 
non-linearity, and multiple testing levels (Level I and II) are the main advantages of the 
CB test.  However, as can be seen from Figure 2.27 the specimen geometry and loading 
configuration are not simple.  And despite the advantage of the chevron-notch it is still a 
rather  difficult  pre-crack  to  machine.    Furthermore,  the  advantages  of  the  CB  test  seem 
only  to  be  significant  in  terms  of  Level  II  testing.    If  Level  I  testing  is  sufficient  then 
other  more  straightforward  techniques  are  available  for  the  rapid  estimation  of  fracture 
toughness.  
 
Chevron Notched Short Rod 
Figure 2.28 shows the configuration and test set-up for the SR test.  The SR test specimen 
has a chevron-notch that is cut parallel to the core axis.  A tensile load is applied at the 
notch  mouth  causing  crack  growth  and  lengthwise  splitting  of  the  specimen.    For  a 
  58
Figure 2.27 The CB specimen and testing configuration 
(After ISRM, 1980 and Whittaker et al., 1992)
complete  background  review  and  detailed  analysis  see  Barker  (1977),  Bubsey  (et  al., 
1982),  Sun  and  Ouchterlony  (1986),  ISRM  (1988),  Ouchterlony  (1989),  Ouchterlony 
(1989a), and Whittaker (et al., 1992).  
 
  The advantages of the SR test are the same as those for the CB test.  The SR method 
also uses a smaller sized specimen and according to the ISRM, two SR specimens can be 
obtained from one fractured CB test (since the length to diameter requirement of the CB 
test  is  4:1  and  for  the  SR  it  is  1.45:1).    This  saves  material  and  allows  for  the 
measurement  of  fracture  toughness  using  two  crack  orientations.    Again,  the 
disadvantages of the SR are that the specimen, notably the chevron-notch, is difficult to 
reproduce consistently and that for Level I type testing it is too complex.  
 
Semi-Circular Bend 
Figure  2.29  depicts  the  specimen  geometry  and  loading  configuration  of  the  SCB  test.  
The  SCB  test  was  developed  and  proposed  by  Chong  and  Kuruppu  (1984)  in  order  to 
provide a test method that was simple to fabricate and load.  The pre-crack can be a saw-
cut notch that is fatigued loaded in order to produce a natural crack, a chevron-notch, or a 
very thin saw-cut notch.  The kinematics of the test are similar to that of the CB test since 
a  vertical  compressive  load  produces  three-point  bending  and  transverse  splitting  of  the 
  59
Figure 2.28 The SR specimen and test configuration 
(After ISRM, 1988 and Whittaker, et al., 1992) 
Figure 2.29 The SCB specimen and test configuration 
(After Karfakis et al., 1986) 
specimen  (Akram,  1991).    A  detailed  description  of  the  test  is  given  by  Chong  and 
Kuruppu (1984), Karfakis et al., 1986, Chong (et al., 1987), and Whittaker (et al., 1982).   
 
Testing of the SCB specimen is simple and does not require sophisticated equipment, 
and  a  large  number  of  samples  can  be  tested  rapidly  and  accurately.    But  the  test  is 
  60
difficult to perform on standard NX core specimens (approximately 50 mm in diameter) 
and larger diameter specimens are usually required in order to obtain stable loading and 
crack growth. 
 
2.5.4 APPLICATIONS OF ROCK FRACTURE TOUGHNESS 
 
Because  of  the  complicated  nature  of  rocks  and  rock  formations,  the  application  of 
fracture  mechanics  to  rock  fragmentation  has  not  been  straightforward.    But  recently 
(within  the  last  25  years)  there  has  been  an  effort  to  employ  aspects  of  fracture 
mechanics,  particularly  fracture  toughness,  to  various  types  of  rock  breakage  analyses.  
This  work  has  been  able  to  correlate  fracture  toughness,  the  intrinsic  material  property 
expressing a materials resistance to crack propagation, with various rock fragmentation 
processes such as blasting, tunnel boring and rock cutting. 
 
  The process of rock cutting involves forcing a cutting tool into rock in order to break 
out  fragments  from  the  surface.    Since  cutting  is  achieved  by  a  fracture  process  it  is 
logical to expect that differences in cuttability of various rocks is related to the variation 
of fracture material properties among the rocks (Nelson and Fong, 1986).  Deliac (1986) 
analyzed  the  chip  formation  due  to  drag  picks  and  found  that  for  sharp,  rigid  picks 
operating in brittle rocks the cutting force can be expressed as a function of the fracture 
toughness of the rock and the cutting depth.  Guo (1990) related the fracture toughness to 
the  penetration  rate  of  a  diamond-coring  machine  and  found  that  rocks  with  higher 
fracture toughness are harder to penetrate resulting in lower penetration rates.  Ingraffea 
(ey  al.,  1982)  has  proposed  using  the  fracture  toughness  of  different  rocks  encountered 
during  three  tunnel  boring  projects  to  predict  the  performance  of  the  TBMs  (Tunnel 
Boring  Machines).    Preliminary  testing  showed  that  fracture  toughness,  in  general, 
showed less variability than other material properties (uniaxial strength, tensile strength, 
and  point  load  strength)  and  that  fracture  toughness  appears  to  be  an  ideal  measure  to 
relate to TBM performance.  It has been shown that the critical energy release rate, G, is 
related  linearly  to  the  penetration  rates  of  TBMs.    Since  G  is  related  to  fracture 
toughness by Poissons ratio and the modulus of elasticity, it follows that the penetration 
rate could also be predicted by fracture toughness (Clark, 1987). 
 
  The  connection  of  fracture  toughness  to  some  common  fragmentation  processes 
indicates that fracture toughness could also be correlated to crushing processes.  Bearman 
(et  al.,  1991)  correlated  various  rock  strength  parameters  to  power  consumption  of  a 
laboratory  cone  crusher  with  a  statistical  significance  of  99.9%  and  fracture  toughness 
was among those parameters.  The same study also showed that fracture toughness could 
be related to product size with a significance of 95% and that K
Ic
 also could be correlated 
with  energy-size  relationships  derived  from  single  particle  breakage  tests.    The  same 
correlations  and  level  of  confidence  should  hold  true  for  laboratory  scale  primary 
crushing equipment (jaw and gyratory crushers) and optimistically for actual, large-scale 
industrial crushers employed at processing plants. 
 
  61
2.5.5 CONCLUSIONS 
 
Irwins  modifications  to  Griffiths  theory,  and  his  demonstration  that  the  strain  energy 
release rate is related to the stress intensity factor, led to the development of the field of 
fracture mechanics.  Fracture mechanics provides quantitative methods for characterizing 
the behavior of an intact material as it fractures due to crack growth.  Although fracture 
mechanics  was  developed  for  metallic  or  man-made  materials  its  extension  into  rock 
mechanics  is  natural  due  to  the  presence  of  inherent  flaws  and  discontinuities  within 
geologic  materials.    These  flaws  control  the  fracture  of  rocks,  and  establishing  a 
relationship between flaw geometry and fracture strength is the most fundamental aspect 
of fracture mechanics.  
 
  Irwins work provided the basis for Linear Elastic Fracture Mechanics.  In LEFM the 
crack tip stresses and displacements can be characterized by the stress intensity factor K 
as long as plastic, or non-linear, deformation ahead of the crack tip is small.  The stress 
intensity  factor  characterizes  the  magnitude  of  the  stresses  near  a  crack  tip  in  a  linear-
elastic,  homogeneous,  and  isotropic  material.    LEFM  provides  a  universal  approach  for 
determining  a  materials  resistance  to  fracture  since  crack  propagation  occurs  when  K 
reaches a critical value.   
 
There  are  three  general  modes  of  crack  tip  displacement  that  are  used  to  describe 
most cases of local crack tip stress and deformation fields.  Mode I is the opening mode 
and  it  is  the  dominant  mode  in  rock  fragmentation.    Using  theory  of  elasticity  a  general 
form of the Mode I stress intensity factor has been determined and shown to be a function 
of  the  applied  stress,  the  crack  length,  and  a  dimensionless  constant,  which  itself  is  a 
function of loading and specimen geometry.  In order to determine fracture toughness an 
explicit  function  that  describes  the  stress  intensity  near  a  crack  tip  for  a  given  crack 
geometry and loading configuration needs to be known.  Once this function is defined the 
fracture  toughness  can  be  determined  experimentally  based  on  the  loading  condition, 
crack geometry, and specimen configuration. 
 
  Fracture  toughness  testing  of  rock  specimens  is  influenced  by  the  development  of  a 
fracture process zone (FPZ) ahead of the crack tip.  The FPZ is formed by the initiation 
and propagation of micro-cracks near the crack tip.  The FPZ is similar to the plastic zone 
that  forms  ahead  of  a  crack  tip  in  metals,  and  likewise  is  used  to  define  specimen 
dimension  requirements,  ensuring  that  the  amount  of  non-elastic  behavior  ahead  of  the 
crack tip is small enough so that LEFM remains applicable.  Fracture toughness testing of 
rock can also be influenced by the microstructure/grain size of the rock, the type of crack 
used, the testing level, and the loading rate.  
 
There are currently no standardized test methods for the determination of Mode I rock 
fracture toughness.  The main requirements of any test method are that it be core based, 
require  simple  specimen  preparation,  be  easy  to  load,  have  a  straightforward  fracture 
toughness  calculation,  and  yield  representative  and  reproducible  values.    The  ISRM  has 
  62
suggested  two  methods  for  standardization,  the  Chevron  Edge  Notched  Round  Bar  in 
Bending (CB) and the Chevron Notched Short Rod (SR).  The Semi-Circular Bend (SCB) 
test is also commonly employed.  Each of these tests has its advantages yet none are used 
exclusively.    The  CB  and  SR  specimens  require  a  chevron-notch,  which  is  difficult  to 
reproduce  consistently,  and  the  SCB  test  is  difficult  to  apply  to  standard  NX  size  rock 
cores.  
 
  Rock  fracture  toughness  has  found  wide  application  in  processes  of  rock 
fragmentation.    Researchers  have  been  able  to  correlate  fracture  toughness  with  the 
penetration rates of tunnel boring machines and rock cutters.  In the field of comminution 
fracture  toughness  has  been  related  to  the  power  consumption  and  product  size 
distribution  of  a  cone  crusher.    Based  on  these  results  it  is  expected  that  fracture 
toughness  can  be  related  to  the  power  consumption,  product  size,  and  capacity  of  jaw 
crushers.    When  individual  rock  particles  are  subjected  to  the  applied  forces  of  size 
reduction  the  amount  of  energy  input  and  the  amount  of  size  reduction  achieved  are 
related  to  the  type  of  loading  and  the  crack  pattern  in  the  material.    Fracture  toughness 
measures a materials resistance to fracture based on the applied load and flaw geometry 
within the rock, thus the same characteristics that control the operating parameters of jaw 
crushers also control an intrinsic material property.  However, in order to investigate the 
use  of  fracture  toughness  in  crusher  applications  a  new  test  method  for  the  rapid 
estimation and comparison of fracture toughness values needs to be developed.   
  63
CHAPTER 3. EXPERIMENTAL PROGRAM 
 
 
In  order  for  any  size  reduction  process  to  be  fully  characterized,  material  properties,  or 
physical expressions describing the nature of the material broken, should be related to the 
pattern  of  breakage  and  resultant  fragment  size  distribution  created  by  crushing.    It  has 
been  suggested  that  a  materials  fracture  properties  should  be  characterized  in  terms  of 
particle strength, specific breakage energy, and the breakage fragment size distribution of 
the material.  It is proposed that the particle strength should be replaced by another, more 
descriptive measure of a materials ability to withstand fracture.  Fracture toughness fully 
characterizes  a  rocks  resistance  to  fracture  and  is  dependent  on  the  fracture  strength  of 
the rock as well as the presence of flaws within the rock.  Therefore, it is proposed that 
the potential of fracture toughness to function as a predictive means for rock crushability, 
and  to  develop  fracture  toughness  based  models  for  the  selection  and  optimization  of 
primary  jaw  crushers  be  examined.    The  aim  of  the  models  is  to  predict  the  power 
consumption,  product  size,  and  capacities  of  jaw  crushers.    Using  the  technique  of 
single  particle  breakage,  discussed  in  section  2.4,  the  specific  comminution  energy  and 
fragment  size  distributions  of  various  aggregates  will  be  determined.    The  associated 
fracture toughness of each rock type will be measured using a newly developed fracture 
toughness  test.    Finally,  a  lab-scale  jaw  crusher  will  be  used  to  test  the  strengths  of  the 
laboratory-based  models,  and,  if  possible,  to  determine  if  scale-up  regression  analysis 
between the laboratory results and actual operating conditions is necessary. 
 
 
3.1 ROCK SPECIMENS 
 
Luck Stone Corporation, the 12
th
 largest producer of crushed stone products in the United 
States,  with  operations  in  Maryland,  Virginia,  and  North  Carolina,  has  provided  rocks 
from  five  of  their  aggregate  quarries:  Boscobel,  Charlottesville  (Shadwell),  Culpeper, 
Leesburg and Spotsylvania (Figure 3.1).  Additional rock was also provided from a sixth 
site not currently under production, identified as the Thornburg property (located near the 
Spotsylvania operation).  Rock from the Culpeper site classifies into two groups, by color 
(red and gray), and although the properties of each are expected to be the same, they are 
to be tested independently in this study. 
 
Each  of  the  rock  types  originates  in  the  same  physiographic  province.    Virginia 
contains  five  physiographic  provinces:  the  Coastal  Plain,  Piedmont,  Blue  Ridge,  Valley 
and  Ridge,  and  Appalachian  Plateaus  (east  to  west  in  Figure  3.2).    The  six  Luck  Stone 
rocks are found in the Piedmont.  The Piedmont extends eastward from the Blue Ridge to 
the  Fall  Line,  where  unconsolidated  sediments  of  the  Coastal  Plain  cover  Paleozoic-age 
and  older  igneous  and  metamorphic  rocks.    The  Piedmont  is  characterized  by  deeply 
weathered, poorly exposed bedrock, and a high degree of geological complexity (DMME, 
2003).    Crystalline  rocks  are  sprinkled  with  a  few  areas  of  much  younger, 
unmetamorphosed,  sedimentary  and  volcanic  rocks  found  in  rift  basins,  with  the  largest 
  64
being the Culpeper Basin.  A more specific geologic description of each rock type to be 
tested follows and has been provided by Luck Stone (Luck Stone, 2003).   
Figure 3.1 Location of rock quarries that provided specimens 
(From Luck Stone, 2003)
 
  65
Figure 3.2 Physiographic provinces of Virginia (Piedmont is green, 2
nd
 from right)
(From VA DMME, 2003) 
Boscobel Granite (BG) 
The  Boscobel  operation  quarries  Petersburg  Granite,  an  igneous  formation  about  330 
million  years  old.    The  rock  was  formed  from  a  molten  mass  and  has  a  pink,  multi-
colored  appearance.    Quart,  orthoclase,  plagioclase,  muscovite,  and  small  amounts  of 
dark colored minerals combine to give the granite its multi-colored appearance. 
 
Shadwell Metabasalt (SMB) 
The  Shadwell  operation  (just  outside  of  Charlottesville)  mines  Catoctin  Greenstone, 
formed  by  lava  flows  over  500  million  years  ago.    The  greenstone  originated  as  basalt 
and  contains  plagioclase  and  pyroxene  minerals.    Epidote  and  chlorite  infused  the 
formation during mountain building giving the green color. 
 
Culpeper Siltstone (CGS and CRS) 
The  Culpeper  quarry  produces  sedimentary  siltstone,  sandstone,  and  shale.    Siltstone  is 
the main rock, but the formation does grade down in particle size to shale and up in size 
to fine-grained sandstone.  Sediments eroded from western uplands were deposited in the 
low-lying  Culpeper  Basin,  which  was  formed  as  a  result  of  continental  land  masses 
pulling apart.  Deposition of the sediments took place 200 million years ago, with lower 
sediments gradually hardening under the effects of weight and pressure.  The red colored 
rock  (found  at  greater  depth)  results  from  the  hematite-cementing  agents  that  bound  the 
sediment grains into rock.  The upper, gray colored rock occured as a result of a lack of 
oxygen during deposition that prevented the cementing agent from becoming red in color. 
 
Leesburg Diabase (LD) 
The Leesburg quarry produces diabase, sometimes referred to as traprock.  The diabase is 
a dense, igneous rock with greenish-black to bluish-black color.  Formed by hot magma 
over 200 million years ago, the rock contains mainly pyroxene and plagioclase. 
 
Spotsylvania Granite (SG) 
The  Spotsylvania  operation  quarries  a  complex  blend  of  igneous  and  metamorphic  rock 
types  (Po  River  Metamorphic  Suite).    The  rock  types  at  Spotsylvania  are  granite-gneiss 
and biotite-gneiss, and are about 400 million years old.  The granite-gneiss is a white to 
gray color and the biotite-gneiss is dark gray to black.          
 
Thornburg Granite (TG) 
The  rock  formation  at  the  Thornburg  property  is  similar  to  that  of  the  Spotsylvania 
granite.  Formed by a later igneous intrusion, the rock is different in its appearance than 
the Spotsylvania granite.  It is a reddish granite with large mineral crystals that contrasts 
with the surrounding rock.  Figure 3.3 depicts each of the seven rocks to be tested. 
 
3.1.1 ROCK PROPERTIES 
 
A  series  of  standard  laboratory  tests  have  been  conducted  on  the  rock  types  in  order  to 
determine  their  mechanical  properties  and  to  provide  a  basis  for  comparison  among 
typical  rock  properties  and  fracture  toughness  as  they  relate  to  the  selection  of  jaw 
crushers.    The  uniaxial  compressive  strength,  indirect  tensile  strength,  and  elastic 
modulus  of  each  rock  was  determined  pursuant  to  ASTM  standard  test  method 
  66
Figure 3.3 Seven quarry rocks to be tested
procedures  (ASTM  D2938-95:  Standard  Test  Method  for  Unconfined  Compressive 
Strength  of  Intact  Rock  Core  Specimens;  ASTM  D3967-95:  Standard  Test  Method  for 
Splitting Tensile Strength of Intact Rock Core Specimens).  The testing was conducted in 
the  Department  of  Mining  and  Minerals  Engineering  Rock  Mechanics  Laboratory  at 
Virginia Tech.  Table 3.1 summarizes the test results.  
 
Table 3.1 Mechanical properties of tested rocks 
Compressive Strength  Tensile Strength  Elastic Modulus 
Rock Type 
MPa  MPa  GPa 
Shadwell Metabasalt  142.0  18.72  29.79 
Boscobel Granite  59.2  7.76  16.76 
Culpeper Grey Siltstone  185.3  21.94  25.93 
Culpeper Red Siltstone   155.0  22.08  23.65 
Leesburg Diabase  225.7  17.61  29.51 
Thornburg Granite  142.6  13.36  27.03 
Spotsylvania Granite  114.3  12.44  25.38 
 
 
3.2 FRACTURE TOUGHNESS TESTING 
 
Although  the  International  Society  of  Rock  Mechanics  has  suggested  that  the  Chevron 
Notched  Short  Rod  and  Chevron  Notch  Round  Bar  in  Bending  be  adopted  as 
recommended  fracture  toughness  tests  (see  section  2.5.3),  the  use  of  these  tests  for  rock 
characterization  and  indexing  purposes  is  not  widespread.    The  complexity  of  these 
existing  tests  is  an  immediate  problem  when  considering  the  application  of  fracture 
  67
toughness  to  gauge  size  reduction  processes.    Thus,  another  method  for  fracture 
toughness  testing  of  rocks  is  necessary,  one  that  will  give  a  representative  toughness 
value  (in  accordance  with  Level  I  testing)  and  yet  be  simple,  core  based,  with  easily 
reproducible  crack  geometry,  and  requiring  only  straightforward  data  collection  and 
evaluation techniques. 
 
  It is proposed that a wedge test utilizing an edge notched disc (END) be employed in 
order  to  acquire  fracture  toughness  values  suitable  for  rapid  indexing  and  comparative 
purposes.    The  test  will  enable  a  large  number  of  rock  types  to  be  tested  so  that 
relationships  among  breakage  energy,  fragment  size  distribution,  and  fracture  toughness 
can  be  investigated  in  order  to  develop  a  crushing  index  capable  of  predicting  and 
evaluating primary crushing equipment performance.  The development and experimental 
verification of the END wedge splitting test, using trial rocks, is covered in the following 
section.   
 
3.2.1 DEVELOPMENT OF THE EDGE NOTCHED DISC WEDGE SPLITTING TEST 
 
Stress intensity factors for edge notched disk 
Recalling section 2.5, a materials resistance to crack propagation can be quantified using 
fracture  toughness,  K
Ic
,  the  critical  stress  intensity  factor.    When  the  stress  intensity 
factor, K
I
, in the region of a crack tip exceeds K
Ic
, fracture initiates and propagates until 
the  stress  intensity  factor  decreases  below  K
Ic
.    Equation  2.38  showed  the  relationship 
between  fracture  toughness,  the  applied  stress,  crack  length,  and  F,  a  dimensionless 
constant  dependent  on  the  geometric  configuration.    When  working  with  applied  loads 
and  planar  geometries,  as  is  this  case  in  rock  testing,  the  stress  intensity  factor  can  be 
expressed according to: 
 
I p
P
K F
t w
=   [3.1] 
where, P is the applied load 
    t is the specimen thickness 
    w is the uncracked ligament length (or the maximum possible crack length) 
    F
p
 is a new dimensionless geometry factor, defined as: 
  (   ) , /
p p
F f geometry a w =   [3.2] 
 
For the testing and calculation of rock fracture toughness using an edge notched disk, 
the  stress  intensity  factor  and  F
p
  must  be  determined.    In  reference  to  the  specimen 
geometry and loading configuration given in  Figure 3.4, Isida et al. (1979) and Gregory 
(1979)  have  independently  characterized  the  stress-strain  region  enclosing  the  crack  tip 
using stress field analysis for the two dimensional case, assuming linear-elastic behavior 
and small-scale yielding (Murakami, 1987).  For an edge cracked circular plate subject to 
concentrated  forces  acting  at  symmetrical  points,  Isida  (et  al.,  1979)  employed  the 
boundary collocation method and found the stress intensity factor, F to be: 
  68
D
a
P P
Figure 3.4 Edge notched disk specimen and loading configuration 
 
I
Isida
K D
F
P a 
=   [3.3] 
Isida (et al., 1979) give values of F for a/D ratios ranging from 0.1 to 0.6 (Table 3.2). 
 
Table 3.2 Values of a/D and corresponding value of F from Isida (et al., 1979) 
a/D  0.1  0.2  0.3  0.4  0.5  0.6 
F
 
11.488  7.721  7.051  7.451  8.636  10.990 
 
  Gregorys  F  equation  was  derived  using  stress  functions  and  it  is  a  closed  form 
solution.  It is applicable to an edge cracked circular disk subjected to pin loading at the 
crack mouth.  F can be written as: 
 
2
2
I
Gregory
K
F
P D
=
a
  [3.4] 
Gregory  gave  the  function  that  defines  F  so  the  stress  intensity  factor  for  any  a/D  ratio 
can be calculated directly.  The expression for F is: 
 
(   )   (   )
3
2 2
1
0.966528 0.355715
Gregory
a
F
D a D a
=   +
 
1
  [3.5] 
 
A plot of F
p
 (as defined in Equations 3.1 and 3.2) versus a/D was developed in order 
to  compare  the  two  solutions  (Figure  3.5).    F
p
  in  terms  of  Isida  and  Gregorys  F  was 
created  by  normalizing  Equations  3.3  and  3.4  against  Equation  3.1  resulting  in  the 
following expressions: 
 
P Isida Isida
a
F
D
  = F   [3.6] 
 
2
2
P Gregory Gregory
D
F
a
  = F   [3.7] 
 
  69
Figure  3.5  shows  that  the  two  solutions  are  almost  identical  for  a/D  ratios  varying 
from  0.1  to  0.6,  and  either  can  be  used  to  determine  the  stress  intensity  factor  for  the 
END.  For application in rock fracture toughness testing it was determined that Gregorys 
solution provides the easiest method for determining K
I
 since a direct expression for F is 
given, and K
I
 can be determined precisely for any a/D ratio. 
  
Load application 
The  previous  stress  analysis  requires  the  edge  notched  disk  to  be  loaded  at  the  crack 
mouth by equal and opposing forces.  This loading configuration can be achieved rather 
simply  by  using  a  wedge.    The  splitting  force  of  the  wedge,  P
sp
,  is  the  horizontal 
component of the force acting on the crack mouth edge and can be determined from force 
equilibrium analysis (Figure 3.6 and Appendix V).  Equation 3.8 shows the splitting force 
in terms of the applied vertical force, P
v
, acting on the wedge, the wedge angle , and the 
coefficient of interface friction . 
 
(   )
  (   )
(   )
1 tan
2
2tan 1 cot
2 2
v
sp
P
P
  
=   
+
  [3.8] 
 
The  wedge  provides  a  mechanical  advantage,  an  advantage  that  increases  as  the 
wedge  angle  decreases.    For  wedging  of  a  crack  in  rock  the  amount  of  friction  between 
Figure 3.5 Comparison of independent solutions of F
p
 vs. a/D for an END
0.00 0.06 0.12 0.18 0.24 0.30 0.36 0.42 0.48 0.54 0.60
5
6
7
8
9
10
11
12
13
14
15
F
p
a/D
 Gregory 
 Isida et al
*(for D = 2)
  70
Figure 3.6 Forces acting at the crack mouth 
P
sp
 
P
P
v
Notch
the 
ble 3.3 Values of  and  from tilt test on hardened steel 
  Large grained granite  Fine grained granite  Dolomitic limestone  Sioux quartzite 
wedge material and the rock can be substantial and frictional losses may occur.  The 
coefficient of friction needs to be quantified for each rock type tested in order to account 
for frictional losses.  A tilt test can be used to determine , where the angle of sliding is  
and tan  is equal to .  The coefficients of friction between four rock types and hardened 
steel are shown in Table 3.3 (rock types and wedge material used during END wedge test 
development).  
 
Ta
  .8  18.5  18.9  19.7  22
  0.335  0.342  0.358  0.420 
 
Test Specimen 
he  test  specimen  can  be  prepared  from  standard  rock  core,  resulting  in  the  geometry 
 3.4.  Specimens are cut using a diamond wheel saw and the sides of the 
  the  following 
ection,  but  guidelines  for  rock  fracture  testing  have  already  been  discussed  (section 
2.5
T
shown in Figure
disk ground until parallel (or to within  0.125 mm) to each other.  The edge notch is cut 
using a diamond wheel saw (with a thickness of 1.5 mm) and requires a fixture that holds 
the disk and ensures that the notch is cut along the centerline of the disk.  
 
Minimum  specimen  dimensions  for  the  END  are  investigated  in
s
.3).    In  general,  values  of  rock  fracture  toughness  do  not  show  the  same  type  of 
dependency on specimen size, namely thickness, as seen in testing of metallic materials.  
But the size cannot be too small and the thickness must not be smaller than the width of 
the  fracture  process  zone.    The  requirements  given  by  Equation  2.40  are  used  for  rock 
fracture  testing  and  these  are  checked  after  an  apparent  fracture  toughness  value  has 
  71
been  measured.    Since  the  specimens  will  be  prepared  from  NX  rock  core  the  diameter 
will  typically  be  around  50  mm.    The  proposed  thickness  is  approximately  25.4  mm.  
Various  notch  lengths  were  proposed  initially  in  order  to  investigate  any  relationship 
between fracture toughness results and specimen size. 
 
Wedging device 
The wedging device used to apply the splitting force along the crack mouth edge consists 
.  The wedge angle is 11.  The wedge, with approximate dimensions, is 
 
Test set-up, loading, and measurement 
he  END  specimen  is  placed  on  the  wedge  (Figure  3.8).    The  vertical  force  is  applied 
under displacement control by an MTS 810 closed loop servo-hydraulic materials testing 
of hardened steel
shown in Figure 3.7. 
Figure 3.7 Wedge configuration
13 mm
76 mm
76 mm
30 mm
6 mm
T
  72
Figure 3.8 Test set-up for END wedge test
(After Donovan and Karfakis, 2003)
system  affixed  with  an  8896  N  load  cell.    Load  and  load-line  displacement  data  are 
recorded digitally by a up). 
 
in terms of stress 
intensity cannot be determined until after testing.  Different rates were tested, by varying 
the 
be  recorded  in  order  to  calculate  K
Ic
.    However,  the  load-displacement  data 
ould  be  recorded  in  order  to  verify  that  the  maximum  load  is  the  critical  applied  load 
that
 PC  ack- and also with an X-Y recorder (for b
Various  loading  rates,  usually  in  terms  of  stress  intensity  per  second,  have  been 
prescribed for rock fracture testing (see section 2.5.3).  The loading rate 
load-line  displacement,  in  order  to  investigate  any  dependency  of  the  results  on 
loading  rate.    The  loading  rate  for  the  END  wedge  test  is  discussed  in  the  following 
section. 
  
In  accordance  with  Level  I  rock  fracture  toughness  testing  only  the  maximum  load 
needs  to 
sh
 results in crack propagation.  For the END wedge test this involves an approximation 
of  the  crack  mouth  opening  displacement  (CMOD).  There  are  no  plans  to  measure  the 
CMOD  directly  but  it  can  be  calculated  from  the  wedge  angle  and  the  load-line 
displacement.  The 5% secant approach for determining the critical load has been shown 
to be satisfactory for rock testing, although smaller specimens may give lower K
Ic
 values.  
This method, along with direct observation of the load-displacement curve (i.e., looking 
for non-linear behavior), was used to verify that the maximum load was the critical load 
when testing trial rocks using the END wedge test. 
 
  73
Calculation of K
Ic 
fracture toughness using the END wedge test is straightforward.  The  The calculation of 
critical  load,  if  not  the  maximum  load,  is  determined  from  the  load  and  the  calculated 
CMOD  curve  (Figure  3.9).    The  apparent  fracture  toughness  is  calculated  using  the 
following  equation,  by  substituting  for  P  in  Gregorys  stress  analysis  of  the  END 
specimen  (Equations  3.4  and  3.5)  with  the  splitting  force,  P
sp
,  applied  by  the  wedge 
(Equation 3.8): 
(   )
  (   )
(   )   (   )   (   )
3 1
2 2
1 tan
1
2
2
2
2tan 1 cot 0.966528 0.355715
2 2
v
q
P D a
K
a
D a D a
|   |
|   | 
   |
   |
=      +
   |
   |
   | +    
\   .
\   .
[3.9] 
The  apparent  fracture  toughness  value  is  considered  valid  (i.e.,  equal  to  K
Ic
)  after 
 
.2.2 EXPERIMENTAL VERIFICATION OF THE END WEDGE SPLITTING TEST 
our  rock  types  were  tested  in  order  to  test  the  validity  of  the  END  specimen  and  test 
checking that the specimen size requirements are met.   
800
Figure 3.9 Example vertical load vs. calculated CMOD curve 
0.000 0.025 0.050 0.075 0.100 0.125 0.150 0.175 0.200 0.225 0.250
0
80
160
240
320
400
480
560
640
720
Dolomitic Limestone
Sample 4
P
v
(N)
Calculated Crack Mouth Opening Dispacement (mm)
3
 
F
configuration: fine-grained granite, large-grained granite, dolomitic limestone, and Sioux 
quartzite.    These  rocks  were  chosen  because  samples  of  the  fine-grained  granite, 
limestone,  and  the  quartzite  each  had  previously  been  tested  using  a  common  fracture 
  74
toughness  test,  the  Semi-Circular  Bend  test  (see  section  2.5.3),  providing  a  basis  for 
comparison  (Akram,  1991).    The  large-grained  granite  was  included  in  order  to 
investigate the influence of notch length and loading rate on fracture toughness. 
 
Fifty-four samples were prepared in accordance with the previous section.  A 15 cm 
diam
ing  of  the  specimens  was  done  using  an  MTS  810  closed  loop  servo-hydraulic 
mat
eter,  diamond  bladed,  radial  arm  saw  was  used  to  cut  25  mm  thick  disks  from  50 
mm diameter rock cores.  The disk sides were ground flat and parallel to each other using 
a  diamond  surfaced  grinding  wheel.    A  1.5  mm  thick  diamond  blade  was  used  to  cut  a 
straight  through  edge  notch  along  the  centerline  of  the  disc.    The  notch  lengths  varied 
from  10.57  mm  to  28.35  mm.    Cutting,  grinding,  and  notching  of  the  samples  took  less 
than 3 hours, resulting in a preparation time of approximately 3 minutes per specimen. 
   
Test
erials  testing  system  affixed  with  an  8.896  kN  load  cell.    The  tests  were  performed 
under  load  line  displacement  control.    The  load-line  displacement  rate  was  0.025  mm/s 
for  the  limestone  and  quartzite,  either  0.003  mm/s  or  0.001  mm/s  for  the  fine-grained 
granite,  and  varied  among  the  previous  three  rates  for  the  large-grained  granite.    After 
testing  the  loading  rates  were  determined  in  terms  of  stress  intensity.    The  maximum 
stress intensity-loading rate occurred with the limestone and was 0.0986 / MPa m s .  The 
minimum  occurred  with  the  large-grained  granite  and  was  0.0011 / MPa ssible 
effects of loading rate on fracture toughness are discussed later. 
 
m s .    Po
The  load-displacement  data  was  collected  and  converted  into  splitting  force  (P
sp
)-
calc
fter testing, all specimen dimensions were tested against Equation 2.40 to verify that 
dim
racture toughness results 
  the  END  wedge  splitting  and  those  of  the  Semi-Circular 
ulated  CMOD  data  in  order  to  verify  that  the  maximum  load  was  also  the  critical 
load.    This  required  the  measurement  of  the  frictional  coefficient  between  the  rock  and 
the hardened steel of the wedge.  For all 54 samples the critical applied load resulting in 
crack propagation was also the maximum load.   
 
A
ension requirements were met (tensile testing of the same rock types was performed 
in order to determine 
t
).  Of the 54 samples tested, 1 fine-grained granite sample and 3 
limestone samples did not meet the minimum requirement for crack length (one of the 3 
limestone  samples  also  did  not  meet  the  thickness  specification).    These  samples  were 
disregarded. 
 
F
Table  3.4  shows  the  results  of
Bend test (Akram, 1991).  Statistical analysis of the results for each individual rock type 
indicates that the mean values of fracture toughness from both tests are not significantly 
different (at a confidence level of 99%).  From the initial analysis it seems reasonable to 
expect  the  END  wedge  split  test  to  yield  representative  fracture  toughness  values, 
particularly in terms of indexing and for comparative purposes. 
 
 
  75
Table 3.4 Fracture toughness values from the END test and SCB test 
END  SCB 
Rock Type 
# of Specimens
Ic
K    MPa m   # of Specimens K
Ic
   MPa m  
13  0.914 3   6  0.884 
Large grained granite  NA  NA  27  0.621  0.101 
Dolomitic limestone  9  1.398  0.186  6  1.33 080  1  0.
Sioux quartzite  4  1.241  0.031  6  1.244  0.071 
Fine grained granite    0.09
   0.022 
 
Based on the test results, there is some dependency on grain size.  The larger grained 
roc
rack length effects 
cture toughness values versus a/D for the two granite rocks tested.  
ks had the lowest fracture toughness values, and this follows the general observation 
that rocks with smaller grains tend to have higher values of fracture toughness (Whittaker 
et  al.,  1992).    Generally,  for  rock  testing,  the  ratio  of  smallest  specimen  dimension  to 
grain size should be at least greater than 10:1.  The END specimens in this case satisfied 
that  requirement,  but  there  is  some  question  as  to  the  effect  of  notch  thickness.    The 
maximum notch width should also be considerably larger than the grain size.  In this case 
a  notch  thickness  of  1.5  mm  was  used,  which  may  not  be  large  enough  relative  to  the 
granite rocks.  
 
C
Figure 3.10 shows fra
Figure 3.10 Fracture toughness vs. a/D 
0.00 0.06 0.12 0.18 0.24 0.30 0.36 0.42 0.48 0.54 0.60 0.66
0
100
200
300
400
500
600
700
800
900
1000
1100
K
Ic
a/D
 Fine-grained granite
 Large-grained granite
  76
There  is  some  dependency  of  K
Ic
  on  crack  length  (D  was  constant  at  approximately  50 
mm).  Usually when the crack length is long enough, the fracture process zone becomes 
fully developed before failure, resulting in the consumption of more energy and a higher 
fracture toughness value.  Figure 3.10 seems to contradict this, as K
Ic
 seems to decrease 
slightly with increasing crack length, but the amount of scatter in the data requires that a 
wider  range  of  crack  lengths  be  tested  before  suggesting  a  relationship  between  K
Ic
  and 
crack length for the END test. 
 
Effect of loading rate 
  that  the  loading  rate  was  varied  among  0.025,  0.003,  and  0.001  It  was  previously  noted
mm/s for the large-grained granite.  In terms of stress intensity, the corresponding loading 
rates were 0.0283, 0.0032, and 0.0011 / MPa m s .  At least 8 samples were tested at each 
rate so that an analysis of variance could be performed to quantitatively assess any effect 
of loading rate on K
Ic
.  At a confidence interval of 99% there is no significant difference 
between the fracture toughness values determined at different loading rates.  For the END 
wedge  split  test  it  is  recommended  that  the  load  line  displacement  be  no  greater  than 
0.025 mm/s. 
 
Test limitations 
In section 2.5.3 it was stated that fracture toughness specimens employing a notch instead 
of  a  sharp  crack  tend  to  underestimate  K
Ic
.    This  is  the  biggest  drawback  to  using  the 
END  specimen.    However,  the  purpose  of  the  END  test  is  to  provide  representative 
values  of  fracture  toughness  for  use  in  indexing  different  rocks  quickly  and  easily,  in 
regards  to  sample  preparation  and  testing.    From  this  standpoint  the  limiting  effect  of  a 
notch  does  not  seem  to  be  an  issue  since  the  results  of  the  END  tests  compared  so 
favorably to the Semi-Circular Bend tests.  For application to Level II testing it would be 
worthwhile  to  investigate  the  use  of  a  sharp  crack  (i.e.,  a  chevron  notch)  with  the  END 
specimen.   
  
3.2.3 APPLICATION OF THE END WEDGE SPLITTING TEST 
 
The  Edge-Notched  Disc  Wedge  Splitting  test  gives  a  representative  toughness  value  (in 
accordance  with  Level  I  testing)  and  is  a  simple,  core  based  specimen,  requiring  only 
straight-through  notch  and  straightforward  data  collection  for  the  determination  of  K
Ic
.  
The END test can be used for indexing and comparative purposes.  The test will enable a 
large  number  of  rock  types  and  specimens  to  be  tested  so  that  relationships  among 
breakage  energy,  fragment  size  distribution,  and  fracture  toughness  can  be  investigated, 
leading  to  the  development  of  an  improved  method  for  prediction  and  evaluation  of 
crushing equipment. 
 
  For  this  study  the  testing  procedure  follows  the  one  used  in  the  development  and 
verification of the END test.  A loading rate of 0.001 mm/s is to be used to ensure slow, 
stable  crack  growth.    T   roc e  tested.    The  average  en  specimens  of  each k  type  are  to  b
  77
specimen  dimensions  for  each  rock,  along  with  the  coefficient  of  friction  between  the 
wedge and rock, are given in Table 3.5. 
 
Table 3.5 Average specimen dimensions and coefficient of friction for tested rocks 
D  a  t   
Rock Type 
mm  mm  mm   
Shadwell Metabasalt  47.422  21.946  24.696  0.384 
Boscobel Granite  47.483  22.007  25.354  0.410 
Culpeper Grey Siltstone  47.597  20.668  25.042  0.386 
Culpeper Red Siltstone   47.589  22.289  25.166  0.372 
Leesburg Diabase  47.597  21.763  25.776  0.350 
Thornburg Granite  47.160  21.237  25.588  0.398 
Spotsylvania Granite  47.468  21.989  25.197  0.394 
 
 
3.3 SINGLE PARTICLE BREAKAGE TESTING 
 
In  any  size  reduction  process  the  breakage  of  any  individual  particle  occurs 
simultaneously  with  that  of  many  other  particles.    However,  each  individual  particle 
breaks  only  as  a  result  of  the  stresses  applied  to  it,  thus  isolation  of  the  single  particle 
acture event represents the most elementary process in size reduction and its study is of 
lication,  and  the  energy  input.    The  purpose  of 
ngle  particle  breakage  in  this  study  is  to  derive  crushing  energy  and  product  size 
d  
fracture toughness of the those rock types. 
 
  Th vices u  single e break dies we cribed in 
sectio f  those  te ystems   used  in  this  study.    Single  particle 
breaka ed  o g  the  Allis-Chalmers Energy ing  Test 
(HEC CT  is  a  s   confi TS  m s  testing 
system f  s eously ring  c   force actuator 
displa   to  accura lculate   energ ed  in  pa crushing 
(Allis .  Additi he HE em allows for flexible control of the 
force  ring  a  that  a  cified  can  be  followed.    It  is 
apable  of  running  under  displacement  control  at  2  m/sec,  using  an  inverted  haversine 
rofile permitting close simulation of the crushing cycle in an actual crusher.  The HECT 
er  sets,  speeds,  and  throws,  a  range  that 
cludes  essentially  all  crusher  operating  conditions,  consequently  distinguishing  itself 
fr
great  importance.    The  characteristics  of  crushing  that  need  to  be  measured  are  force 
applied at fracture, total fracture energy, specific fracture energy (taking into account the 
mass  of  the  particle),  and  the  product  size  distribution.    The  product  size  distribution  is 
used to determine the breakage function, which is influenced by the material properties of 
the  particle,  the  nature  of  the  stress  app
si
istribution data for the seven rock types, data which will ultimately be compared to the
e most common de sed for  particl age stu re des
n  2.4.2.    None  o sting  s   will  be
ge  tests  will  be  carri
T)  system.    The  HE
ut  usin
system 
  High 
gured  M
  Crush
aterial pecially
  that  is  capable  o imultan   measu rushing   and 
cement  in  order tely  ca   the  net y  utiliz rticle 
-Chamlers, 1985) onally t CT syst
or  displacement  du test  so  pre-spe profile 
c
p
system  can  simulate  a  wide  range  of  crush
in
  78
from other testing systems.  The Department of Mining and Minerals Engineering Rock 
Mechanics Laboratory at Virginia Tech is equipped with a HECT system. 
 
3.3.1 CONCEPTS OF THE HIGH ENERGY CRUSHING TEST 
 
The  purpose  of  the  HECT  test  is  to  obtain  the  breakage  functions  for  various  feed  sizes 
over  a  range  of  crushing  energies  or  size  reduction  levels.    A  number  of  pieces  of 
material  are  broken,  the  crushing  energy  is  recorded  for  each  piece,  and  the  products 
pooled  together.    The  size  distribution  of  the  pooled  pieces  is  determined  by  sieve 
analysis and the breakage function is determined in accordance with the methods outlined 
in section 2.4.3. 
 
  Since  the  crushing  energy  is  dependent  upon  the  amount  of  size  reduction  and  the 
e
ich  carries  the 
st material up towards a fixed platen.   A load cell affixed to the upper platen measures 
lly by a data acquisition system. 
uck Stone provided core boxes of the rock types to be used in this study.  Therefore the 
e  state  of  stress  within  a 
article  under  point  contact  loading  is  essentially  independent  of  the  particle  shape  and, 
m chanical  properties  of  the  material,  the  range  of  HECT  tests  is  defined  by  various 
reduction ratios.  The reduction ratio is the ratio of the geometric feed size to the closed 
side  set  of  the  HECT  device.    Allis-Chalmers  recommends  that  HECT  tests  be  run  for 
reduction  ratios  between  1.6  and  3.3.    It  is  also  recommended  that  various  feed  sizes, 
representing what might be seen in an actual crusher, be tested.  
 
  The  HECT  system  is  very  flexible  and  can  accommodate  a  wide  range  of  crusher 
conditions.    The  HECT  system  requires  the  specification  of  three  conditions:  the  closed 
side set, the throw, and the speed (or frequency).  These parameters are set to match the 
operating conditions of the crusher application in question. 
 
  The  HECT  system  itself  measures  only  two  parameters  during  testing,  the  crushing 
force  and  the  displacement  of  the  hydraulic  actuator.    A  linear  variable  differential 
transducer  (LVDT)  measures  the  displacement  of  the  hydraulic  ram,  wh
te
the crushing force.  These two signals are recorded digita
 
  The  following  section  outlines  the  exact  HECT  test  procedure,  set-up,  data 
acquisition and analysis, and specimen geometry used in this study. 
 
3.3.2 HECT TEST SET-UP AND PROCEDURE 
 
Test specimens 
L
specimen  geometry  for  the  HECT  test  is  limited  by  the  cylindrical  geometry  of  rock 
cores.  Normally single particle breakage is performed using irregular shaped particles or, 
in some instances, spherical particles are used.  There is no evidence in the literature that 
cylindrical specimens have been used previously for single particle breakage.  The main 
limitation of any specimen with a uniform, geometric shape is that it is not representative 
of  the  irregular  particles  fed  to  an  actual  crusher.    However,  th
p
  79
as  with  irregular  particles,  cylindrical  particles  fracture  due  to  induced  tensile  stresses 
that act along the load-line.   
 
  Table  3.6  gives  the  average  specimen  dimensions  for  each  rock  tested  using  the 
ECT  system.    A  cylindrical  specimen  with  a  thickness  approximately  half  of  the 
Reduction Ratio 1  Reduction Ratio 2 
H
diameter  was  chosen.    The  specimens  are  cut  using  a  radial  armed  saw  outfitted  with  a 
diamond wheel blade.  Approximately 20 specimens, or at least enough to exceed 2000 g 
by mass (total), are prepared for each test (i.e., varying reduction ratios) per rock type. 
 
Table 3.6 Average HECT test specimen dimensions 
D  t  Mass  D  t  Mass 
Rock Type 
mm  mm  g  mm  mm  g 
Shadwell Meta lt  7  basa 47.47 25.883  126.8  47.481  26.109  128.8 
Boscobel Gran 2  ite  47.49 25.668  117.0  47.464  26.041  117.5 
Culpeper Grey Siltsone  47.624  26.601  128.3  47.621  25.575  124.0 
Culpeper Red Siltsone   47.631  26.034  127.6  47.605  25.740  125.5 
Leesburg Diabase  47.587  26.575  141.2  47.608  26.038  139.1 
Thornburg Granite  47.174  26.728  125.1  47.192  26.064  119.9 
Spotsylvania Granite  47.454  26.673  128.3  47.431  26.069  125.2 
 
Test equipment and set-up 
The  HECT  system  is  a  uniquely  designed  MTS  810  materials  testing  system  (Figure 
3.11).  It is a closed loop servo-hydraulic materials testing system affixed with a 250 kN 
load cell.  The force-generating device in the HECT system is a high velocity hydraulic 
actuator.    The  actuator,  or  ram,  is  capable  of  speeds  up  to  2  m/s  and  has  a  maximum 
roke length of 152 mm.  The actuator is rated at 150 kN.  The displacement of the ram 
s, 
ith  the  closed  side  set  for  reduction  ratio  1  being  31.75  mm  and  for  reduction  ratio  2, 
16.00  mm.    Base erage  reduction 
ratios are 1.50 and 2.97.  The t istance travele  is fixed 
at 50.8  rdance with Allis-Chalmers  ion tha l tests b un at t ow 
(Allis-Chalmers,  1985).    The  speed  of  the  test  is  to  be  representative  of  an  actual  jaw 
cr  an a spe enc a sin gle ushe the 
te   will  be lose   spe he  sys uire the 
sp req  thu st is  at  ing cy  Hz 
(2
 
st
is measured by an LVDT mounted inside the hydraulic piston. 
 
  The closed side set, throw, and speed are the only parameters that need to be defined 
for  the  HECT.    The  closed  side  set  is  determined  based  on  the  required  amount  of  size 
reduction.  For this study it is proposed that tests be run at two different reduction ratio
w
d  on  the  average  dimensions  given  in  Table  3.6,  the  av
hrow (the linear d d by the actuator)
mm in acco direct t al e r hat thr
usher.  225 rpm is verage  ed/frequ y for  gle-tog  jaw cr r and 
sts  outlined  here   run  c   to  that ed.    T HECT  tem  req s  that 
eed be entered as a f
28 rpm).  
uency, s the te  set-up a crush  frequen of 3.8 
  80
The  test  is  run  under  displacement  control  with  the  actuator  displacement  following 
an inverted haversine profile, shown in Figure 3.12.  The combination of the given throw 
and frequency results in an average loading rate of 386 mm/s.  Breakage of one particle 
takes 0.26 seconds with full-scale displacement occurring at 0.13 seconds. 
 
Each particle is placed inside the crushing zone, as defined by the small platen affixed 
to the load cell (Figure 3.13).  The crushing chamber is surrounded by a plastic enclosure 
that  is  closed  and  locked  after  a  particle  is  loaded.    The  particle  is  then  broken.    After 
breakage the particle fragments are cleaned from the chamber and transferred to a plastic 
bag.  
 
Data acquisition 
A  digital  acquisition  system  accompanies  the  HECT  system  (upper  left  hand  corner  in 
Figure 3.11) and records and analyzes the raw data sent to it during each test.  The raw 
data is the displacement as measured by the LVDT and the load as measured by the load 
cell.    The  DATA  6000  Waveform  Analyzer  acquires  this  data  during  testing  and  is  also 
capable of processing it after testing.  The DATA 6000 has a sampling rate of 10 s. 
 
 
Figure 3.11 HECT system 
  81
  82
  The  DATA  6000  can  be  programmed  manually  or  by  using  a  program  supplied 
by Allis-Chalmers.  In each case the DATA  6000  is  configured  to  collect  the  raw  data 
Figure 3.12 Displacement profile of HECT actuator 
Frequency = 3.8 Hz
50.8
Time (s)
-50.8
0
0.132
A
c
t
u
a
t
o
r
 
d
i
s
p
l
a
c
e
m
e
n
t
 
(
m
m
)
Crushing Zone
Actuator
Fixed platen and Load cell
CSS + Throw
Figure 3.13 Schematic and picture of HECT loading configuration 
  83
fter  a  certain  triggering  level  has  been  met  and  to  process  the  data  according  to  the 
input  program.    For  this  study  the  DATA  6000  is  programmed  to  calculate  the  crushing 
energy  from  the  force  and  displacement  data.    The  DATA  6000  differentiates  the 
stroke/displacement signal in order to obtain the actuator velocity curve.  The force data 
is  multiplied  by  the  velocity  curve  in  order  to  determine  the  power  curve.    The  DATA 
6000  then  determines  the  points  at  which  crushing  began  and  ended.    This  interval  is 
bounded by the first non-zero force value and the maximum displacement of the actuator 
(i.e.,  the  point  at  which  the  actuator  begins  to  retract).    The  waveform  analyzer  then 
integrates  the  power  curve  over  this  interval  in  order  to  determine  the  total  crushing 
energy.    Figure  3.14  illustrates  a  typical  set  of  DATA  6000  recorded  and  processed 
curves.       
 
 
After  each  breakage  event  an  HP  7470A  plotter  produces  a  graph  similar  to  Figure 
3.14.  In addition to the plot, the DATA 6000 also provides numerical results for the total 
crushing  energy  and  the  peak  crushing  force.    These  values  are  recorded  for  each 
specimen.  The crushing energy is given in points, where 1 point is equal to one watt-
second.  The crushing energy is converted to specific comminution energy with units of 
kW-hours per metric ton.  Using Equation 2.31 and the peak crushing force, the dynamic 
tensile strength of the specimen is calculated. 
 
  Upon completion of one series of testing (all specimens tested for a given closed size 
set/reduction  ration),  a  sieve  analysis  is  performed  on  the  pooled  breakage  fragments.  
For  this  study  a 
a
2 series  will  be  used  with  an  initial  sieve  opening  of  38.1  mm.    The 
Figure 3.14 Recorded and processed curves from DATA 6000 
End of crushing
Start of crushing
  83
material will be sieved down to 0.074 mm (200 mesh).  A Ro-Tap sieve shaker is used to 
provide  uniform  circular  tapping  motion  to  the  sieves.    The  particle-size  distribution  is 
recorded  as  the  weight  percentage  retained  on  each  of  the  sieves  of  decreasing  size  and 
the percentage passed of the finest size.  This data is converted to the weight percentage 
passing each screen in order to determine the breakage function (see section 2.4.3).  For 
the cylindrical disk specimens of this study the parent size is taken to be the diameter of 
the disk. 
 
 
3.4 DATA REDUCTION AND MODEL DEVELOPMENT 
 
The  results  of  the  Edge  Notched  Disk  test  and  the  High  Energy  Crushing  Test  will  be 
compared  in  order  to  investigate  and  develop  relationships  between  fracture  toughness 
and  specific  comminution  energy  and  fracture  toughness  and  breakage  distribution.  
Since  the  closed  side  set  is  the  only  HECT  parameter  that  is  variable,  the  specific 
comminution  energy,  E
c
,  and  breakage  function,  B,  are  expected  to  be  functions  of 
fracture  toughness  and  reduction  ratio.    Regression  analysis  will  be  performed  using 
Microcal Origin 6.0. 
 
  The  fracture  toughness  results  will  also  be  used  to  modify  Equation  2.12,  which  is 
used to estimate capacity of a jaw crusher and is dependent upon a parameter related to 
the nature of the material crushed.  It is proposed that fracture toughness can be used to 
define  this  parameter.    Lab-scale  crushing  tests  will  be  conducted  in  order  to  determine 
how fracture toughness can be used in this regard. 
 
 
3.5 LABORATORY SCALE CRUSHING TESTS 
 
The power of the models developed from analysis of the END test and HECT test results 
will  be  assessed  by  comparing  the  predicted  and  actual  results  generated  from  lab-scale 
crushing  tests.    The  laboratory  crusher  results  will  also  be  used  to  develop  a  fracture 
toughness based model for the prediction of jaw crusher capacity.   
 
3.5.1 CRUSHING EQUIPMENT 
 
The  tests  will  be  performed  using  a  102  mm  by  152  mm  single-toggle,  overhead 
eccentric, Blake-type, Morris jaw crusher with a feed opening of 75 mm.   
 
3.5.2 ROCK SAMPLES 
 
Rock  samples  from  Luck  Stones  Boscobel,  Culpeper,  Shadwell,  and  Spotsylvania 
operations  will  be  used.    The  samples  were  collected  from  aggregate  product  piles.  
Figure 3.15 shows the feed size distribution of the four samples.   
 
  84
   
3.5.3 TEST PROCEDURE AND DATA ACQUISITION 
 
Each test will use approximately 35 kg of material.  The material is fed to the crusher and 
the time required to reduce the size of the feed material is recorded.  The open and closed 
side sets of the crusher are noted.  The crushed product is collected and a sieve analysis is 
conducted in order to determine the product size distribution. 
90
100
Figure 3.15 Feed size distribution of material fed to laboratory jaw 
4 5 6 7 8 9 10 20 30 40
0
10
20
30
40
50
60
70
80
 SMB
 BG
 CR
 SG
%
 
p
a
s
s
i
n
g
Particle size (mm)
 
 
  85
CHAPTER 4. EXPERIMENTAL RESULTS AND DISCUSSION 
 
 
4.1 FRACTURE TOUGHNESS 
 
Ten specimens of each rock type where tested in accordance with the END test procedure 
set  forth  in  section  3.2.    The  results  are  shown  in  Table  4.1.    Figure  4.1  displays  the 
tatistical scatter of test data for each rock (data points represent valid tests).  Appendix I 
ontains complete test data. 
Table 4.1 Fracture toughness results 
s
c
 
K
Ic
 ( SD) 
Rock Type 
No. of specimens 
tested 
No. of valid 
specimens 
MPa m  
Shadwell Metabasalt (SMB)  10  7  1.411  0.215 
Boscobel Granite (BG)  10  5  0.601  0.083 
Culpeper Gray Siltstone (CRS)  10  10  1.454  0.195 
Culpeper Red Siltstone (CGS)  10  10  1.501  0.262 
Leesburg Diabase (LD)  10  10  1.254  0.101 
Thornburg Granite (TG)  10  10  1.045  0.037 
Spotsylvania Granite (SG)  10  8  0.843  0.165 
 
Of the seventy rock specimens tested, ten did not qualify as valid specimens.  Three 
specimens of the Shadwell metabasalt and two specimens of the Spotsylvania granite did 
ot  meet  the  specimen  size  requirements  set  forth  by  Equation  2.40.    The  five  invalid 
ecimens of Boscobel granite did not fail at the notch (i.e., crack propagation ahead of 
o the friable nature of the Boscobel granite, 
hich  consists  of  large  grains  with  multiple  planes  of  weakness.    BG  samples  were 
hness  values  show  a 
ependence  on  grain  size.    The  results  in  Table  4.1  support  this  trend.    The  Culpeper 
d  sedimentary  rock  and  has  the  highest  fracture  toughness 
alue.    The  metabasalt  and  diabase  have  slightly  larger  grain  sizes  and  correspondingly 
ure toughness.  
here  may  be  a  propensity  for  the  K
Ic
  of  coarse-grained  rocks  to  be  underestimated  in 
ained  rocks  when  a  notched  specimen  is  used.    This  point  will  be 
onsidered when comparing fracture toughness and specific comminution energy. 
hat the two rocks are 
irtually  identical.    Statistical  analysis  of  the  separate  siltstone  results  shows  that  at  a 
n
sp
the crack tip did not occur).  This was due t
w
difficult to prepare as often times the rock core would crumble or break apart while being 
sawed. 
 
As  was  noted  in  section  2.5  and  section  3.2,  rock  fracture  toug
d
siltstone  is  a  very  fine-graine
v
have  slightly  lower  K
Ic
  values.    The  granites  follow  in  the  same  manner,  with  the 
Boscobel granite, consisting of fairly large grains, having the lowest fract
T
comparison  to  fine-gr
c
 
Figure 4.1 indicates that the Culpeper siltstone rocks (red and gray) had very similar 
results.  The mechanical properties given in Table 3.1 also show t
v
  86
3.0
+SD
2.5
Median value
-SD
Figure 4.1 Fracture toughness data for all valid specimens 
SMB TG SG LD BG CGS CRS
0.0
0.5
1.0
1.5
2.0
K
I
c
 
(
M
P
a
 
m
0
.
5
)
Rock Types
confidence  level  of  99%  the  average  fracture  toughness  values  are  not  significantly 
different.    Unless  significant  differences  in  red  and  gray  siltstone  behavior  are  observed 
during  single  particle  breakage,  the  test  values  of  the  two  rocks  will  be  pooled  and  the 
rocks considered as one type/sample. 
 
  For  all  valid  test  specimens  fracture  initiated  at  the  notch  tip.    Fracture  propagation 
did  not  a   surface 
planar.  The crack surfaces of the Spotsylvania and Thornburg granites wandered towards 
lanes  of  foliation  but  never  directly  along  them.    Crack  propagation  occurred  along 
lanes  of  weakness  represented  by  narrow  veins  of  mineral  infilling  in  the  Boscobel 
granite.    The  granites  were  more  susceptible  to  changes  in  heterogeneity  and  the 
continuity of crack growth was affected by changes in local structure ahead of a crack tip.  
The  metabasalt  and  diabase  demonstrated  somewhat  more  parallel  crack  propagation, 
with the Culpeper rocks almost always exhibiting parallel propagation.  All the rocks had 
somewhat non-planar crack surfaces, most likely due to friction and inter-locking.  
 
Figures  4.2  through  4.4  show  the  correlation  between  fracture  toughness  and  the 
major  mechanical  properties  of  each  rock  (compressive  strength,  tensile  strength,  and 
elastic modulus).  As would be expected, fracture toughness and tensile strength exhibit a 
strong correlation since tensile stresses are responsible for the opening of flaws or cracks 
lways  occur  in  a  direction  parallel  to  the  notch  tip  nor  was  the  crack
p
p
  87
Figure 4.2 Correlation between fracture toughness and tensile strength 
5.0 7.5 .0 12.5 15.0 20.0 22.5
0.8
K
I
c
 
(
M
T ile strength (MPa)
2.0
1.8
1.6
1.0
1.2
1.4
P
a
 
m
0
.
5
)
0.0 2.5
0.0
10 17.5 25.0
0.2
0.4
0.6
ens
2.0
1.0
1.2
1.4
1.6
1.8
M
P
a
 
m
0
.
5
)
0.8
K
I
c
 
(
within rock materials, particularly under conditions of point contact loading.  There does 
not seem to be much of a link between K
Ic
 and elastic modulus but it is interesting to note 
that  the  Culpeper  siltstones,  with  the  highest  fracture  toughness  values,  are  among  the 
softer rocks or lower modulus rocks.  Two vertical lines have been added to Figure 4.4, 
Figure 4.3 Correlation between fracture toughness and elastic modulus 
0 4 8 12 16 20 24 28 32 36 40
0.0
0.2
0.4
0.6
Elastic modulus (GPa)
  88
  89
one  at  69  MPa  and  the  other  at  138  MPa.    The  lines  represent  the  cut-offs  for  material 
ranked  soft  and  material  ranked  average,  for  crushing  applications,  according  to  Table 
2.2.  Above 138 MPa rocks are considered hard and supposed to require full power draw.  
In  Figure  4.4  it  can  be  seen  that five of the rocks would be considered hard but that the 
three rocks that showed the highest resistance to fracture are not the hardest according to 
the rankings.  
    
4.2 SINGLE PARTICLE BREAKAGE 
 
High  Energy  Crushing  Tests  were  conducted  in  accordance  with  the  procedure  set  forth 
in  section  3.3.    Approximately  twenty  specimens  of  each  rock  were  tested  at  each 
reduction ratio.  The reduction ratio for the first series of tests was 1.50, corresponding to 
a  closed  side  set  of  31.75  mm.    The  second  series  of  tests was conducted at a reduction 
ratio of 2.97 with the closed side set equal to 16.00 mm.  Data collection included particle 
strength  (tensile),  specific  comminution  energy,  and  breakage  size  distribution.    A 
comprehensive catalog of HECT data can be found in Appendix II and Appendix III. 
 
  Again,  the  red  and  gray  Culpeper  siltstones  did  not  exhibit  different  behavior,  with 
the  data  for  each  rock  not  statistically  different.    From  this  point  forward  the  individual 
results for the red and gray siltstones will be shown since they were considered separate 
rocks when tested, but for comparative purposes (section 4.3) their results will be pooled 
and presented as one rock, Culpeper siltstone. 
 
Figure 4.4 Correlation between fracture toughness and compressive strength
0   25   50   75   100   125   150   175   200   225   250
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
2.0
K
I
c
 
(
M
P
a
 
m
0
.
5
)
Compressive strength (MPa)
4.2.1 PARTICLE STRENGTH 
 
Table  4.2  summarizes  the  dynamic  tensile  strength  results  from  both  HECT  series.  
Particle  strength  data  was  collected  in  order  to  compare  traditional  rock  properties 
collected under conditions of slow displacement with a strength property measured under 
the  dynamic  or  high  strain  rate  conditions  seen  in  a  jaw  crusher.    The  compressive 
strength,  tensile  strength,  and  elastic  modulus  of  each  rock  were  tested  under 
displacement  control  at  a  rate  of  0.003  mm/s.    As  was  mentioned  in  section  3.3,  the 
crushing  tests  were  conducted  at  a  rate  of  386  mm/s.    Additionally  the  dynamic  tensile 
strength  was  compared  to  the  fracture  toughness  in  order  to  verify  that  mode  I  crack 
opening is applicable to the crack displacement behavior of particles under point contact 
loading. 
 
Table 4.2 Tensile strength measured in HECT tests 
Reduction Ratio 1  Reduction Ratio 2 
Tensile Strength ( SD)  Tensile Strength ( SD) 
Rock Type 
MPa   MPa  
Shadwell Metabasalt (SMB)  21.69  4.10  24.80  6.89 
Boscobel Granite (BG)  11.90  5.84  10.03  3.46 
Culpeper Gray Siltstone (CRS)  27.51  3.22  26.29  2.85 
Culpeper Red Siltstone (CGS)  28.89  3.69  26.41  4.93 
Leesburg Diabase (LD)  26.45  3.93  25.72  3.28 
Thornburg Granite (TG)  17.60  4.72  16.16  2.88 
Spotsylvania Granite (SG)  13.58  4.13  15.71  3.86 
 
  Care  had  to  be  taken  when  determining  the  load  that  caused  initial  fracture.    The 
HECT/DATA 6000 set-up only records the peak load, which in some instances occurred 
after  fracture  had  already  initiated  and  represented  a  smaller  daughter  fragment  being 
loaded.    In  these  instances  the  force  curve  produced  by  the  DATA  6000  acquisition 
system  was  used  to  distinguish  the  point  at  which  fracture  first  initiated  and  the 
corresponding load on the disk specimen at that time.  In all other instances the peak load 
recorded by the DATA 6000 represented the failure load.   
 
On average the increased displacement rate resulted in about a 29% increase in tensile 
strength for each rock.  The data in table 4.2 indicates that, for each rock, there is little to 
no  change  in  the  dynamic  tensile  strength  at  the  two  reduction  ratios.    The  change  in 
reduction ratio would not be expected to influence the measured particle strength, as there 
is  no  c he  ratios,  only  a  c   peak 
fractur uency 
hange  in  strain  rate  between  t hange  in  total  strain.    The
e load would be more affected by a change in either the crushing speed/freq
or the throw. 
 
Figures 4.5 through 4.8 show the relationships between dynamic tensile strength and 
tensile  strength,  elastic  modulus,  compressive  strength,  and  fracture  toughness.    Figures 
  90
4.5,  4.6,  and  4.7  indicate  that,  for  the  rocks  tested  here,  the  relationships/correlations 
between dynamic tensile strength and tensile strength, elastic modulus, and compressive 
strength  are  almost  identical  to  those  between  fracture  toughness  and  those  latter 
properties (compare to Figures 4.2  4.4).  This is evident also by the strong correlation 
between dynamic tensile strength and fracture toughness seen in Figure 4.8.  The strong 
correlations of dynamic tensile strength with tensile strength and fracture toughness again 
signify  that  fracture  under  point  contact  loading  is  due  to  induced  tensile  stresses.    It  is 
also clear that under rapid loading the magnitude of the induced tensile stresses increases 
but  the  fundamental  mechanical  behavior  of  rocks  is  unchanged  (i.e.,  the  increase  in 
tensile strength is approximately proportional for all rocks). 
 
  The  relationships  among  fracture  toughness,  dynamic  tensile  strength,  and  tensile 
strength  highlight  the  application  of  fracture  toughness  as  a  strength  parameter.    When 
the  proper  crack  opening  displacement  mode  is  identified  for  a  given  situation,  fracture 
toughness effectively characterizes a materials resistance to fracture.  In the case of jaw 
crushers,  tensile  stresses  open  cracks  and  the  ability  of  mode  I  fracture  toughness  to 
assess a rocks resistance to breakage is shown by the strong correlation between fracture 
toughness  and  both  tensile  strength  measurements.    In  other  instances  though,  for 
example  cone  crushing,  particles  may  also  be  subjected  to  shear  stresses  and  mode  III 
(mixed-mode/tearing) fracture toughness may be more applicable.  The ability of fracture 
toughness  to  characterize  a  rocks  resistance  to  fracture  is  dependent  upon  an 
understanding of the stress situation within the particle under  the load action of the size 
reduction machine. 
Figure 4.5 Correlation between dynamic tensile strength and tensile strength
2.5 15.0 17.5 20.0 22.5 25.0
8
32
40
D
y
h
 
(
P
a
)
ength (MPa)
o 1 36  Reduction rati
 Reduction ratio 2
28
M
12
16
20
24
n
a
m
i
c
 
t
e
n
s
i
l
e
 
s
t
r
e
n
g
t
0.0 2.5 5.0 7.5 10.0 1
0
4
Tensile Str
  91
Figure 4.6 Correlation between dynamic tensile strength and elastic modulus
0 4 8 12 16 20 24 28 32 36 40
0
4
8
12
16
D
y
n
a
m
i
c
 
t
e
n
Elastic modulus (GPa)
40
20
24
28
32
36
s
i
l
e
 
s
t
r
e
n
g
t
h
 
(
M
P
a
)
 Reduction ratio 1
 Reduction ratio 2
0
36
n
s
 Reduction ratio 1
40
20
24
28
32
i
l
e
 
s
t
r
e
n
g
t
h
 
(
M
P
a
)
 Reduction ratio 2
4
8
12
16
D
y
n
a
m
i
c
 
t
e
0 25 50 75 100 125 150 175 200 225 250
Compressive strength (MPa)
Figure 4.7 Correlation between dynamic tensile strength and compressive strength
 
  92
 
4.2.2 SPECIFIC COMMINUTION ENERGY 
 
The  DATA  6000  waveform  analyzer  developed  a  power  curve  for  each  test  specimen 
using the stroke and force input signals.  The area under the power curve, initiating at the 
point  of  first  fracture  and  ending  at  the  moment  the  actuator  began  to  retract,  was 
calculated  by  the  DATA  6000  in  order  to  determine  the  net  crushing  energy.    The  net 
crushing energy divided by the mass of the specimen is the specific comminution energy, 
the energy required to reduce the size of the particle for the given operating conditions (in 
kilowatt-hours  per  metric  ton).    Table  4.3  lists  the  specific  comminution  energy  results 
for each rock at each reduction ratio.  Individual specimen data can be found in Appendix 
II. 
 
Figures  4.9  through  4.15  are  sample  HECT  plots  representing  typical/average 
behavior for each rock at each reduction ratio.  The visual quality of the plots is poor and 
the  velocity  curve  developed  by  the  DATA  6000  is  quite  ragged.    This  is  a  result  of 
differentiation  of  the  digitally  recorded  stroke  signal  processed  from  an  analog  input 
signal.  Despite the ragged appearance there is little to no difference between the digitally 
processed  velocity  curve  and  one  developed  from  direct  differentiation  of  the  analog 
signal (Allis-Chalmers, 1985).  Also note that for the stroke signal (BUF.A1) one volt is 
equivalent to 10.16 mm and for the load signal (BUF.A2) one volt is equivalent to 15.57 
kN. 
Figure 4.8 Correlation between dynamic tensile strength and fracture toughness
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
0
4
8
12
16
D
y
n
a
m
i
c
 
t
e
n
s
K
Ic
 (MPa m
0.5
)
20
24
28
32
36
40
e
 
s
t
r
e
n
g
t
h
 
(
M
P
a
)
 Reduction ratio 1
 Reduction ratio 2
i
l
  93
 
Table 4.3 Specific comminution energy results from the HECT 
Reduction Ratio 1  Reduction Ratio 2 
E
c 
E
c 
Rock Type 
kWh/t  kWh/t 
Shadwell Metabasalt  0.407  0.550 
Boscobel Granite  0.105  0.230 
Culpeper Grey Siltstone  0.292  0.350 
Culpeper Red Siltstone   0.298  0.320 
Leesburg Diabase  0.371  0.477 
Thornburg Granite  0.272  0.384 
Spotsylvania Granite  0.233  0.356 
 
Figures 4.9 through 4.15 show the behavior of the rocks under jaw crusher conditions.  
At  the  start  of  the  crushing  cycle  the  actuator,  which  represents  the  swing  jaw,  moves 
towards the particle following the predefined haversine path.  When the actuator comes in 
contact with the particle there is a sharp increase in the load and a corresponding increase 
in power consumption.  Failure of the particle occurs quite rapidly (within approximately 
1 s) and is followed by a sharp decrease in load.  The power also drops off after failure.  
Since particle failure requires only small amount of strain input the actuator continues to 
move  upward  (i.e.,  the  crusher  chamber  continues  to  close).    Fragmented  particles 
remaining  inside  the  crushing  chamber  (Figure  3.13)  with  dimensions  greater  then  the 
closed  side  set  are  crushed  under  the  continued  displacement  of  the  actuator.    This  is 
depicted  by  post-failure  increases  in  the  load  and  power  signals.    After  traveling  the 
throw distance the actuator begins to retract and particle crushing ends. 
 
  The HECT plots along with the results in Table 4.3 reveal that the rocks that required 
the  most  energy  to  crush  also  experienced  the  most  secondary  breakage.    Secondary 
breakage  is  the  crushing  of  daughter  fragments  after  particle  failure.    The  Culpeper 
siltstones  exhibited  the  smallest  amounts  of  secondary  breakage.    Even  the  Boscobel 
granite,  which  on  average  required  less  energy  to  crush  than  the  siltstones,  underwent 
more  secondary  crushing,  i.e.,  a  larger  percentage  of  the  total  specific  comminution 
energy  was  a  result  of  secondary  breakage.    The  siltstones  required  more  energy  to 
initiate fracture due to their high strength, low elastic modulus behavior, and thus more of 
the actuator displacement was used during particle failure.  In comparison to the other 
rocks  the  Culpeper  siltstones  take  approximately  43%  to  94%  more  strain  to  fail  (based 
on tensile strength and elastic modulus). 
 
The  strain  at  failure  is  very  small  in  comparison  to  the  total  displacement  of  the 
actuator and a more significant explanation for the siltstones behavior is needed.  During 
testing  it  was  observed  that  the  siltstones  tended  to  shatter  producing  thin,  flat  daughter 
fragments.      This  was  a  result  of  ideal  disk  splitting.    Under  diametrical  compression 
  94
Figure 4.9 Typical BG HECT behavior: CSS 1 (top) and CSS 2 (bottom) 
  95
Figure 4.10 Typical CRS HECT behavior: CSS 1 (top) and CSS 2 (bottom) 
  96
Figure 4.11 Typical CGS HECT behavior: CSS 1 (top) and CSS 2 (bottom) 
  97
Figure 4.12 Typical LD HECT behavior: CSS 1 (top) and CSS 2 (bottom) 
  98
Figure 4.13 Typical SG HECT behavior: CSS 1 (top) and CSS 2 (bottom) 
  99
Figure 4.14 Typical SMB HECT behavior: CSS 1 (top) and CSS 2 (bottom) 
  100
 
Figure 4.15 Typical TG HECT behavior: CSS 1 (top) and CSS 2 (bottom) 
  101
  102
energy required for fracture is dependent upon, among other things, the stress applied to 
the particle, the strength of these correlations is not unexpected.  More significantly, the 
correl
fracture initiates at the center of a disk and typically propagates in a straight line towards 
the  load  contact  points.    Additional  displacement  results  in  fracture  planes  on  each  side 
and parallel to the first plane.  The result is rectangular plate-like fragments.  Under rapid 
displacement  rates  of  the  HECT  system  these  fragments  move  outward  and  away  from 
the  contact  points  at  velocities  greater  than  the  velocity  of  the  actuator.    The  daughter 
fragments of the Culpeper rocks either exited the crushing chamber or fell in such a way 
that  their  relative  thinness,  in  comparison  to  the  CSS,  prevented  any  further  crushing.  
Crack  propagation  in  the  other  rocks,  based  on  observations  from  fracture  toughness 
testing,  most  likely  did  not  occur  in  a  straight  path,  resulting  in  larger  and  more 
irregularly  shaped  daughter  fragments  that  were  subject  to  additional  breakage.    These 
rocks experienced more cleavage fracture, particularly at the lower reduction ratio.   This 
also  implies  that  shatter  and  cleavage  are  not  only  functions  of  the  type  of  crushing 
machine used but also of the fracture behavior of the material.  
 
The  increase  in  reduction  ratio  resulted  in  an  increase  in  specific  comminution 
energy.  The increase in energy was due to the larger amount of secondary breakage that 
occurs at larger reduction ratios.  Since the energy required for initial fracture essentially 
did not vary between reduction ratios (assuming load at first fracture and dynamic tensile 
strength  are  representative  of  the  energy  required  to  initiate  fracture),  the  amount  of 
crushing that occurs after initial breakage is dependent upon the post-failure strain input 
into  the  particle.    A  decrease  in  the  closed  side  set  of  the  HECT  system  results  in  more 
post-peak displacement and significant secondary crushing of daughter fragments. 
 
Although  an  energy  increase  due  to  an  increase  in  reduction  ratio  was  seen  for  each 
rock changes in crusher reduction ratio (or CSS) cannot be expected to have a universal 
effect  on  all  rocks.    Table  4.3  indicates  that  the  Culpeper  siltstones  saw  the  smallest 
increase  in  specific  comminution  energy,  an  average  of  0.040  kWh/t.    The  other  five 
rocks  exhibited  an  increase  that  ranged  between  0.106  -  0.143  kWh/t  (average  of  0.122 
kWh/t).  The dissimilar behavior, or variance in the degree of behavior, of the Culpeper 
siltstone  is  a  result  of  the  mechanical  and  fracture  properties  of  the  rock.    Again  the 
shattering behavior shown by the siltstones resulted in the exiting of daughter fragments 
from the crushing chamber.  More secondary breakage of the siltstone is only possible if 
the crusher speed is increased such that the actuator is able to close in and make contact 
with daughter fragments before they exit the crushing chamber.    
 
The specific comminution energy results were compared to the mechanical properties 
of the rocks, as well as the fracture toughness results and dynamic tensile strength results.  
Figures  4.16  -  4.20  show  the  correlations  between  E
c
  and  these  properties  for  each 
reduction ratio.  The strongest correlations are between E
c
 and tensile strength and E
c
 and 
fracture  toughness.    Since  tensile  strength  and  mode  I  fracture  toughness  are  most 
representative of the stress situation within a particle under point contact loading and the 
  E
c
.   ation  between  E
c
  and  K
Ic
  is  stronger  than  that  between  tensile  strength  and
  103
ratio  are  made.    The  energy  required  to  crush  rock  material  is  a  clearly  function  of  its 
resistance to fracture and the amount of strain input by the crushing machine. 
Assessing the fracture resistance of a rock particle in terms of a property that relates the 
fracture strength of the particle to the presence of flaws within the particle normalizes the 
effect and variation of stress magnitude.  Tensile strength is an ultimate strength criterion 
and  is  based  strictly  on  the  magnitude  of  stress  that  a  particle  can  withstand.    It  is  very 
applicable  to  situations  where  tensile  stresses  lead  to  fracture  by  opening  cracks  and 
flaws  but  it  is  subject  to  more  variation  since  no  account  is  made  for  the  presence  or 
geometry of those flaws.  
 
A limiting relationship between E
c
 and elastic modulus is evident in Figure 4.17, with 
a  stronger  correlation  for  the  larger  reduction  ratio.    The  specific  comminution  energy 
seems to increase exponentially with elastic modulus, although more very stiff rocks (E > 
28  GPa)  would  need  to  be  tested  in  order  to  suggest  such  a  relationship.    Some 
correlation  between  specific  comminution  energy  and  compressive  strength  is  indicated 
in Figure 4.18 for reduction ratio 1.   The correlation loses its power significantly with an 
increase in reduction ratio.  Compressive strength most likely correlates fairly well with 
E
c
 at low reduction ratios where power consumption is more dependent upon the material 
strength.  At higher reduction ratios the specific comminution energy is more dependent 
on the fracture and elastic properties, i.e., fracture strength, crack propagation, and strain 
energy  release  rate.    Figure  4.20  illustrates  the  relationship  between  E
c
  and  dynamic 
tensile  strength,  which  is  more  scattered  than  expected  due  to  the  low  dynamic  tensile 
strength of the Shadwell metabasalt.   
 
The relationship between E
c
 and K
Ic
 in Figure 4.19 may be skewed somewhat by the 
results of the Edge Notched Disk test.  It is widely accepted that rock fracture toughness 
measurements and values show a dependence on grain size.  This dependency is evident 
in  the  fracture  toughness  values  presented  here  and  was  discussed  in  section  4.1.    The 
data presented in Figure 4.19 show that the Culpeper siltstones have the highest fracture 
toughness  values  yet  the  corresponding  specific  comminution  energy  values  for  those 
rocks  fall  in  the  middle  to  bottom  third  of  the  overall  results.    In  fact  the  relationship 
between E
c
 and K
Ic
 is essentially linear except for the Culpeper siltstones.  Based on the 
grain size effect, the fracture toughness values of the other, larger grained rocks may be 
underestimated,  resulting  in  a  left-shift  of  their  data.    If  this  is  the  case  the  correlation 
between E
c
 and K
Ic
 may in fact be stronger than is demonstrated. 
 
The  most  important  aspect  of  the  HECT  specific  comminution  energy  results  is  that 
fracture toughness correlates the strongest with E
c
 in comparison to other rock properties.  
Based on the physics of fracture and the fact that the energy required to fracture a particle 
is  dependent  on  the  stresses  applied  to  it,  its  mechanical  properties,  and  the  presence  of 
flaws, it is no revelation that Mode I fracture toughness is related to the energy required 
to  reduce  the  size  of  a  particle  under  point  contact  loading.    Another  important  result  is 
that the relationship between E
c
 and K
Ic
 remains evident even when changes in reduction 
  104
0.0 2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0 22.5 25.0
0.00
0.06
0.12
0.18
0.24
0.30
0.36
0.42
0.48
0.54
0.60
E
c
 
(
k
W
h
/
t
)
Tensile strength (MPa)
 Reduction ratio 1
 Reduction ratio 2
Figure 4.16 Correlation between E
c
and tensile strength 
Figure 4.17 Correlation between E
c
 and elastic modulus 
0 4 8 12 16 20 24 28 32 36 40
0.00
0.06
0.12
0.18
0.24
0.30
0.36
0.42
0.48
0.54
0.60
E
c
 
(
k
W
h
/
t
)
Elastic modulus (GPa)
 Reduction ratio 1
 Reduction ratio 2
Figure 4.18 Correlation between E
c
 and compressive strength 
0 25 50 75 100 125 150 175 200 225 250
0.00
0.06
0.12
Compressive strength (MPa)
0.18
0.24
0.30
0.36
0.42
0.48
0.54
0.60
E
c
 
(
k
W
h
/
t
)
 Reduction ratio 1
 Reduction ratio 2
Figure 4.19 Correlation between E
c
 and fracture toughness 
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
0.00
0.06
0.12
0.18
K
Ic
 (MPa m
0.5
)
0.60
0.24
0.30
0.36
0.42
0.48
0.54
E
c
 
(
k
W
h
/
t
)
 Reduction ratio 1
 Reduction ratio 2
  105
0 3 6 9 12 15 18 21 24 27 30
0.00
0.06
0.12
0.18
0.24
0.30
E
c
 
(
k
W
h
/
t
)
Dyna
   
4.2.3 BREAKAGE FUNCTION 
 
After  single  particle  breakage  was  conducted  the  pooled  samples  for  each  rock  were 
sieved in order to determine the fragmented size distribution.  A  2 sieve series was used 
starting  with  an  initial  opening  size  of  38.1  mm  and  finishing  with  a  0.075  mm  mesh 
opening size.  The breakage distribution was determined by normalizing the fragmented 
sizes  with  the  parent  size.    The  parent  size  was  taken  to  be  the  average  diameter  of  the 
tested specimens.  Appendix III contains the fragmented size distribution data. 
 
  he breakage functions for each rock, and for both reduction ratios, are displayed in 
Fig
ccurs in an 
actual  crusher.    If  the  irregular  particles  crushed  in  actual  machines  are  idealized  as 
T
ures  4.21  and  4.22.    In  each  instance  all  material  passed  at  a  fragmented  size  of 
approximately  0.8  of  the  parent  size.    In  fact  for  each  rock  100%  of  the  fragmented 
material would have passed at a size close to 0.54 of the parent size.  This is because disk 
specimens  with  thicknesses  approximately  one  half  of  their  diameter  were  used  in  this 
study.  Under (small displacement) point contact loading, a disk, with dimensions DDt, 
tends  to  split  in  half  and  the  resultant  fragments  have  approximate  dimensions  of 
D0.5Dt.    When  the  thickness  is  close  to  the  radius  of  the  disk,  100%  of  the  broken 
fragments are retained on a sieve that has an opening equal to one-half of the parent size.  
There was some debate as to whether or not this behavior is typical of what o
0.36
0.42
0.48
0.54
0.60
 Reduction ratio 1
 Reduction ratio 2
Figure 4.20 Correlation between E
c
 and dynamic tensile strength 
mic tensile strength (MPa)
  106
  107
spheres,  then  disk  specimens  with  thicknesses  near  one  half  of  their  diameters  are  also 
representative  of  irregular  particles.    Small  displacement,  point  contact  loading  of  a 
sphere will result in four lune pieces with dimensions D0.5D0.5D, which also results 
in 100% of the fragmented pieces being retained on a sieve that has an opening equal to 
one-half of the parent size.  Of course a sphere, or an equivalent disk, may not adequately 
describe  irregular  particles  so  the  results  of  this  study  may  be  limited/effected  by  the 
chosen test specimen.  
 
For  the  smaller  reduction  ratio  (Figure  4.21)  it  can  be  seen  that  the  Boscobel, 
Shadwell,  Leesburg,  Spotsylvania,  and  Thornburg  rocks  exhibit  similar  distributions 
between  the  parent  size  and  the  0.1  size.    At  sizes  smaller  than  1/10
th
  of  the  parent  size 
the distributions diverge.  At larger fragment sizes these rocks experience approximately 
the  same  degree  of  cleavage  but  the  proportion  of  material  in  the  smaller  size  ranges 
varies  according  to  the  degree  of  shatter  each  rock  is  subjected  to.    Figure  4.21  also 
touches upon the observation made in the previous section that fracturing of the Culpeper 
siltstones  was  a  shatter-dominated  event.    The  siltstones  are  shown  to  have  a  uniform 
distribution  and  their  breakage  functions  are  not  a  mixture  of  separate  size  populations.  
The  other  five  rocks  exhibit  multi-modal  behavior  and  their  breakage  functions  are  the 
more typical m . 
Figure 4.21 Breakage functions for each rock at reduction ratio 1 
0.0 1
0.01
0.1
1
1 0.1
B
 
(
d
i
,
d
1
)
d
i
/d
1
 BG
 CGS
 CRS
 LD
 SMB
 SG
 TG
ixture of separate populations caused by shatter and cleavage
  108
 
Increasing  the  reduction  ratio  resulted  in  a  breakage  distributions  consisting  of  a 
larger  proportion  of  fine  progeny  sizes  (i.e.,  the  percent  passing  each  progeny  size 
increased).  Figure 4.22 indicates that the distribution modulus of the larger progeny size 
populations  is  approximately  the  same  for  each  rock  but  the  proportion  of  fragmented 
material in the larger sizes is not.  There is no consistency among the rocks distribution 
modulus  or  material  proportion  in  the  finer  sizes  (less  than  1/10
th
  the  parent  size).    The 
Culpeper  siltstones  again  yielded  fairly  uniform  breakage  functions.    The  siltstones  also 
did  not  experience  as  great  of  an  increase  in  the  percentage  passing  each  fragment  size.  
This is most likely a result of the small degree, relative to the other rocks, of secondary 
crushing  the  siltstones  were  exposed  to  and  corresponds  to  the  proposition  made  in  the 
previous section that an increase in reduction ratio does not affect the siltstones breakage 
behavior as much as the other rocks. 
  Figures  4.23    4.29  are  the  separate  breakage  functions  for  each  rock  at  each 
reduction  ratio.    The  change  in  breakage  distribution  with  reduction  ratio  can  be  seen 
more  clearly  for  each  rock.    At  the  lower  reduction  ratio  the  breakage  functions  are  a 
mixture of two separate size populations, except for the Culpeper rocks as noted earlier.  
At the higher reduction ratio the larger size population becomes bimodal due to increased 
secondary bre
Figure 4.22 Breakage functions for each rock at reduction ratio 2 
0.01 0.1 1
0.01
0.1
1
B
 
(
d
i
,
d
1
)
d
i
/d
1
 BG
 CGS
 CRS
 LD
 SMB
 SG
 TG
akage and shattering of progeny fragments. 
0.01 0.1 1
0.01
0.1
1
 Reduction ratio 1
Figure 4.23 Breakage functions for Boscobel granite 
 Reduction ratio 2
B
 
(
d
i
,
d
1
)
d
i
/d
1
0.01 0.1 1
0.01
B
 
(
d
d
i
/d
1
Figure 4.24 Breakage functions for Culpeper gray siltstone 
0.1
1
 Reduction ratio 1
 Reduction ratio 2
i
,
d
1
)
  109
Figure 4.25 Breakage functions for Culpeper red siltstone 
0.01 0.1 1
0.01
0.1
1
 Reduction ratio 1
 Reduction ratio 2
B
 
(
d
i
,
d
1
)
d
i
/d
1
0.1
 Reduction ratio 2
i
,
d
1
)
0.01 0.1 1
0.01
1
B
 
(
d
d
i
/d
1
Figure 4.26 Breakage functions for Leesburg diabase 
 Reduction ratio 1
  110
 
Figure 4.27 Breakage functions for Spotsylvania granite 
0.01 0.1 1
0.01
0.1
B
 
(
d
i
,
d
1
)
1
 Reduction ratio 1
 Reduction ratio 2
d
i
/d
1
Figure 4.28 Breakage functions for Shadwell metabasalt 
0.01 0.1 1
0.01
d
i
/d
1
0.1
1
 Reduction ratio 1
 Reduction ratio 2
B
 
(
d
i
,
d
1
)
  111
  Conventional  wisdom,  based  on  prior  research  (see  section  2.4),  dictates  that  an 
increase in energy intensity changes only the proportion of the separate populations that 
make up a breakage function and not their distribution.  Agreement with this perception 
is  found  in  the  preceding  figures.    The  distribution  of  the  finer  size  population  (sizes 
below  1/10
th
  of  the  parent  size)  does  not  change  but  the  proportion  of  material  in  that 
population does change.  The proportion of material in the larger size fractions decreases 
because of rebreakage of cleavage progeny fragments.  The distribution modulus does not 
change, although a change in the distribution above sizes half the parent size can be seen 
in some instances.  This indicates that rebreakage of daughter particles initially produced 
by  cleavage  is  occurring  (note  that  the  slope/distribution  modulus  above  0.5  is 
approximately the same as that of the fine size population). 
 
  In  section  2.4.3  the  t
10 
parameter  was  introduced.    It  is  the  percent  of  progeny 
particles passing a size one-tenth of the initial original particle size.  t
10
 is employed as a 
characteristic size reduction parameter and has been related to the energy absorbed during 
a  single  particle  breakage  test.    For  crushing  applications  t
10
  is  usually  in  the  range  of 
10% to 20%.  In order to determine t
10
, and other t
n
 values, from the breakage distribution 
data,  a  MathCAD  program  was  developed  so  that  cubic  spline  interpolation  of  the 
measured data could be performed.  Table 4.4 lists the value of t
10
 for each rock at each 
reduction  ratio.    As  expected  t
10
  increased  with  an  increase  in  reduction  ratio  and  the 
results fit close to the range of values typical of crushing.  In conjunction with Table 4.3 
Figure 4.29 Breakage functions for Thornburg granite 
0.01 0.1 1
0.01
0.1
1
 Reduction ratio 1
 Reduction ratio 2
B
 
(
d
i
,
d
1
)
d
i
/d
1
  112
  113
the results listed below show that the rocks that had the largest percentage increase in E
c
 
also had the largest percentage increase in t
10
.  
 
Table 4.4 Values of t
10
 for each rock and reduction ratio 
Reduction Ratio 1  Reduction Ratio 2 
t
10 
t
10 
Rock Type 
%   %  
Shadwell Metabasalt  8.1  15.0 
Boscobel Granite  7.4  17.0 
Culpeper Grey Siltstone  11.7  16.8 
Culpeper Red Siltstone   12.9  12.8 
Leesburg Diabase  8.7  15.8 
Thornburg Granite  8.7  15.7 
Spotsylvania Granite  8.5  20.4 
 
  For  each  specific  rock  the  t
10
  parameter  is  a  function  of  the  energy  input.    On  the 
whole,  looking  at  the  grouped  results  for  all  rocks,  t
10
  is  more  so  a  function  of  the 
reduction  ratio.    Figure  4.30  shows  that  t
10
  increases  with  an  increase  in  reduction  ratio 
but  that  at  each  reduction  ratio  it  does  not  change  significantly  as  E
c
  increases.    This  is 
because  at  a  given  reduction  ratio  specific  comminution  energy  is  a  function  of  the 
material behavi ws that t
10
  varies 
Figure 4.30 Relationship between t
10 c
and E
0.00 0.06 0.12 0.18 0.24 0.30 0.36 0.42 0.48 0.54 0.60
0.0
2.5
5.0
7.5
10.0
12.5
15.0
17.5
20.0
22.5
25.0
t
1
0
 
(
%
)
Specific comminution energy (kWh/t)
 Reduction ratio 1
 Reduction ratio 2
or.  Thus a plot of t
10
 versus fracture toughness also sho
  114
very little as K
Ic
 changes, although the change in t
10
 with an increase in reduction ratio is 
still  evident  (Figure  4.31).    Figures  4.32  and  4.33  indicate  that  other  t
n
  values  show 
similar behavior when compared to fracture toughness, particularly smaller values.  t
2
 and 
t
4
  show  a  fairly  wide  degree  of  variation  as  fracture  toughness  changes.    t
25
,  t
50
,  and  t
75
 
fluctuate in a manner similar to t
10
.   
 
It is difficult to observe any noticeable trend or correlation between t
n
 values and K
Ic
.  
In  regards  to  t
n
  and  K
Ic
,  the  conventional  thinking  would  be  that  as  fracture  toughness 
increases  the  percent  passing  each  t
n
  value  decreases,  or  tougher  rocks  would  fracture 
more coarsely.  However, high fracture toughness rocks require more strain energy input 
in  order  to  propagate  flaws  and  thus  tend  to  shatter  upon  fracture,  which  produces  finer 
sized  fragments.    The  siltstone  is  an  example  of  this  behavior  and  in  Figures  4.32  and 
4.33 it can be seen that more material passes at t
2
 and t
4
 for the tough siltstones.  The 
relationship  between  t
n
  and  K
Ic
  is  complicated  further  by  grain  size.    Rock  fracture 
toughness  has  been  shown  to  be  dependent  upon  grain  size  (in  this  study  and  in  the 
literature), as large grained rocks tend to have lower K
Ic
 values.  Subsequently, on the one 
hand,  large  grained  rocks  are  easy  to  fracture  and  should  produce  a  more  finely  sized 
breakage distribution based on their fracture toughness.  On the other hand, large grained 
rocks  tend  to  fracture  along  grain  boundaries  and  progeny  fragment  size  will  be 
dependent on the grain size and may end up being coarser in comparison to tougher, fine-
grained rocks.  The Boscobel granite is a good example of this behavior.  In Figures 4.32 
Figure 4.31 Relationship between t
10
 and K
Ic
 
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
0.0
2.5
5.0
7.5
10.0
12.5
15.0
17.5
20.0
22.5
25.0
t
1
0
 
(
%
)
Fracture toughness (MPa m
0.5
)
 Reduction ratio 1
 Reduction ratio 2
and  4.33  the  Boscobel  granite,  which  has  the  lowest  fracture  toughness,  has  the  highest 
percentage passing at t
2
 and t
4
.  This indicates that multiple crack surfaces, occurring as 
grain boundaries, are propagated upon fracture.  However, at lower t
n
 values the Boscobel 
has  a  smaller  percentage  passing.    These  sizes  are  near  the  grain  size  of  the  granite  and 
the  progeny  fragments  at  these  sizes  were  most  likely  produced  at  first  fracture  or  soon 
there  after.    In  other  words,  large  grained  rocks  shatter  more  upon  initial  fracture 
producing  daughter  fragments  that  are  not  subjected  to  secondary  breakage.    In  order  to 
determine  a  relationship  between  fracture  toughness  and  breakage  function,  as 
represented  by  a  series  of  t
n
  values,  observation  of  the  fracture  behavior  and 
structure/grain size of each rock is required. 
 
A  final  observation  from  the  breakage  distribution  results  is  that  the  t
10
  parameter 
measured  under  jaw  crusher  conditions  can  be applied  in  order  to  determine  a  complete 
size distribution.  This is the one-parameter family of curves method described in section 
2.4.3,  where,  using  only  a  given  value  of  t
10
,  the  full  product  size  distribution  can  be 
reconstructed.    Figure  4.34  shows  the  HECT  breakage  results  in  terms  of  t
n
  versus  t
10
.  
The  solid  lines  represent  Narayanans  (1985)  work  presented  previously  in  Figure  2.21.  
Narayanans work was for tumbling mills and there seems to be some agreement, at least 
trend-wise,  between  that  work  and  the  results  from  this  study.    The  t
n
  values  from  the 
HECT jaw crusher set-up are higher than those given by Narayanan, most likely because 
of  the  different  size  reduction  application  (jaw  crusher  vs.  tumbling  mill),  different 
degrees of reduction ratio, and the fact that Narayanans data represents only one type of 
material,  where  as  the  data  in  this  study  covers  six  different  rock  types  (although 
Naraynans relationships have since been verified for a range of rock/ore types).  Despite 
these  differences  it  is  clear  that  t
10
  is  mostly  a  function  of  reduction  ratio,  and  not 
necessarily  of  energy  input  or  size  reduction  machine,  and  that  if  t
10
  and  fracture 
toughness can be related, then all other t
n
 values can be determined simply from t
10
.  
  115
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
0
10
20
30
40
50
60
70
80
90
100
 t
2
 t
4
 t
10
 t
25
 t
50
 t
75
 
Figure 4  ratio 1  .32 Relationship between various t
n
 and K
Ic
 for reduction
Ic
t
n
 
(
%
)
K  (MPa m
0.5
)
Figure 4.33 Relationship between various t
n
 and K
Ic
 for reduction ratio 2 
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
0
10
20
30
K
Ic
 (MPa m
0.5
)
40
50
60
70
80
90
100
t
n
 
(
%
)
 t
2
 t
4
 t
10
 t
25
 t
50
 t
75
  116
 
Figure 4.34 t
n 
vs. t
10 
overlapped with Narayanans one-parameter family of curves
0 10 20 30 40 50
0
20
40
80
100
60
t
n
 
(
%
 
p
a
s
s
i
n
g
)
 t
t
75
t
50
t
25
t
10
t
4
t
2 t
2
 t
4
 t
10
 t
25
50
 t
75
t
10
 (%)
  117
CHAPTER 5. MODEL DEVELOPMENT AND EXPERIMENTAL VERIFICATION 
 
 
5.1 MODELS FOR JAW CRUSHER POWER CONSUMPTION AND PRODUCT SIZE 
 
5.1.1 POWER CONSUMPTION 
 
The  results  presented  in  section  4.1  indicated  a  strong  correlation  between  fracture 
toughness,  K
Ic
,  and  specific  comminution  energy,  E
c
.    This  relationship  can  be  used  to 
develop an empirical model for the prediction of jaw crusher power consumption that can 
account for changes in reduction ratio. 
 
  As  was  noted  earlier  the  Culpeper  gray  and  red  siltstones  were  tested  separately  but 
when no statistical difference was found between their mechanical properties or fracture 
toughness values it was decided that they would be considered as one rock.  Therefore the 
HECT  results  for  the  gray  and  red  siltstone  have  been  pooled  together  and  model 
development is based only on data for six rock types. 
 
Figure  5.1  is  a  plot  of  specific  comminution  energy  versus  fracture  toughness  for 
reduction  ratio  1.    It  displays  a  best-fit  line  of  the  data  along  with  the  upper  and  lower 
95%  confidence  limits  of  the  correlation.    The  linear  fit  has  been  forced  through  the 
origin  based  on  the  assumption  that  a  material  with  no  fracture  toughness  would  not 
require  energy  in  order  to  be  fractured.    The  linear  relationship  between  E
c
  and  K
Ic
  is 
given by the following equation: 
    [5.1] 
The coefficient of determination for the relationship given in Equation 5.1 is 0.89.  When 
the reduction ratio is so changes.  This is 
en  in  Figure  5.2.    The  same  type  of  relationship  holds,  with  E
c
  and  K
Ic
  being  linearly 
rela
0.255   {for a reduction ratio of 1.50}
c Ic
E K =
 increased the relationship between E
c
 and K
Ic
 al
se
ted.  The relationship can be expressed by: 
  0.342   {for a reduction ratio of 2.97}
c Ic
E K =   [5.2] 
The  coefficient  of  determination  for  the  relationship  between  E
c
  and  K
Ic
  for  reduction 
ratio  2  is  0.80.    An  increase  in  reduction  ratio  results  in  an  increase  in  the  slope,  m,  of 
Equation  5.1.    In  order  to  determine  a  model  that  is  inclusive  of  all  reduction  ratios  an 
expression relating the slopes of Equations 5.1 and 5.2 needs to be determined. 
 
  In addition to the two reduction ratios used in this study, it can be assumed that at a 
reduction ratio of one (i.e, the particle size equals the closed side set and no strain energy 
is input into the particle), the slope of the relationship between E
c
 and K
Ic
 is also zero.  A 
plot  of  this  data  is  given  in  Figure  5.3  and  has  been  fit  with  two  expressions,  one  for 
reduction ratios between 1 and 1.5 and the other for reduction ratios greater than or equal 
to 1.5 .  The general relationship between E
c
 and K
Ic
, covering all reduction ratios (RR), is 
then given by the following equation: 
  118
  119
 
Figure 5.1 Linear fit of E
c
 and K
Ic
 data for reduction ratio 1 
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
0.00
0.05
0.10
0.15
0.20
E
K
Ic
 (MPa m
0.5
)
0.48
0.54
0.60
 Linear fit for reduction ratio 2
 Upper 95% Co fidence Limit n
 Lower 95% Confidence Limit
E
c
 
(
k
W
h
/
t
)
K
Ic
 (MPa m
0.5
)
c
 
(
k
0.25
0.30
0.35
0.40
0.45
0.50
 Linear fit for reduction ratio 1
 Upper 95% Confidence Limit
 Lower 95% Confidence Limit
W
h
/
t
)
Figure 5.2 Linear fit of E
c
and K
Ic
data for reduction ratio 2 
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
0.00
0.06
0.12
0.18
0.24
0.30
0.36
0.42
  120
 
|   |
0.428
0.511 0.511         for 1 1.5
0.215                 for  1.5
c Ic
c Ic
E RR K
E RR K RR
=     +      <
   ( =   
   
RR
  [5.3] 
 
An  allometric  equation  was  chosen  for  reduction  ratios  greater  than  or  equal  to  1.5.  
Although the measured data only suggest a linear relationship between reduction ratios of 
1.5  and  2.97  it  has  been  assumed  that  a  limiting  power  law,  or  possibly  exponential,  
relationship  between  specific  communution  energy  and  reduction  ratio  exists  at  higher 
reduction ratios.  This is mainly due to the fact that the rate of strain input decreases with 
increasing reduction ratio.   
 
Equation  5.3  governs  the  relationship  between  fracture  toughness  and  specific 
comminution energy based on the HECT results, where E
c
 is given in terms of kilowatt-
hours  per  metric  ton  and  reduction  ratio  is  defined  as  the  particle  size  divided  by  the 
closed  side  set.    In  order  to  determine  the  power  consumption  of  a  jaw  crusher  the  feed 
size distribition would have to be known, as well as the mass flow at each size, and the 
the  idling  power  of  the  crusher.    Equation  5.3  can  be  adjusted  to  account  for  these 
variables and an equation for determining the power consumption of a jaw crusher, based 
on fracture toughness, can be written as: 
Figure 5.3 Change in coefficient m with reduction ratio 
0 1 2 3 4 5 6 7 8 9 10
0.00
0.06
0.12
0.18
0.24
0.30
0.36
0.42
0.48
0.54
0.60
m = -0.511+0.511RR
m = 0.215RR
0.428
S
l
o
p
e
,
 
m
Reduction ratio, RR
|   |
1
0.428
1
0.511 0.511     for 1 1.5 
0.215             for  1.5
j
c i Ic i i n
i
j
c i Ic i i n
i
P RR K C x P
P RR K C x P RR
=
=
=      +   +      <
   (
=   +
   
RR
  [5.4] 
where, P
c
 is the power consumption of the crusher in kW 
    RR
i
 is the reduction ratio for a particle of size i 
    C
i
 is the probability of breakage for particle size i 
    x
i
 is the mass flow of particle size i in metric tons per hour 
    P
n
 is the power drawn by the crusher under no load in kW. 
Equation 5.4 accounts for the wide range of reduction ratios seen in a jaw crusher due to 
the  non-uniformity  of  the  feed  size  distribution.    It  is  difficult  in  practice  to  sample  the 
feed size entering a jaw crusher and often times the feed size distribution is unknown.  In 
these  instances  Equation  5.4  is  not  applicable.    If  an  estimate  of  the  average  reduction 
seen by the feed can be made, which infers some kowledge of the feed size distribution,  
then the following equation can be used to approximate the power consumption of a jaw 
crusher: 
 
RR
  [5.5] 
w
    T is the total mass flow in metric tons per hour. 
 
  Equation 5.4 is limited since the E
c
 and K
Ic
 relationship is based on six rocks and two 
reduction ratios.  Figures 5.1 and 5.2 indicate that the relationship may be exponential or 
allometric, with a limiting amount of specific comminution energy required to crush high 
fracture toughness rocks (i.e., a ceiling is reached in terms of E
c
).  The strength of the 
linear  relationships  suggested  in  Equations  5.1  and  5.2  is  influenced  by  the  behavior  of 
the  Culpeper  siltstone,  which  at  each  reduction  ratio  has  E
c
  values  outside  the  95% 
confidence limit of the linear fit model.  As discussed in section 4.2, the siltstone was not 
exposed to as much secondary breakage as the other rocks.  This is mainly due to the low 
elastic  modulus  of  the  siltstone.    Though  the  siltstone  was  the  only  tough  rock  that 
exhibited this type of behavior, it is believed that E
c
 also has a strong dependency on the 
elastic  modulus  and  over  the  course  of  more  testing,  linear,  or  non-linear,  relationships 
between  E
c
  and  K
Ic
  may  be  strengthened  if  they  are  defined  for  specific  intervals  of 
elastic modulus.  Another option may be to use the strain energy release rate, G, which is 
dependent on fracture toughness and the elastic modulus (and in the case of plane strain, 
Poissons ratio).  G, or an estimate of G based on the measured K
Ic
 and elastic modulus 
values for each rock, was investigated for this study but the correlation, and the strength 
of the relationship, between strain energy release rate and E
c
 was not as significant as the 
one between E
c
 and K
Ic
. 
 
0.428
0.511 0.511     for 1 1.5 
0.215              for  1.5
c avg Ic n
c avg Ic n
P RR K T P
P RR K T P RR
   ( =      +   +      <
   
   (
=   +
   
here, RR
avg
 is the avereage reduction ratio for the entire feed 
  121
5.1.2 PRODUCT SIZE 
 
A  direct  model  for  the  product  size  appearing  from  a  jaw  crusher  is  not  possible.    Any 
n, and the 
reakage  function.    The  model  developed  here  is  strictly  for  the  prediction  of  the 
onjunction with the feed size distribution and classification 
nction can be used to find the product size distribution from a jaw crusher. 
the  one-parameter  family  of  curves  approach 
opularized by the JKMRC (Julius Kruttschnitt Mineral Research Center).  Each method 
age functions (2 for each rock) was 
rduous. 
 
these  relationships.    Thus  the  one-parameter  family  of 
curves  can  be  used  to  determine  the  breakage  function  for  a  ro
ults  are  pooled  and  presented  as  one  rock
because the percent passing at a particle size one-tenth the parent si
e weaker 
cks  that  are  activated  upon  fracture  and  produce  progeny  fragments  that  are  not 
sly.  This process is shatter fracture and results in 
prediction of product size is dependent upon the feed size, classification functio
b
breakage function, which in c
fu
 
  In  section  2.4.3  methods  for  describing  the  breakage  function  were  covered.    The 
most  common  method  is  to  express  the  breakage  function  as  a  mixture  of  separate  size 
populations.    A  second  method  is 
p
was  given  due  consideration  for  the  breakage  data  collected  in  this  study  but  the  final 
model  used  here  employs  an  approach  similar  to  that  of  the  one  parameter  family  of 
curves.  For the mixed population approach, relationships between the coefficients K
i 
and 
n
i
,  used  to  describe  different  size  populations,  and  fracture  toughness  were  difficult  to 
discern and curve fitting of the twelve measured break
a
  The model developed here is based on the characteristic size reduction parameter t
10
.  
In  section  4.2.3  a  plot,  inclusive  of  all  reduction  ratios,  of  t
n
  values  versus  t
10
  was 
presented  that  showed  relationships  exist  between  different  t
n
  values  and  t
10
.    It  was 
determined that if t
10 
is a function of fracture toughness and reduction ratio, then other t
n
 
values  can  be  determined  using 
ck  given  the  fracture 
toughness and reduction ratio. 
 
  Figure  4.31  in  section  4.2.3  showed  the  relationship  between  t
10
  and  fracture 
toughness at each reduction ratio.  A correlation between the two is hard to observe, even 
when  the  siltstone  res .    This  is  most  likely 
ze, or any size for that 
matter, is dependent upon the reduction ratio, the rocks fracture behavior, and the rocks 
grain size/structure.  The data presented in section 2.4.3 indicated that at small t
n
 values 
the percent passing increases with increasing fracture toughness for the smaller reduction 
ratio.  One explanation for this behavior is the increased presence of flaws in th
ro
subjected  to  secondary  breakage.    This  corresponds  with  the  behavior  observed  at  the 
higher  reduction  ratio  that  showed  a  decrease  in  percent  passing  with  increased  fracture 
toughness  at  the  small  t
n
  values.    The  effect  of  grain  size/structure  is  nullified  when 
significant amounts of secondary breakage occur.  A second explanation is based on the 
ability of the rock particle to absorb energy.  It is possible that before fracture occurs at 
the  weakest  flaw  in  a  rock  particle,  excess  energy  can  be  applied  such  that  fracture  of 
many flaws occurs almost simultaneou
  122
finer progeny fragments.  Again, at higher reduction ratios this behavior is hidden due 
large amounts of secondary breakage. 
 
  Based  on  the  previous  qualitative  description,  the  effect  of  fracture  behavior  and 
microstructure  on  the  breakage  distribution  should  be  accounted  for  when  investigating 
the  relationship  between  t
10
  and  K
Ic
.    The  only  approach  that  can  be  taken  here  is  to 
account  for  a  rock  particles  ability  to  absorb  energy  by  estimating  the  strain  energy 
release rate, G, of each rock.  Recalling Equation 2.34, G can be calculated using fracture 
toughness, elastic modulus, and Poissons ratio.  For conditions of plane stress, only  K
Ic
 
and  the  elastic  modulus  are  needed.    Poissons  ratio  for  the  rocks  used  in  this  study  are 
not known thus a universal value of 0.3 was assumed since Poissons ratios for rocks are 
fairly small and usually very comparable.  Table 5.1 lists the G values for the rocks used 
in this study, along with their K
Ic
 values for reference. 
 
Table 5.1 Strain energy release rate of each rock 
G  K
Ic
 
Rock Type 
2
J m   MPa m  
Shadwell Metabasalt  60.15  1.411 
Boscobel Granite  15.98  0.601 
Culpeper Siltstone  79.54  1.477 
Lees . burg Diabase  47 95  1.254 
Thornburg Granite  36.36  1.045 
Spotsylvania Granite  25.22  0.843 
 
  Figures 5.4 and 5.5 are plots of t
10
 versus  G for each reduction ratio.  An allometric 
expression has been fit to the data in each case.  For a reduction ratio of 1.50 t
10 
can be 
determined from: 
 
0.251
10
3.54
'
Ic
K
t
E
|   |
=
     |
\   .
  [5.6] 
where E is the elastic modulus.  For a reduction ratio of 2.97 the relationship between t
10
 
and K
Ic
 can be written as: 
 
0.139
10
27.2
'
Ic
K
t
E
|   |
=
     |
\   .
  [5.7] 
 
It  is  clear  that  the  relationship  between  t
10
  and  K
Ic
2
/E  is  not  well  defined  but  the 
coefficient  of  determination  is  higher  than  that  for  any  relationship  defined  between  t
10
 
and  K
Ic
.    The  relationship  shown  in  Figure  5.4  demonstrates  that  t
10
  increases  with  G.  
This  is  because  for  rocks  with  higher  G  values  shatter  fracture  is  more  prevalent  and 
more  energy  is  released  as  new  crack  surface  energy.    Figure  5.5  indicates  that  as  the 
reduction ratio is increased, the relationship between t
10
 and G inverses, and t
10
 decreases 
with increasing strain energy release rate.  The low G rocks are subjected to more 
  123
 
Fi
10
1  gure 5.4 Allometric fit of t  versus G for reduction ratio 
0 10 20 30 40 50 60 70 80 90 100
0.0
2.5
5.0
7.5
10.0
12.5
15.0
17.5
20.0
22.5
25.0
t
1
0
 
(
%
 
p
a
s
s
i
n
g
)
G (J/m
2
)
 Allometric fit for reduction ratio 1
Figure 5.5 Allometric fit of t
10
 versus G for reduction ratio 2 
0 10 20 30 40 50 60 70 80 90 100
22.5
25.0
12.5
15.0
17.5
20.0
 
p
a
s
s
i
n
g
)
0.0
2.5
5.0
7.5
10.0
t
1
0
 Allometric fit for reduction ratio 2
 
(
%
G (J/m
2
)
  124
secondary  breakage  than  the  higher  G  rocks  because  the  low  G  rocks  initially  produce 
large  progeny  fragments  as  a  result  of  cleavage.    These  progeny  undergo  additional 
breakage in comparison to the first fragments of the high G rocks, prod
and having more particles of varying size. 
 
7 are functions of the reduction 
expressions  relating  them  to  the 
  a,  is  plotted  against  reduction 
  1  means  that  no  size 
d and all t
n
s are zero.  
e  functions  used  to  relate  the  coefficient  a  and  the  reduction  ratio.    Two 
expressions were used, one for reduction ratios between 1 and 1.5, and one for reduction 
ratios greater than or equal to 1.5. 
The exponent, defined as b, in Equations 5.6 and 5.7 is more difficult to relate to the 
reduction  ratio.    In  this  case  b  decreases  and  becomes  negative  between  the  reduction 
uced by shatter 
The coefficients and exponents in Equations 5.6 and 5.
ratio.    Thus,  as  was  done  for  the  E
c 
model,  general 
reduction  ratio  need  to  be  determined.    The  coefficient,
ratio  in  Figure  5.6.    As  was  noted  previously  a  reduction  ratio  of
reduction takes place, therefore no daughter fragments are produce
Again,  since  no  reduction  ratios  greater  than  3  were  used  some  assumptions  have  to  be 
made about what occurs as the reduction ratio increases.  In this instance it is known that 
t
10
 cannot exceed 100%, thus a has  an  upper  limit  as  the  reduction  ratio  increases.    It  is 
expected that a will increase at about the same rate as defined by the two measured data 
points  up  to  a  reduction  ratios  of  about  5.    Above  a  reduction  ratio  of  5,  a  will  not 
increase as rapidly and will approach 100 at a reduction ratios greater than 10.  Figure 5.6 
shows  th
 
Figure 5.6 Change in coefficient a with reduction ratio 
0 1 2 3 4 5 6 7 8 9 10
0
10
20
30
40
50
60
70
80
a = 100 - 100/(1+(RR/3.97)
3.40
)
a = -7.095 + 7.905RR
C
o
e
f
f
i
c
i
e
n
t
,
 
a
Reduction ratio
90
100
  125
ratios of 1.5 and 2.97.  But for a reduction ratio of one b is zero, as is a, and at some large 
duction ratio b equals zero again (in order to satisfy the t
10
 upper limit of 100%).  For  re
reduction ratios between 1 and 1.5 the change in b is assumed linear.  For reduction ratios 
greater  than  1.5  b  decreases  in  a  exponential/linear  manner  up  to  a  reduction  ratio  of  3.  
Beyond  a  reduction  ratio  of  3  b  increases  and  approaches  zero  in  an  exponential/linear 
manner.  The functions governing the change in b with reduction are given in Figure 5.7. 
0.5
Figure 5.7 Change in exponent b with reduction ratio 
3 4 5 6 7 8 9 1
-0.1
b = -0.190 + 0.0127RR + 13.4RR(0.0998)
RR
Reduction ratio
0.0
0.1
0.2
0.3
0.4
b = -0.501 + 0.501RR
 
E
x
p
o
n
e
n
t
,
 
b
0 1 2
-0.5
-0.4
-0.3
-0.2
0
 
Substituting  the  functions  that  relate  the  coefficient  a  and  the  exponent  b  to  the 
reduction  ratio  into  Equations  5.6  or  5.7  gives  the  following  general  expression  relating 
t
10
 and K
Ic
2
/E: 
|   |
  (   )
(   )
(   )
0.501 0.501
2
10
0.190 0.0127 13.4 0.0998
2
10 3.40
7.01 7.01                                for 1 <1.5
'
100
100        for  1.5
'
1
3.97
RR
RR
Ic
RR
Ic
K
t RR RR
E
K
t RR
E
RR
   +
   +   +
|   |
=     +   
   |
\   .
   (
   (
|   |
   (
=      
   |
   (
   (
   
  [5.8] 
where,  RR  is  the  reduction  ratio.    Equation  5.8  is  applicable  only  over  the  range  of 
fracture toughness values determined in this study.  What occurs outside of this range, at 
K
Ic
2
/E values approaching zero or values greater than 80 J/m
2
, is hard to assess based on 
|   |
  \   .
+
   |
\   .
the  given  data.    More  testing  is  required  in  order  to  establish  a  wider  ranging  t
10
  and 
  126
K
Ic
2
/E relationship, as well as to strengthen or modify the suggested one.  Additionally, 
HECT tests run at higher reduction ratios need to be performed in order to verify that the 
nctions  proposed  in  Figures  5.6  and  5.7  are  acceptable.    Despite  these  limitations  it  is 
age  function  more  t
n
  values  are 
eeded.  Using the one-parameter family of curves approach t
10
 can be related to other t
n
 
values.  Figure 4.3 g of that data was 
used  to  develop  Figures  5.8  and  5.9,  whi show  the  relationship etween  various  t
n
 
values  and  t
10
tionships  are  also  given  in  Figures 
5.8 and 
    [5.9] 
The series of equations given in 5.9 can be used along
complete  breakage  function  for  a  rock  of  known  fracture  toughness  exposed  to  a  given 
e  data  and  a  classification  function  in  order  to  predict  the 
product  size  distribution  produced  by  a  jaw  crusher  when  crushing  the  given  rock 
materia  
 
fu
believed  t
10
  is  dependent  upon  the  fracture  behavior  of  the  rock  as  defined  by  either 
fracture toughness or the strain energy release rate, which is related to K
Ic
.  The fracture 
toughness of a rock is dependent upon the presence, distribution, and geometry of flaws 
within the rock, as well as the energy/stress applied to the rock.  These factors also dictate 
the size and distribution of new crack surfaces present after fracture or breakage and are 
represented  as  t
n
  values.    Further  study  will  enhance  the  inter-relationship  between 
fracture toughness and fractured size distribution. 
 
  Equation  5.8  defines  t
10
  for  a  given  reduction  ratio,  fracture  toughness,  and  elastic 
modulus.    In  order  to  determine  the  complete  break
n
4 showed the measured t
n
 versus t
10
 data.  Curve fittin
ch  s  b
.    The  expressions  describing  those  rela
5.9 and are: 
0
7.68
10
t
40.9
2.17
0.281
1.5 10
0.800
2 1
4 10
10
25 10
75 10
100 10
0.556
0.271
t t
t t
t t
t
t t
t t
=
=
=
=
=
=
50 10
0.384
0.315
t t
t t
=
=
 with Equation 5.8 to determine the 
level  of  size  reduction  in  a  jaw  crusher.    The  breakage  function  can  then  be  used  in 
conjunction  with  feed  siz
l.
 
  127
Figure 5.9 Plot of functions relating small t
n
values and t
10
 
0 2 4 6 8 10 12 14 16 18 20
0
1
2
3
4
5
6
7
8
9
10
t
50
 = 0.384t
10
t
75
 = 0.315t
10
t
100
 = 0.271t
10
t
100
t
75
t
50
t
n
 
(
%
)
t
10
 (%)
Figure 5.8 Plot of functions relating large t
n 
values and 
10
0 2 4 6 8 10 12 14 16 18 20
0
10
20
30
40
50
60
70
80
90
100
t
1.5
 = 40.94(t
10
)
0.281
t
2
 = 7.68(t
10
)
0.800
t
4
 = 2.174t
10
t
10
 = t
10
t
25
 = 0.556t
10
t
10
t
25
t
4
t
2
t
1.5
t
n
 
(
%
)
t  (%)
  128
 
5.2 EXPERIMENTAL VERIFICATION OF MODELS 
 
5.2.1 LABORATORY  VERIFICATION OF  E
C
  AND BREAKAGE FUNCTION MODELS 
 
It was concluded that a brief assessment  of  the  HECT  derived E
c 
and breakage function 
models  needed  to  be  performed  before  laboratory  jaw  crushing  was  conducted.    This 
allows  for  an  investigation  of  the  strength  of  the  models  strictly  from  a  laboratory,  or 
baseline, perspective before scale-up variables can have an effect on the predicted results 
(i.e,  feed  rate,  jaw  crusher  geometry).    Samples  of  dolomitic  limestone  where  tested 
because  it  had  been  used  in  the  development  of  the  Edge  Notched  Disk  fracture 
toughness  test  and  a  K
Ic
  value  was  available.    The  fracture  toughness  of  the  dolomitic 
limestone is 1.397 MPa m and it has an elastic modulus of 39.7 GPa.  
 
  Ten samples of the dolomitic limestone were tested in the High Energy Crushing Test 
system.  Except for the closed side set, which was 35.2 mm, the conditions of the HECT 
system were exactly the same as used for the six Luck Stone rocks.  The 10 samples were 
disk specimens with an average diameter of 50.22 mm and thickness of 24.82 mm.  The 
average  reduction  ratio  was  1.43.    The  average  specific  comminution  energy  of  the  ten 
samples was determined and the pooled fragments sieved for breakage function analysis.  
The specific comminution energy results are shown in table 5.2 along with the predicted 
value of E
c 
as given by Equation 5.3.  The measured t
10
 value and the t
10
 value predicted 
by  Equation  5.9  are  also  in  Table  5.2.  The  actual  breakage  function  and  the  predicted 
breakage function given by Equation 5.9 are shown in Figure 5.10. 
 
Table 5.2 Actual and predicted E
c 
and t
10 
for dolomitic limestone 
Parameter  Measured  Predicted 
E
c
  0.295 kWh/t  0.305 kWh/t 
t
10
  6.0 %  6.6 % 
 
  Table  5.2  indicates  that  the  E
c
  model  developed  from  the  experimental  program  is 
fairly  accurate  for  the  given  reduction  ratio  of  1.43.    Since  this  reduction  is  within  the 
bounds of the measured data it is not surprising the actual and predicted E
c
 values are in 
good  agreement.    It  remains  to  be  seen  if  the  model  holds  at  higher  reduction  ratios, 
which  could  not  be  tested  here  due  to  the  limited  availability  of  rocks  and  the  re-
configuration  of  the  HECT  system  required  for  small  closed  side  sets  and  intense 
secondary crushing. 
 
  The  measured  and  predicted  breakage  functions  also  show  fairly  good  agreement, 
particularly  at  the  larger  progeny  sizes.    As  the  daughter  size  decreases  past  0.1  (which 
represents  t
10
)  t ge  distribution 
increases.  This is because the relationships between t
10
 and the smaller t
n
 values (t
50,  75, 
100
)
he  deviation  between  the  measured  and  predicted  breaka
 given in Equation 5.9 are not as strong as those between t
10
 and the larger t
n
 values.  
  129
The average coefficient of determination for the relationships between t
10
 and t
n
s coarser 
than t
10
 is 0.85, and between t
10
 and the finer t
n
s it is 0.60.  
1
 Measured
 Predicted
 
5.2.2 PREDICTIO L  
 
A  l
ure  toughness  based  model  for  the  prediction  of  capacity.    The  crushing 
tests  were  conducted  on  rock  samples  from  Boscobel,  Culpeper,  Shadwell,  and 
Spotsylvania  in  accordance  with  the  procedure  outlined  in  section  3.5.    Appendix  IV 
contains data collected from the laboratory crushing trials. 
 
Verification of breakage function (product size) model 
The  closed  side  set  and  open  side  set  of  the  crusher  were  7.94  mm  and  14.29  mm 
respectively.    The  throw  of  the  crusher  was  6.35  mm  and  the  crusher  speed  was 
approximately  300  rpm.    Based  on  the  feed  size  distributions  shown  in  Figure  3.15  the 
Figure 5.10 Measured and predicted breakage function for dolomitic limestone
0.01 0.1 1
0.01
0.1
B
 
(
d
i
,
d
1
)
d
i
/d
1
N OF PRODUCT SIZE USING A  ABORATORY JAW CRUSHER
aboratory  scale  crusher  was  used  to  test  the  ability  of  the  breakage  function  model, 
when  used  with  the  feed  size  distribution  and  a  classification  function,  to  predict  the 
product size coming out of an actual jaw crusher.  The lab-scale crusher was also used to 
develop  a  fract
maximum  reduction  ratio  seen  by  any  particle  in  the  crusher  was  around  4.    Using 
Whitens crusher model, determined from mass balance equations, and given in Equation 
2.4,  the  product  size  was  predicted  using  the  breakage  function  model  given  in  the 
previous  section  along  with  a  classification  function  and  the  feed  size  distribution.    The 
  130
breakage  function  for  each  particle  size  in  the  feed  was  determined  using  Equations  5.8 
and 5.9 where reduction ratio was defined as the particle size divided by the closed side 
set.  The classification function was calculated for each feed particle using Equation 2.5.  
A  MathCAD  program  was  developed  in  order  to  perform  matrix  calculation  of  the 
product  size  using  the  feed  size  distribution  (mass  fraction  in  each  size)  vector,  the 
diagonal classification function matrix, and the strictly lower triangular breakage function 
matrix.  The product size distribution vector was determined in terms of the mass fraction 
retained  in  each  size  class.    This  vector  in  turn  was  converted  to  a  percent  passing 
distribution in order to compare it to the actual, measured size distribution of the product. 
 
  The  actual  product  size  distribution  from  the  jaw  was  determined  by  sampling  and 
f the cylindrical disk specimen during HECT testing.  Since most of 
the feed material was in the 19.1 to 25.4 mm size range, the breakage function developed 
in this study using the disk specimen implies that a large portion, if not all, of the product 
will  pass  at  sizes  half  of  19.1  to  25.4,  or  9.5  to  12.7  mm.    In  each  case  100%  of  the 
product was predicted to pass at 12.7 mm.  If irregular particles were used during single 
particle  breakage  analysis  more  agreement  would  be  expected  between  the  actual  and 
predicted  distributions  at  the  larger  sizes.    The  
distributions  are  very  similar  in  each  case  with  the  major  difference  between  the  two 
curves  being  in  the  proportion  of  material  in  each  size  class.    The  Boscobel  granite 
product distributions showed the most deviation, with about a constant 5% difference (by 
% passing) between the actual and predicted percentage passing.   
 
  are  minimal.    The  rocks  crushed  in  the  lab  jaw  were  also  the  same  as  ones 
used  in  the  development  of  the  breakage  function  model,  thus  the  applicability  of  the 
odel to rocks outside the measured data range is still in question.  In the same vein, the 
maximum  reduction  seen  in  the  lab  crushing  experiments  was  about  4,  which  is  outside 
the  range  used  in  model  development  but  not  too  far  (maximum  HECT  reduction  ratio 
was  3).    Since  jaw  crushers  can  see  higher  reduction  ratios  up  the  applicability  of  the 
breakage function to all jaw crushing reduction ratios remains to be tested.  Despite these 
limitations the fracture toughness based breakage function model can be used to provide 
a reasonable estimate of the product size coming out of a jaw crusher given the feed size 
distribution and the closed side set of the crusher. 
sieve analysis.  Approximately 9 kg of product material was sampled, sieved using a Ro-
Tap,  and  weighed  for  each  rock  tested.    The  actual  and  predicted  results  are  shown  in 
Figures 5.11 through 5.14. 
 
  It  can  be  seen  from  Figures  5.11  -  5.14  that  the  actual  and  predicted  results  are  in 
fairly  good  agreement.    The  predicted  product  size  distribution  deviates  at  the  12.7  mm 
size due to the use o
shapes  of  the  predicted  and  actual 
  Based on the results of the laboratory jaw crushing experiment, the suggested fracture 
toughness based model for the determination of the breakage function seems reasonable.  
However,  lab  crushing  is  a  small-scale  attempt  at  reproducing  very  large-scale 
conditions.    The  feed  size  used  may  not  be  representative  of  what  is  seen  in  actual 
crushing  conditions  and  the  effects  of  jaw  plate  wear,  throughput,  and  other  operating 
conditions
m
  131
  132
Figure 5.11 Pr i vania granite edicted and actual product s ze distributions for Spotsyl
1 10 30
0
10
20
30
40
50
60
70
80
90
100
%
 
p
a
s
s
i
n
g
Particle size (mm)
Spotsylvania granite
 Actual
 Predicted
Figure 5.12 P l metabasalt redicted and actual product size distributions for Shadwel
1 10 30
0
10
20
30
40
50
60
70
80
90
100
%
 
p
a
s
s
i
n
g
Particle size (mm)
Metabasalt
 Actual
 Predicted
 
Figure 5.13 Predicted and actual product size distributions for Culpeper siltstone 
1 10 30
0
10
20
30
Particle size (mm)
100
100
90  Actual
 Predicted
Siltstone
40
50
60
70
80
%
 
p
a
s
s
i
n
g
Figure 5.14 Predicted and actual product size distributions for Boscobel granite
90  Actual
 Predicted
 
1 10 30
0
10
Particle size (mm)
Boscobel granite
20
70
80
30
40
50
60
%
 
p
a
s
s
i
n
g
  133
5.3 MODEL FOR VOLUMETRIC CAPACITY OF A JAW CRUSHER 
 
In  section  2.1.3  equations  were  given  for  the  volumetric  capacity  of  a  jaw  crusher 
based  on  the  dimensions  of  the  crusher,  its  operational  settings  (closed  side  set  and 
throw), its speed, the average feed size, and an undefined parameter related to the nature 
of  the  material.    The  laboratory  jaw  crusher  trials  were  used  to  determine  the  actual 
volumetric  capacity  of  the  jaw,  for  a  given  rock,  with  the  expectation  that  the  K
3
 
parameter could be defined in terms of fracture toughness. 
 
  For each rock the average particle size and bulk density of the feed were determined, 
and the time required to crush the feed material was recorded.  The volumetric capacity 
for each rock was then calculated.  Table 5.3 lists the measured capacities for each rock. 
 
Table 5.3 Rock specific volumetric capacities of laboratory jaw crusher 
Bulk density  Mass  Volume  Time  Vol. capacity 
Rock Type 
kg/m
3
  kg  m
3
  sec  m
3
/hr 
Culpeper siltstone  1702  38.80  0.0228  160  0.513 
Shadwell metabasalt  1837  45.97  0.0250  195  0.462 
Spotsylvania granite  1691  42.97  0.0254  169  0.541 
Boscobel granite  1670  47.14  0.0282  172  0.591 
 
  Equation 2.12 gives the volumetric capacity of a jaw crusher.  The predicted capacity 
is  the  ideal  capacity  (i.e.,  strictly  a  function  of  the  crusher  geometry  and  operational 
set ed 
ze, throw and gape, and the nature of the material, respectively.  Since K
3
 is undefined it 
ctual  capacity 
easured  using  the  lab  crusher  is  a  function  of  K
3
.    This  can  be  expressed  as  (Sastri, 
tings) normalized by three parameters, K
1
, K
2
, and K
3
, which are functions of the fe
si
is  assumed  that  the  deviation  between  the  predicted  capacity  and  the  a
m
1994): 
 
3
actual
predicted
V
K
V
=   [5.10] 
where (recalling Equations 2.12 and 2.13), 
(   )
1 2
2
450
60 0.5
predicted
g
V Nw CSS T K K
N
|   |
   |
 
1
0.85
avg
K
     |
2.5
F
  .
|   |
6.5
2
1.92 10
T
G
G
K
  
\   .
=   
and,  N is the speed of the crusher in rpm 
    w is the width of the jaws in m 
    CSS is the closed side set in m 
    T is the throw in m 
=   +
\
=      [5.11] 
  134
    D is the vertical depth between the jaws in m 
    G is the crusher gape in m 
g is gravitational acceleration in m/s
2
. 
Equation  5.10  was  used  to  determine  the  predicted  volumetric  capacity  for  each  rock, 
which  in  turn  was  compared  to  the  actual  volumetric  capacity  using  Equation  5.11.    It 
was anticipated that K
3
 would be a function of either fracture toughness or strain energy 
release  rate.    Figure  5.15  is  plot  of  K
3
  versus  fracture  toughness.    It  can  be  seen  that  K
3
 
decreases  with  fracture  toughness,  or  in  other  words,  the  volumetric  capacity  decreases 
1.0
with increasing fracture toughness.  K
3
 decreases from a value of one since it is expected 
that  the  capacity  of  a  material  with  little  to  no  hardness  is  strictly  a  function  of  the 
crusher  geometry,  operational  settings,  and  feed  size.    Again,  the  Culpeper  siltstone 
shows unique behavior in that it has the highest fracture toughness yet its value of K
3
, and 
in  turn  the  volumetric  capacity,  was  not  the  lowest  among  the  four  rocks  crushed.    The 
upward trend is a result of the large amount of shatter occurring in the siltstone particles 
upon fracture.  This behavior has been discussed in detail in the previous sections.  The 
increased shatter occurring in the siltstone results in a uniform progeny size distribution 
consisting  of  more  finely  sized  fragments.    In  regards  to  volumetric  capacity,  these 
fragments  pass  through  the  crusher  without  experiencing  much  secondary  breakage  and 
exit  the  crushing  chambering  rather  quickly.    Additionally,  the  lack  of  large  fragments 
that  result  primarily  from  cleavage  enables  more  feed  material  to  enter  the  crushing 
chamber.  Based on this observation K
3
 was plotte
Figure 5.15 K
3
 parameter versus fracture toughness 
Ic
d against the strain energy release rate, 
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
0.0
0.1
0.2
0.3
0.4
0.5
0.6
K
 
3
1.8 2.0
0.7
0.8
0.9
K  (MPa M
0.5
)
  135
Figure 5.16 K
3
 parameter versus strain energy release rate 
0 10 20 30 40 50 60 70 80 90 100
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
K
 
3
G (J/m
2
)
  136
which  was  previously  related  to  the  product  size  and  takes  into  account  the  dominant 
type of fracturing seen in a rock (i.e, cleavage, shatter).  Figure 5.16 shows K
3
 versus G 
for the four rocks crushed. 
 
Figure 5.16 indicates that the volumetric capacity decreases with an increase in strain 
energy release rate to a limiting value (which corresponds to K
3
 equal to 0.54) and then 
begins  to  increase  as  the  strain  energy  release  rate  increases  past  50  J/m
2
.    The 
relationship  between  K
3 
and  G  (where  G  equals  K
Ic
2
/E)  can  be  fit  with  an  exponential 
expression of the following form: 
  (   )
2 2
'
3
0.0987 0.0055 0.8995 0.9674
'
Ic
K
Ic E
K
K
E
|   |
=   +   +
   |
\   .
  [5.12] 
The volumetric capacity of a jaw crusher can be determined by substituting Equation 5.12 
into Equation 2.12. 
 
  The  relationship  given  in  5.12  is  limited  by  the  small  number  of  rock  types  (4) 
crushed  in  the  laboratory  jaw  crusher.    Verification  of  the  fracture  toughness  based  K
3
 
and volumetric capacity models were not performed.  Additional testing on the laboratory 
scale, or even on a large scale using actual jaw crushers, needs to be performed in order 
to assess the applicability of the given model.   
CHAPTER 6. CONCLUSIONS AND RECOMMENDATIONS 
 
 
6.1 CONCLUSIONS 
 
In  todays  crushing  environment  the  need  for  optimization  of  jaw  crusher  selection  and 
operation  to  meet  the  economic  requirements  of  quarrying  has  become  imperative.  
Aggregate  processing  plants  require  rugged,  massive,  and  expensive  equipment,  and 
misjudgments  in  design  are  difficult  and  expensive  to  correct.    The  selection  of  a  jaw 
crusher for application in the aggregate industry is primarily based on technical literature 
provided by crusher manufacturers, experience, and cost.  Charts and graphs provide data 
on  electric  power  requirements,  crusher  size,  as  well  as  expected  capacities  for  a  given 
material  and  closed  side  set.    These  charts  and  graphs  have  also  been  incorporated  into 
computer  programs  to  aid  in  crusher  selection.    Jaw  crusher  selection  is  also  heavily 
influenced  by  the  subjective  judgment/experience  of  individuals,  which  results  in  the 
conservative selection and operation of jaw crushers. 
 
In order to account for material variations manufacturers rely on a suite of laboratory 
tests.    The  uniaxial  compression  and  Bond  impact  crushability  test  are  two  of  the  most 
common tests used to rank materials relative to their hardness/crushability.  Each of these 
tests has limitations and neither adequa aterials resistance to fracture o
e breakage process that occurs in a jaw crusher.  The plates of a jaw crusher are used to
tely describes a m r 
th  
apply  compressive  forces  that  induce  tensile  stresses  within  particles,  causing  fracture.  
Any description of the nature of rock material, relative to jaw crushing, should be based 
on, at the very least, tensile strength.  Although jaw crushers are used extensively the lack 
of  understanding  relative  to  their  operational  characteristics,  as  well  as  a  reliance  on  an 
inappropriate,  single  set  of  material  properties,  makes  selection  of  a  proper  machine 
difficult, and optimization of in-place crushers almost non-existent. 
 
The focus of jaw crusher selection should be on the prediction of power consumption, 
product  size,  and  capacity.    Each  of  these  parameters  is  a  function  of  the  mechanical 
aspects  of  the  jaw  crusher  as  well  as  the  rock  material  being  crushed,  and  prediction  of 
these parameters requires the proper characterization of the mechanical nature of rock in 
regards to the jaw crushing environment and operational settings.  This will allow for the 
rapid estimation of power consumption, product size, and capacity for a given rock being 
crushed under a given set of operating conditions.  
 
Consideration  of  the  physics  of  particle  fracture  and  the  mechanisms  that  drive 
fracture  initiation  is  necessary  in  the  evaluation  and  improvement  of  jaw  crushing 
operations.    Rocks  broken  in  jaw  crushers  fail  at  stress  levels  well  below  what  is 
predicted based on stress-strain behavior or compressive strength, due to induced tensile 
stresses (rocks are much weaker in tension than compression) and the presence of cracks.  
Rocks by their nature contain inherent flaws in the form of grain boundaries, voids/pores, 
be ss  dding planes, foliations, etc., on both small and large scales.  These cracks act as stre
  137
concentrators, and the increase in stress at these locations is proportional to the geometry 
f the crack. 
 
.  The importance in terms of 
w  crushing  operations  is  that  fracture  is  dependent  on  applied  loads  (i.e.,  energy)  and 
The mechanisms of fracture also control the distribution of progeny particle sizes and 
specifi sics is 
the  foundation  for  describing  size  reduction  processes.    Energy-size  reduction 
relationships are related to Griffiths energy criterion and the presence of Griff cks. 
The  energy  criterion  states  that  enough  potential  energy  must  be  released  in  order  to 
overcome a materials resistance to crack propagation, requiring an increase in the work 
d rces  a   on  the  al.    Th e  amou f  energy   into 
re  a par  Therefo  energy ed for 
of  size  reduction  achieved,  or  the  size  distributi lting  f oth 
ependent upon the presence and distribution of Griffith flaws. 
  of  rocks,  and  establishing  a  relationship  between  flaw  geometry  and  fracture 
strength  is  the  most  fundamental  aspect  of  fract
defined  by  fracture  toughness,  the  critical  level  a
terizes  the  crack  opening  displacements  that 
crusher.    Induced  tensile  stresses  open  pre-
existing  flaws  and  ultimately  lead  to  crack  propagation  and  particle  failure.    When 
considering  jaw  crushing,  mode  I  fracture  toughness  is  the  material  property  most 
representative  of  the  breakage  process  occurring  due  to  point  contact  loading  and  it  is 
dependent  upon  the  same  parameters  (applied  stress/energy  inpu
geometry) that control the power consumption and product size of jaw crushers. 
 
In order to relate rock fracture toughness to the power consumption, product size, and 
etric  capacity  of  jaw  crushers,  a  series  of  laboratory  experiments,  comprised  of 
le breakage, and small scale jaw crushing, had to 
re toughness test had to be developed and verified 
tive, estimation of rock fracture toughness could be achieved.  
o
  Griffiths  theory  expounds  a  relationship  between  crack  size  and  fracture  stress.  
Griffith hypothesized that fracture occurs when the energy supplied by an external force, 
or  by  the  release  of  stored  strain  energy,  is  greater  than  the  energy  of  the  new  crack 
surface.  Griffiths work defined a constant critical value that when equal to the product 
of the applied stress and crack length will result in fracture
ja
the  presence,  distribution,  and  size  of  flaws,  and  both  need  to  be  accounted  for  when 
characterizing a rock materials resistance to fracture.   
 
c fracture mechanisms produce specific fragment sizes, and thus particle phy
ith cra
one  by  external  fo cting materi is  is  th nt  o   input
ducing the size of ticle.  re the  requir fracture and the amount 
on  resu rom  fracture,  are  b
d
 
Modifications to Griffiths theory have led to the development of the field of fracture 
mechanics.    Fracture  mechanics  provides  quantitative  methods  for  characterizing  the 
behavior of an intact material as it fractures due to crack growth.  Fracture mechanics is 
applicable  to  problems  of  rock  breakage  since  the  presence  of  inherent  flaws  and 
discontinuities  within  geologic  materials  result  in  failure.    These  flaws  control  the 
fracture
ure  mechanics.    This  relationship  is 
bove  which  crack  extension  occurs.  
Mode  I  fracture  toughness  properly  charac
occur  in  particles  being  broken  in  a  jaw 
t,  flaw  distribution  and 
volum
fracture toughness testing, single partic
be undertaken.  In addition a new fractu
so that rapid, and representa
  138
A major drawback in studying size reduction  processes from a fracture mechanics point 
 complexity of rock fracture toughness testing. 
 
comparison.    When  compared  to  the  Semi-Circular  bend  test,  a  more  common  and 
accepted fracture toughness test, the END test yields almost identical results. 
 
  Seven rocks, provided by aggregate producer Luck Stone Corporation, were tested for 
fracture  toughness,  specific  comminution  energy,  and  breakage  function.    Specific 
comminution  energy  and  breakage  function  were  determined  using  a  unique  single 
particle  breakage  device,  the  High  Energy  Crushing  Test  (HECT)  system.    The  HECT 
system  is  a  specially  configured  materials  testing  system  that  is  capable  of  simulating  a 
wide  range  of  crushing  operations.    For  this  study  the  HECT  system  was  configured  to 
deliver  a  sinusoidal  crushing  blow  at  a  speed  of  225  rpm,  average  jaw  crushing 
conditions.    Since  power  consumption  and  product  size  are  functions  of  the  crusher 
settings  as  well  as  the  material  being  crushed,  the  closed  side  set  of  the  HECT  system 
was varied, resulting in single particle breakage tests run at two reduction ratios, 1.5 and 
3.  Core based, disk specimens were employed in the HECT system in order to eliminate 
the  effect  of  microstructure  changes  due  to  blasting  (since  the  END  test  also  used  rock 
core).    Although  some  size/geometry  dependency  was  seen  in  the  breakage  function 
results,  the  experimental  program  performed  in  this  study  provides  a  unique  core  based 
approach that can be used to determine the jaw crushing requirements of a given rock. 
 
  The  results  of  the  fracture  toughness  and  HECT  system  tests  indicate  a  strong, 
proportional  relationship  between  fracture  toughness  and  specific  comminution  energy.  
Additionally, fra inution energy 
more strongly th trength.  For the 
the road in terms of required breakage energy.  It was observed during HECT breakage 
of view has been the difficulty and
  The Edge-Notched Disk Wedge Splitting test yields representative fracture toughness 
values  (in  accordance  with  Level  I  testing)  using  a  simple,  core  based  specimen.    It 
requires  little  specimen  preparation  time  since  it  employs  a  straight-through  notch.  
Testing, load application, and data collection are straightforward.  The test enables a large 
number  of  rock  types  and  specimens  to  be  tested  allowing  for  quick  indexing  and 
rocks tested in this study a linear relationship most adequately describes the relationship 
between  fracture  toughness  and  specific  comminution  energy.    As  fracture  toughness 
increases  so  does  the  energy  required  to  achieve  a  certain  amount  of  size  reduction  (as 
defined by the reduction ratio, the ratio of the initial particle size to the closed side set).  
As  the  reduction  ratios  increases  so  does  the  specific  comminution  energy.    Therefore, 
the  specific  comminution  energy  is  function  of  both  the  fracture  toughness  of  the  rock 
material  and  the  operational  characteristics  of  the  jaw  crusher,  as  defined  by  the  closed 
side set.   
 
  Of the rocks tested only one, a siltstone, did not fall within the 95% confidence limit 
for  the  suggested  linear  relationship  between  fracture  toughness  and  specific 
comminution energy.  The siltstone had the highest fracture toughness yet was middle of 
cture toughness was shown to be related to specific comm
an any other material property tested, including tensile s
  139
of  the  siltstone  that  the  rock  tended  to  shatter  upon  fracture,  producing  a  wide  range  of 
progeny  sizes.    The  lack  of  daughter  fragments  created  by  cleavage  resulted  in  less 
secondary crushing and consequently lower levels of specific comminution energy.  This 
observation  was  substantiated  quantitatively  by  the  uniform  distribution  of  progeny 
particles seen in the siltstones breakage function.       
 
When  the  the  feed  size  distribution,  the  mass  flow  at  each  size,  the  probability  of 
breakage  and  the  idling  power  are  taken  into  account,  the  relationship  between  fracture 
toughness  and  specific  comminution  energy  can  be  used  to  define  a  model  for  the 
prediction of jaw crusher power consumption, given by: 
 
|   |
1
0.428
1
0.511 0.511     for 1 1.5 
0.215             for  1.5
j
c i Ic i i n
i
j
c i Ic i i n
i
P RR K C x P
P RR K C x P RR
=
=
=      +   +      <
   (
=   +
   
RR
   
where, P
c
 is the power consumption of the crusher in kW 
    RR
i
 is the reduction ratio for a particle of size I 
    K
Ic
 is the fracture toughness of the rock being crushed in  MPa m  
    C
i
 is the probability of breakage for particle size i 
    x
i
 is the mass flow of particle size i in metric tons per hour 
    P
n
 is the power drawn by the crusher under no load in kW. 
Since the max  to be made 
in  regards  to  what  occurs  at  higher  reduction  ratios,  and  the  applicability  of  the  power 
tween  fracture  toughness  and  specific  comminution  energy.  
he  common  procedure  of  expressing  breakage  functions  as  mixtures  of  separate  size 
pop
ck  was  taken  into  account.    Additional  relationships 
between  t
10
  and  various  t
n
  values  were  also  evident,  thus  allowing  for  the  determination 
of a rocks breakage function based soley on its fracture toughness and
 
nergy release rate of the rock.  This is because for rocks with higher 
rain energy release rates shatter fracture is more prevalent and more energy is released 
rogeny  undergo  additional  breakage  in 
imum reduction ratio used in this study was 3, assumptions had
consumption model to higher reduction ratios remains to be tested.  
 
  The relationship between breakage function and fracture toughness was not as strong 
or  as  evident  as  the  one  be
T
ulations (fine and coarse) was not employed due to the lack of a relationship between 
the size proportion and distribution parameters and fracture toughness.  Instead, the one-
parameter family curves approach was used.  In this case a relationship between t
10
, the 
characteristic size reduciton parameter, and fracture toughness was evident but only when 
the  elastic  modulus,  E,  of  the  ro
 elastic modulus. 
    The  relationship  demonstrated  by  the  data  shows  that  t
10
  increases  with  K
Ic
2
/E, 
which is the strain e
st
as new crack surface energy.  As the reduction ratio is increased, the relationship between 
t
10
 and K
Ic
2
/E inverses, and t
10
 decreases with increasing strain energy release rate.  The 
low  strain  energy  release  rate  rocks  are  subjected  to  more  secondary  breakage  than  the 
higher  K
Ic
2
/E  rocks  because  the  low  K
Ic
2
/E  rocks  initially  produce  large  progeny 
fragments  as  a  result  of  cleavage.    These  p
  140
comparison  to  the  fragments  of  the  high  strain  energy  release  rate  rocks,  which  are 
roduced by shatter and have more particles of varying size.   
e breakage function is: 
p
 
The model for th
|   |
(   )
(   )
10
0.190 0.0127 13.4 0.0998
2
10 3.40
0.281
1.5 10
7.01 7.01                                for 1 <1.5
'
100
100        for  1.5
'
1
3.97
40.9
RR
Ic
RR
Ic
t RR RR
E
K
t RR
E
RR
t t
   +   +
=     +   
   |
\   .
   (
   (
   (
=      
   |
   (
   (
   
=
0.800
2 10
7.68
2.17
t t
t t
=
=
(   ) 0.501 0.501
2
RR
K
     +
|   |
4 10
|   |
|   |
  \   .
+
   |
\   .
10 10
25 10
50 10
75 10
100 10
0.556
0.384
0.315
0.271
t t
t t
t t
t t
t t
=
=
=
=
=
 
where, t
10
 is the percent passing at a progeny size 1/10
th
 of the parent size 
    K
Ic
 is the fracture toughness of the rock being crushed in  MPa m  
    t
n
 values are percent passing at a progeny size 1/n
th
 of the parent size 
and,  E is given by, 
'              for conditions of plane stress E E =
   
2
'        for conditions of plane strain
1
E
E
v
=
|   |
=   
   |
\   .
=   
|   |
   |
\   .
 
where, F
avg
 is the average feed size in m. 
 
  Full verification of the models for power consumption and volumetric capacity were 
not  performed.    Verification  of  the  model  strictly  for  specific  comminution  energy  was 
performed,  along  with  verification  of  the  breakage  function  model  using  the  HECT 
system.  The breakage functon model was tested further using the laboratory jaw
=   +   +
 crusher, 
y  comparing  the  actual  product  size  distribution  with  the  predicted  product  size 
dis
b
tribution based on Whitens crusher model and the breakage function. 
 
  For  the  HECT  system  verification  experiment,  dolomitic  limestone,  which  had 
previously been tested for fracture toughness using the END test, was used.  The fracture 
toughness  based  models  for  specific  comminution  energy  and  t
10
  predicted  that  the 
  142
dolomitic  limestone  would  require  0.305  kWh  per  metric  ton  of  energy  to  crush,  with 
brekage  resulting  in  a  t
10
  value  of  6.6%.    The  actual  values  were  0.295  kWh/ton  and 
.0%, respectively.  The entire predicted breakage function also showed good agreement 
 agreement 
ith  the  actual  product  size  produced  by  the  crusher.    There  is  some  deviation  at  larger 
on 
quirements.    This  would  result  in  lower  capital  cost  and  lower  operating  costs.    It  is 
p
iven. 
 
6
with  the  measured  function,  with  some  deviation  occurring  at  large  t
n
  values  since  the 
breakage function model loses power with increasing t
n
. 
 
  The  prediction  of  product  size  for  the  laboratory  scale  jaw  crusher  was  performed 
using  four  rock  types.    Employing  Whitens  crusher  model,  based  on  a  mass  balance 
approach,  the  product  size  was  predicted  using  the  breakage  function,  classification 
function, and feed size distribution.  The predicted product size showed good
w
sizes due to the effect of the disk specimen used to develop the breakage function model.  
In two cases the predicted percent passing was slightly higher than measured, and in the 
other two instances it was slightly lower.  Overall the predicted product size distribution 
was within  5%, in terms of percent passing (by weight) a given particle size. 
 
  Unfortunately  full  scale  verification,  using  jaw  crushers  in  operation  at  Luck  Stone 
sites,  could  not  be  performed.    One  reason  is  because  the  feed  size  distribution  entering 
Luck  Stones  jaw  crushers  has  never  been  determined.    Secondly,  Luck  Stone  employs 
essentially  the  same  size  jaw  crusher  at  each  of  its  operations  and  comparison  of  the 
predicted  power  consumption  to  the  installed  electric  drive  would  be  baseless  since  the 
same drive is used at each operation.  But at the same time this is the argument for using 
a  new  approach  to  selecting  jaw  crushers.    Based  on  the  model  suggested  here,  the 
toughest  rock  tested  would  consume  230%  more  power  than  the  weakest  for  the  same 
feed  size  and  average  reduction  ratio.    In  turn,  a  smaller  crusher,  driven  by  a  smaller 
motor,  could  be  used  for  the  weaker  rock  as  long  as  it  is  able  to  meet  producti
re
im ortant  to  realize  that  this  would  also  require  a  shift  away  from  avoiding  secondary 
breakage at all costs, the main reason large crushers are employed almost universally.  Or 
it  would  require  a  change  in  blast  specifications  such  that  the  feed  material  entering  the 
jaw crusher will not bridge.   
 
For cases such as Luck Stones, where jaw crushers are already in place, the breakage 
function  and  volumetric  capacity  models  developed  in  this  study  can  be  used  to 
determine  feed  size  distributions  and  operational  settings  (i.e.,  closed  side  set)  that 
optimize  the  product  size  and  production  capacity.    This  would  improve  downstream 
processes since control over their feed sizes would be excercised.  Blasting would also be 
improved since guidelines for muck size disitribution, which becomes the feed material, 
could be g
This  study  clearly  demonstrates  that  a  fracture  mechanics  approach  to  the  study  of 
size reduction processes is appropriate, and most thoroughly characterizes the process of 
rock  fragmentation  in  crushing  equipment.    Rock  fracture  toughness  is  related  to  the 
amount  of  energy  required  to  reduce  a  particle  to  a  given  size.    It  can  also  be  related  to 
  143
the  resultant  fractured  size  distribution  and  the  volumetric  capacity  of  a  jaw  crusher.  
Equipment  selection  and  optimization  can  now  be  performed  using  a  material  property 
that  properly  represents  the  fracture  process  occurring  in  a  jaw  crusher.    Fracture 
toughness  can  be  rapidly  determined  using  the  END  test,  and  along  with  feed  size  data, 
and  the  average  amount  of  size  reduction,  K
Ic
  can  be  used  to  determine  the  power 
onsumption,  product  size,  and  the  volumetric  capacity  of  a  jaw  crusher  processing  a 
giv
6.2 RECOMMENDATIONS 
 
All  size  reduction  processes  are  dependent  upon  the  mechanical  parame
reduction and the nature of the rock being broken.  Therefore all size reduction and rock 
fragmentation  processes,  beginning  with  blasting,  should  be  analyzed  using  fracture 
ed  using  some  measure  of  its 
in energy release rate. 
  test  was  verified  in  this  study, 
size  and  notch  radius.  
ticularly  in  regards  to 
veloped in this study 
eed  to  either  verified  or  modified,  based  on  additional  testing,  for  reduction  ratios 
e 
ower  consumption  of  cone  crushers  require  scale-up  factors,  and  the  same  may  be 
c
en  rock.    The  process  can  also  be  reversed  so  that  fracture  toughness,  along  with 
product size and/or desired capacity, can be used to optimize the feed size or closed side 
set of the crusher. 
 
ters  of  the  size 
mechanics,  with  the  nature  of  the  rock  being  characteriz
ability to resist fracture, either fracture toughness or stra
 
  Although  the  applicability  of  the  Edge  Notched  Disk
there  is  some  evidence  that  the  results  are  dependent  on  grain 
Additional  refinement  of  the  END  test  needs  to  be  performed,  par
using a chevron notch instead of a straight through notch. 
 
  The models for power consumption and the breakage function de
n
greater than 3.  The breakage function model also requires more testing in order to further 
clarify  the  relationship  between  fragmented  product  size  and  fracture  toughness/strain 
energy release rate.  Testing at higher reduction ratios will also improve the relationships 
between  t
10
  and  other  t
n
  vales  used  to  determine  the  breakage  size  distribution.    Since 
only  6  rock  types  were  used  in  this  study  the  models  proposed  could  be 
strengthened/adjusted simply by testing more rocks that are both inside and outside of the 
measured range.  Specifically, more sedimentary rocks types should be tested in order to 
determine the how representative of sedimentary rocks the Culpeper siltstone is.   
 
  All  the  models  proposed  here  require  full-scale  verification.    Similar  models  for  th
p
expected for the jaw crusher power consumption model.  The volumetric capacity model 
is based only on the results of lab-scale crushing of four rocks, and it too may be subject 
to a scale-up factor since actual jaw crushers are on the order of 10 times bigger than the 
lab jaw used in this study.  The feed size distribution used in the lab crushing experiments 
was  fairly  narrow  and  the  scale-up  effect  of  feed  size  and  distribution  must  also  be 
accounted for.  
 
  144
   This investigation relied solely on the testing of rock core specimens (except for the 
laboratory jaw tests).  In reality, rocks fed  to a jaw crusher have undergone a change in 
eir  microstructure  due  to  blasting.    Although  it  would  be  expected  that  similar 
rela hness,  specific  comminution  energy,  and 
breakage  function,  a  similar  study  should  be  undertaken  using  rock  specimens  prepared 
from blasted rock.  The effect of blasting on the distribution and size of flaws, and thus 
fracture toughness, can be examined.  
 
  Likewise,  single  particle  breakage  testing  should  be  done  on  irregular  particles  that 
more  closely  resemble  those  found  in  the  feed.    The  influence  of  the  disk  specimen 
geometry  on  the  breakage  distribution  can  then  be  determined,  as  well  as  it  ability  to 
mimic an irregular particle under point contact loading. 
 
  It  is  clear  that  fracture  mechanics  and  fracture  toughness  can  be  used  to  model  and 
predict  the  behavior  of  both  rock  and  machine  when  considering  jaw  crushing 
applications.    Jaw  crushers  can  now  be  selected  or  optimized  based  on  only  a  few 
parameters as long as the fracture toughness of the rock being broken is known.  Proper 
characterization of the rock, along with the exclusion of overemphasized factors such as 
design  experience  and  secondary  breakage,  allows  for  the  selection  of  suitably  sized 
equipment and/or the optimization of the product and production capacity.  This results in 
reduced  capital  costs,  reduced  operating  costs,  efficient  production,  and  an  overall 
improvement in the entire size reduction process.      
 
ort in studying 
eds  to  be  carried  out  in 
pplication of 
ther size reduction applications.   
 
 
th
tionships  hold  between  fracture  toug
It is also apparent that this study is just the beginning, and that more eff
size  reduction  from  a  rock  fracture  mechanics  point  of  view  ne
order to refine and verify the findings outlined here, as well as to widen the a
fracture mechanics to o
  145
REFERENCES 
 
Akram,  M.  (1991)  The  Effect  of  Zero  Point  of  Charge  Environment  On  Rock  Fracture 
Behavior, MS Thesis - Virginia Tech, Blacksburg, VA. 
 
Allis-Chalmers (1985) Allis-Chalmers High Energy Crushing Test System User Manual, 
Version 1.0. 
 
Andersen, J.S. (1988) Development of a Cone Crusher Model, MEng Thesis  University 
f Queensland, Australia. 
rd
 
Mill Operators Conference, AusIMM, Cobar, Australia. 
 
Andersen, J.S. and Napier-Munn, T.J. (1990) The Influence of Liner Condition on Cone 
Crusher Performance, Minerals Engineering, Vol. 3, No. 1-2, pp. 105-116. 
 
.A. (1969) Single Fracture of Brittle Spheres, 
AIME, New York, pp. 118-133. 
.  (1983)  A  Comparison  of  the  the  Bond  Method  for  Sizing 
odel, Powder Technology, Vol. 34, 
Barker,  L.M.  (1977)  A  Simplified  Method  for  Measuring  Plane  Strain  Fracture 
Toughness, Engineering Fracture Mechanics, Vol. 9, pp. 361-369. 
 
Barton,  C.C.  (1983)  Variables  in  Fracture  Energy  and  Toughness  of  Rock,  Proceedings 
of the 24
th
 U.S. Symposium On Rock Mechanics, Berkeley, CA, pp. 449-462. 
 
Bearman, R.A., Pine, R.J., and Wills, B.A. (1989) Use of Fracture Toughness Testing in 
Characterizing  the  Comminution  Potential  of  Rock,  Proceedings  of  MMIJ/IMM  Joint 
 12, 
p. 1243-1256. 
o
 
Andersen,  J.S.  and  Napier-Munn,  T.J.  (1988)  Power  Prediction  for  Cone  Crushers,  3
Arbiter, N., Harris, C.C., and Stamboltzis, G
Transactions of SME/ AIME, Vol. 244, 
 
Austin,  L.G.  and  Brame,  K
Wet Tumbling Ball Mills With An SMB Simulation M
pp. 261-274. 
 
Symposium, Kyoto, pp. 161-170. 
 
Bearman,  R.A.,  Barley,  R.W.,  and  Hitchcock,  A.,  (1990)  The  Development  of  a 
Comminution Index for Rock and The Use of an Expert System to Assist The Engineer in 
Predicting Crushing Requirements, Minerals Engineering, Vol. 3, No. 1-2, pp. 117-127. 
 
Bearman,  R.A.,  Barley,  R.W.,  and  Hitchcock,  A.  (1991)  Prediction  of  Power 
Consumption and Product Size in Cone Crushing, Minerals Engineering, Vol. 4, No.
p
 
  146
Bearman,  R.A.,  Briggs,  C.A.,  and  Kojovic,  T.  (1997)  The  Application  of  Rock 
Mechanics  Parameters  to  the  Prediction  of  Comminution  Behavior,  Minerals 
Engineering, Vol. 10, No. 3, pp. 255-264. 
 
Bearman, R.A. (1998) The Use of the Point Load Test for the Rapid Estimation of Mode 
Fracture  Toughness,  International  Journal  of  Rock  Mechanics  and  Mining  Sciences, 
e  Particle 
rushing, Proceedings of 5
th
 Rock Mechanics Symposium, pp. 155-172. 
d  Herbst,  J.A.  (1992)  Low-Impact-Energy  Single-Particle 
racture,  Comminution    Theory  and  Practice,  ch.  8,  (edited  by  S.  Komar  Kawatra). 
SM
 Bearman, R.A. (1996) An Investigation of Rock Breakage and Damage 
in Comminution Equipment, Minerals Engineering, Vol. 9, No. 5, pp. 489-497. 
 
I 
Vol. 36, pp. 257-263. 
 
Bergstrom,  B.H.,  Sollenberger,  C.L.,  and  Mitchell,  W.  (1962)  Energy  Aspects  of  Single 
Particle Crushing, Transactions AIME, Vol. 223, pp. 367-372. 
 
Bergstrom,  B.H.  and  Sollenberger,  C.L.  (1962)  Kinetic  Energy  Effect  in  Single  Particle 
Crushing, Transactions AIME, Vol. 223, pp. 373-379. 
 
Bergstrom,  B.H.  (1965)  Energy  and  Size  Distribution  Aspects  of  Singl
C
 
Berry, P., Dantini, E.M., and Masacci, P. (1984) Influences of Mechanical Characteristics 
of  Rock  on  Size  Reduction  Processing,  Proceedings  of  Mineral  Processing  and 
Extractive Metallurgy, Beijing, pp. 15-26. 
 
Beke, Bela. (1964) Principles of Comminution, Akademiai Kiado, Budapest. 
 
Black,  R.A.  and  Hoek,  E.  (1967)  Status  of  Rock  Mechanics  as  Applied  to  Mining, 
Proceedings  of  the  9
th
  Symposium  on  Rock  Mechanics,  AIME,  Colorado  School  of 
Mines, Golden, CO. 
 
Bond, F.C. (1952) The Third Theory of Comminution, Transactions of AIME/SME, Vol. 
193, pp. 4884-494. 
 
Bourgeois,  F.,  King,  R.P.,  an
F
E, Littleton, CO, pp. 99-108. 
 
Bowers,  L.R.,  Broaddus,  W.R.,  Dwyer,  J.,  Hines,  J.,  and  Stansell,  R.  (1991)  Processing 
Plant  Principles,  The  Aggregate  Handbook,  ch.  8,  (edited  by  Richard  D.  Barksdale), 
NSA, Washington D.C. 
 
Briggs, C.A. and
Bubsey,  R.T.,  Munz,  D.,  Pierce,  W.S.,  Shannon,  J.L.  (1982)  Compliance  Calibration  of 
the  Short  Rod  Chevron-Notch  Specimen  for  Fracture  Toughness  Testing  of  Brittle 
Materials, International Journal of Fracture, Vol. 18, no. 2, pp. 125-133. 
  147
 
Carey, W.F. and Bosanquet, C.H. (1933) A Study of Crushing Brittle Solids, Journal of 
the Society of Glass Technology, Vol. 17, pp. 348.  
 
Charles,  R.J.  (1957)  Energy-Size  Relationships  in  Comminution,  Transactions 
AIME/SME, Vol. 208, pp. 80-88. 
 
Choi,  Woo-Zin.  (1982)  Comminution  and  Liberation  Studies  of  Complex  Sulfide  Ores, 
MS Thesis  Virginia Tech, Blacksburg, VA. 
s of Strain Rate on Oil Shale Fracturing, International Journal 
f Rock Mechanics, Mining Science, and Geomechanics Abstracts, Vol. 17, pp. 35-43. 
hong,  K.P.,  Kuruppu,  M.D.,  and  Kuszmaul,  J.S.  (1987)  Fracture  Toughness 
ey and Sons, New York. 
for  the 
apid  Determination  of  Mode  I  Fracture  Toughness  and  the  Assessment  of  Relative 
de Tourenq). Balkema, Rotterdam. 
 
Chong, K.P. (1980) Effect
o
 
Chong,  K.P.  and  Kuruppu,  M.D.  (1984)  New  Specimen  for  Fracture  Toughness 
Determination of Rock and Other Materials, International Journal of Fracture, Vol. 26, 
pp. R59-R62. 
 
C
Determination  of  Rocks  with  Core  Based  Specimens,  SEM/RILEM  International 
Conference on Fracture of Concrete and Rock, Houston, pp. 13-25.  
 
Clark, C.B. (1987) Principles of Rock Fragmentation, Wil
 
Csoke, B., Petho, S., Foldesi, J., and Meszaros, L. (1996) Optimization of Stone-Quarry 
Technologies, International Journal of Mineral Processing, Vol. 44-45, pp. 447-459. 
 
DeDiemar,  R.B.  (1990)  New  Concepts  in  Crusher  Technology,  Minerals  Engineering, 
Vol. 3, No. 1-2, pp. 67-74. 
 
Deliac,  E.P.  (1986)  Optimization  Des  Machines  Dabattage  a  Pics,  PhD  Thesis   
University of Paris, France. 
 
Donovan,  J.G.  and  Karfakis,  M.G.  (2003)  Adaptation  of  a  Simple  Wedge  Test 
R
Fracture  Resistance,  International  Journal  of  Rock  Mechanics  and  Mining  Sciences, 
Technical Note, In Press.  
 
Dowling,  N.E.  (1999)  Mechanical  Behavior  of  Materials    Engineering  Methods  for 
Deformation, Fracture, and Fatigue, Prentice Hall, Upper Saddle River, NJ. 
 
Duthoit,  Vincent  (2000)  Crushing  and  Grinding,  Aggregates,  ch.  9,  (edited  by  Louis 
Primel and Clau
 
  148
Eloranta,  J.  (1997)  The  Efficiency  of  Blasting  Versus  Crushing  and  Grinding, 
Proceedings of the 23
rd
 Annual Conference on Explosives and Blasting Techniques, Las 
Vegas, pp. 157-163. 
 
Fairs, G.L. (1954) A Method of Predicting the Performance of Commercial Mills in the 
Fine  Grinding  of  Brittle  Materials,  Transactions  of  the  Institute  of  Mining  and 
Metallurgy, Vol. 63, pp. 211-240. 
 
Fenghui, W. (2000) Prediction of Intrinsic Fracture Toughness for Brittle Materials From 
the Apparent Toughness of Notched-Crack Specimen, Journal of Materials Science, Vol. 
35, pp. 2443-2546. 
 
Gaudin,  A.M.  and  Hukki,  R.T.  (1946)  Principles  of  Comminution,  Transactions 
and  Comminution  of  Brittle  Solids: 
urther Experimental Results, Transactions AIME/SME, Vol. 226, pp. 412-422. 
 
riffith,  A.A.  (1921)  The  Phenomenon  of  Rupture  and  Flow  in  Solids,  Philosophical 
Transactions of the Royal Physical Society, Vol. 221, pp. 163-198.  
 
Guo,  H.  (1990)  Rock  Cutting  Using  Fracture  Mechanics  Principles,  PhD  Thesis  -  The 
University of Wollongong, Australia. 
 
Hanish,  J.  and  Schubert,  H.  (1986)  Compressive  Comminution  of  Particle  Beds, 
Aufbereitungs Technik, No. 10, pp. 535-539.  
 
Herbst, J.A. and Fuersternau, D.W. (1972) Influence of Mill Speed and Ball Loading on 
the  Parameters  of  the  Batch  Grinding  Equation,  Transactions  AIME/SME,  Vol.  252,  pp. 
169-176. 
 
Herbst,  J.A.  and  Lo,  Y.C.  (1992)  Microscale  Comminution  Studies  for  Ball  Mill 
Modeling,  Comminution    Theory  and  Practice,  ch.  11,  (edited  by  S.  Komar  Kawatra). 
SME, Littleton, CO, pp. 137-152. 
AIME/SME, Vol. 223, pp. 40-43. 
 
Gaudin, A.M. and Meloy, T.P. (1962) Model and a Comminution Distribution Equation, 
Transactions AIME, Vol. 169, pp. 67-87. 
 
Gilvarry, J.J. (1961) Fracture of Brittle Solids  Distribution Function for Fragment Size 
in Single Fracture (Theoretical), Journal of Applied Physics, Vol. 32, pp 391-399. 
 
Gilvarry,  J.J.  and  Bergstrom,  B.H.  (1962)  Fracture 
F
Gregory,  R.D.  (1979)  The  Edge-Cracked  Circular  Disc  Under  Symmetric  Pin-Loading, 
Mathematical  Proceedings  of  the  Cambridge  Philosophical  Society,  Vol.  85,  pp.  523-
538. 
 
G
  149
 
ersam,  E.A.  (1923)  Factors  Controlling  the  Capacity  of  Rock  Crushers,  Transactions 
ofler,  Andreas.  (1990)  Fundamental  Breakage  Studies  of  Mineral  Particles  With  An 
olmes, J.A. (1957) A Contribution to the Study of Comminution  A Modified Form of 
ondros, J.R. (1959) The Evaluation of Poissons Ratio and the Modulus of Materials of 
ence, no. 10, pp. 243-268. 
08. 
,  Proceedings  of 
e 23rd U.S. Symposium on Rock Mechanics, The University of California, pp. 463-470. 
ineers' Working Party Concerned With the Theory and Practice of the Size 
eduction of Solid Materials, (edited by V.C. Marshall), I. Chem. E. Services, London. 
s and Sharp Corners, 
ransactions of the Institute of Naval Architects, pp. 219-230.  
  Society  of 
etals Symposium, Chicago, pp. 147-166.    
nics, Vol. 24, pp. 361-364. 
bstracts, Vol. 25, pp. 73-96. 
Jomoto,  K.  and  Majima,  H.  (1972)  Criterion  of  Comminution,  CIM  Transactions,  Vol. 
75, pp. 326-331. 
H
AIME, Vol. 68, pp. 463-476. 
 
H
Ultrafast Load Cell Device, PhD Dissertation  University of Utah, Salt Lake City, UT. 
 
H
Kicks Law, Transactions of the Institute of Chemical Engineers, Vol. 35, pp. 125-156. 
 
H
a  Low  Tensile  Resistance  by  the  Brazilian  (Indirect  Tensile)  Test  With  Particular 
Reference to Concrete, Australian Journal of Applied Sci
 
Hukki,  R.T.  (1961)  Proposal  for  a  Solomonic  Settlement  Between  the  Theories  of  von 
Rittenger, Kick, and Bond, Transactions of AIME/SME, Vol. 220, pp. 403-4
 
Ingraffea,  A.R.,  Gunsallus,  K.L.,  Beech,  J.F.,  Nelson,  P.  (1982)  A  Fracture  Toughness 
Testing  System  for  Prediction  of  Tunnel  Boring  Machine  Performance
th
 
Institution  of  Chemical  Engineers  (1974)  Comminution:  A  Report  by  the  Institution  of 
Chemical Eng
R
 
Inglis, C.E. (1913) Stresses in a plate Due to the Presence of Crack
T
 
Irwin,  G.R.  (1948)  Fracture  Dynamics,  Fracturing  of  Metals  -  American
M
 
Irwin, G.R. (1957) Analysis of Stresses and Strains Near the End of a Crack Transversing 
a Plate, Journal of Applied Mecha
 
Isida,  M.,  Imai,  R.,  and  Tsuru,  H.  (1979)  Symmetric  Plane  Problems  of  Arbitrarily 
Shaped Plates With an Edge Crack, Transactions of the Japanese Society of Mechanical 
Engineers, Vol. 45, 395A, pp. 743-749. 
 
ISRM  (1988)  Suggested  Methods  for  Determining  the  Fracture  Toughness  of  Rocks, 
(coordinator  F.  Ouchterlony),  International  Journal  of  Rock  Mechanics  and 
Geomechanics A
 
  150
 
Jowett  A.  and  van  der  Waerden,  N.M.  91982)  Adaptation  of  Single  Particle  Hardness 
Test  for  Application  to  Iron  Ores  and  Associated  Rocks,  Powder  Technology,  Vol.  32, 
p. 9-19. 
f  Rocks,  Proceedings  of  the  27
th
  U.S.  Symposium  on  rock 
echanics, University of Alabama, SME, pp. 3-10. 
rushers,  XIV  International 
ineral  Processing  Congress:  Worldwide  Industrial  Application  of  Mineral  Processing 
elly, E.G and Spottiswood, D.J. (1982) Introduction to Mineral Processing, John Wiley 
elly,  E.G  and  Spottiswood,  D.J.  (1990)  The  Breakage  Function;  What  Is  It  Really?,  
e  Derivation  of  a  Mathematical  Model  for 
reakage  in  Small,  Continuous  Wet  Ball  Mill,  Joint  AIChE  and  Institute  of  Chemical 
Engineering, Vol. 33, no. 3, pp. 179-206. 
s Journal, Vol. 250, pp. 141-145. 
ojovic,  T.,  Napier-Munn,  T.J.,  and  Andersen,  J.S.  (1997)  Modeling  Cone  and  Impact 
limpel,  R.R.  and  Austin,  L.G.  (1965)  The  Statistical  Theory  of  Primary  Breakage 
nott,  J.F.  (1972)  Fundamentals  of  Fracture  Mechanics,  John  Wiley  and  Sons,  New 
p
 
Karfakis,  M.G.,  Chong,  K.P.,  and  Kuruppu,  M.D.  (1986)  A  Critical  Review  of  Fracture 
Toughness  Testing  o
M
 
Karra,  V.J.  (1982)  A  Process  Performance  Model  for  Cone  C
M
Technology, Toronto, paper III-6. 
 
K
and Sons, New York. 
 
K
Minerals Engineering, Vol. 3, No. 5, pp. 405-414. 
 
Kelsall,  D.F.  and  Reid,  K.J.  (1965)  Th
B
Engineers Meeting, London, Paper 4.2. 
 
Khan, K. and Al-Shayea, N.A. (2000) Effect of Specimen Geometry and testing Method 
on Mixed-Mode I-II Fracture Toughness of a Limestone Rock From Saudi Arabia, Rock 
Mechanics and Rock 
 
Kick,  F.  (1883)  The  Law  of  Proportional  Resistance  and  Its  Application  to  Sand  and 
Explosions, Dingler
 
King, R.P. (2001) Modeling and Simulation of Mineral Processing Systems, Butterworth-
Heinemann, Boston. 
 
K
Crushers  Using  Laboratory  Determined  Energy-Breakage  Functions,  Comminution 
Practices, ch. 25, (edited by S. Komar Kawatra), SME, Littleton, CO. 
 
K
Distribution for Brittle Materials, Transactions AIME/SME, Vol. 232, pp. 88-94. 
 
K
York. 
 
  151
Krough,  S.R.  (1980)  Determination  of  Crushing  and  Grinding  Characteristics  Based  On 
im,  I.  L.,  Johnston,  I.  W.,  Choi,  S.  K.,  Boland,  J.  N.  (1994)  Fracture  Testing  of  a  Soft 
g Sciences, and Geomechanics Abstracts, Vol. 31, pp. 
85-197. 
, CRC Press, Cleveland, Ohio. 
Testing of Single Particles, AIME Transactions, Vol. 266, pp. 1957-1962. 
 
L
Rock  With  Semicircular  Specimens  Under  Three  Point  Bending  -  Part  1,  International 
Journal of Rock Mechanics, Minin
1
 
Lowrison, G.C. (1974) Crushing and Grinding
 
Luck Stone Corporation (2003) http://www.luckstone.com/crushed. 
 
Lynch,  A.J.  (1977)  Mineral  Crushing  and  Grinding  Circuits:  Their  Simulation, 
assacci G., and Rossi, G. (1983) Energy Requirements of Rock Breaking 
  Mining  Operations    Considerations  and  Proposals,  1
st
  International  Symposium  on 
pression, 
1
st
 U.S Symposium on Rock Mechanics, pp. 410-419. 
etso Minerals (2003) Nordberg C Series Jaw Crushers, Technical publication. 
ng, 
echnical literature. 
iction  of  Crusher  Power  Requirements  and  Product  Size 
nalysis,  Design  and  Installation  of  Comminution  Circuits,  (edited  by  A.L.  Mular  and 
orrell, S., Napier-Munn, T.J., and Andersen, J.S. (1992) The Prediction of Power Draw 
s intensity Factors Handbook, Pergamon Press, New York.  
nd  Bending, 
roningen P. Noordhoff. 
oratory  Single  Particel  Breakage 
echnique  and  Its  Application  to  Ball  Mill  Modeling  and  Scale-Up,  PhD  Dissertation   
University of Queensland. 
Optimization, Design, and Control, Elsevier, New York. 
 
Manca, P.P, M
in
Rock Fragmentation by Blasting, Lulea, Sweden, pp. 705-724. 
 
Mehrim,  M.R.  and  Khalaf,  F.  91980)  Kinetics  of  Comminution  in  Slow  Com
2
 
M
 
Metso  Minerals  (2003a)  Your  Comprehensive  Source  for  Aggregate  Processi
T
 
Moore,  D.C.  (1982)  Pred
A
G.V. Jergensen), SME/AIME, pp. 218-227.  
 
M
for  Comminution  Machines,  Communition    Theory  and  Practice,  (edited  by  S.  Komar 
Kawatra), AIME, pp. 405-426. 
 
Murakami, Y. (1987) Stres
 
Muskhelishvili,  N.I.  (1963)  Some  Basic  Problems  of  the  Mathematical  Theory  of 
Elasticity;  Fundamental  Equations,  Plane  Theory  of  Elasticity,  Torsion,  a
G
 
Narayanan,  S.S.  (1985)  Development  of  a  Lab
T
  152
 
Narayanan,  S.S.  and  Whiten,  W.J.  (1988)  Determination  of  Characteristics  from  Single 
Particle  Breakage  Test  and  Its  Application  in  Ball  Mill  Scale-up,  Transactions  of  IMM, 
7, Section C, pp. C1115-C124. 
eration  and  Optimization,  Julius  Kruttschnitt  Mineral 
esearch Center, Queensland. 
acture  Material  Properties,  Proceedings  of  the  27
th
  U.S.  Symposium 
n Rock Mechanics, pp. 846-852. 
n,  R.B.  (1987)  Optical  Detection  of  Crack  Patterns  in 
e Opening Mode Fracture of Marble, International Journal of Rock Mechanics, Mining 
ordberg (1995) Compression Test, Laboratory Test Guidelines, pp. 3-9. 
racture 
echanics, Canadian Metallurgical Quarterly, Vol. 9, pp. 429-339. 
ics, Vol. 
2, pp. 185. 
onic Research Foundation, Stockholm. 
  By  Blasting,  Lulea, 
weden, pp. 199-223. 
ness Testing of Rock With Core Based Specimens 
The  Development  of  an  ISRM  Standard,  Fracture  Toughness  and  Fracture  Energy, 
uchterlony,  F.  (1989a)  On  the  Background  to  the  Formula  and  Accuracy  of  Rock 
3.  
9
 
Napier-Munn,  T.J.,  Morrell,  S.,  Morrison,  R.D.,  Kojovic,  T.  (1999)  Mineral 
Comminution  Circuits    Their  Op
R
 
Nelson,  P.P.  and  Fong,  F.L.C.  (1986)  Characterization  of  Rock  For  Boreability 
Evaluation  Using  Fr
o
 
Nolen-Hoeksema,  R.C.  and  Gordo
th
Sciences, and Geomechanics Abstracts, Vol. 24, no. 2, pp. 135-144. 
 
N
 
Oka,  Y.  and  Majima,  W.  (1970)  A  Theory  of  Size  Reduction  Involving  F
M
 
Orowan, E. (1949) Fracure and Strength of Solids, Reports on Progress in Phys
1
 
Ouchterlony,  F.  (1980)  A  New  Core  Specimen  for  the  Fracture  Toughness  Testing  of 
Rock, SveDeFo Report DS1980:17, Swedish Det
 
Ouchterlony,  F.  and  Sun,  Z.  (1983)  New  Methods  of  Measuring  Fracture  toughness  on 
Rock  Cores,  1
st
  International  Symposium  on  Rock  Fragmentation
S
 
Ouchterlony, F. (1989) Fracture Tough
- 
(edited by Mihashi et al.), Balkema, Rotterdam, pp. 231-251. 
 
O
Fracture Toughness Measurements Using ISRM Stnadard Core Specimens, International 
Journal  of  Rock  Mechanics,  Mining  Science,  and  Geomechanics  Abstracts,  Vol.  26,  no. 
1, pp. 13-2
 
Pauw,  O.G.  and  Mare,  M.S.  (1988)  The  Determination  of  Optimum  Impact  Breakage 
Routes for an Ore, Powder technology, vol. 54, pp. 3-13. 
 
  153
ennsylvania  Crusher  Corporation  (2003)  Crusher  Selection,  Handbook  of  Crushing  -  P
http://penncrusher.com/1sol1_3.html.   
 
Peter Rit A., Hagg, H., and van Brakel, J. (1983) Characterization of Mechanical Strength 
d 
onference, Birmingham, UK, pp. 1-19.  
d 
pplications, WIT Press, Boston. 
ances  in 
ineral Processing, (edited by P. Somasundaran), SME/AIME, pp. 575-593. 
ooke,  D.P.  and  Cartwright,  D.J.  (1976)  Compendium  of  Stress  Intensity  Factors, 
e  Impact  of  Limestone 
pheres,  Rock  Mechanics  as  a  Multidisciplinary  Science  Proceedings  of  the  32
nd
  U.S. 
astri, S.R (1994) Capacities and Performance Characteristics of Jaw Crushers, Minerals 
chmidt,  R.A.  (1980)  A  Microcrack  Model  and  Its  Significance  to  Hydraulic  fracturing 
chmidt,  R.A  and  Rossmantihm  H.P  (1983)  Basics  of  Rock  Fracture  Mechanics,  Rock 
choenert,  K.  (1972)  Role  of  Fracture  Physics  in  Understanding  Comminution 
of  Coal  Particles,  Instiution  of  Chemical  Engineers  EME  Symposium  Series  69, 
POWTECH  83  -  9
th
  International  Powder  Bulk  Solids  Technical  Exhibit  an
C
 
Pook,  L.P.  (2000)  Linear  Elastic  Fracture  Mechanics  for  Engineers:  Theory  an
A
 
Prasher,  C.L.  (1987)  Crushing  and  Grinding  Process  Handbook,  John  Wiley  and  Sons, 
New York. 
 
Rimmer,  H.W.,  Swaroop,  S.,  Flavel,  M.D.,  Leverance,  N.C.  (1986)  Equipment  Sizing 
and  Process  Design  Procedures  for  Crushing  and  Screening  Circuits,  Adv
M
 
von Rittenger, P.R. (1867) Lehrbuch der Aufbereitungskonde, Berlin. 
 
R
H.M.S.O, London. 
 
Rumpf, H. (1966) Auferbereitungs-Technik, Vol. 7, pp. 421. 
 
Santurbao,  R.  and  Fairhurst,  C.  (1991)  Fracture  Mechanics  In  the  Context  of  Rock 
Crushing:  preliminary  Experimental  Results  Concerning  th
S
Symposium (edited by Roegiers), Balkema, Rotterdam, pp. 441-449 
 
S
and Metallurgical Processing, Vol. 11, No. 2, pp. 80-86.  
 
S
and  Fracture  toughness  Testing,  Proceedings  of  the  21
st
  U.S.  Symposium  on  Rock 
Mechanics, pp. 581-590.  
 
S
Fracture Mechanics, ch. 1, (edited by H.P. Rossmanith), Springer-Verlag, New York, pp. 
2-29. 
 
S
Phenomena, Transactions SME/AIME, Vol. 252, pp. 21-26. 
 
  154
Schoenert,  K.  (1979)  Aspects  of  the  Physics  of  Breakage  Relevant  to  Comminution,  4
th
 
, K. (1987) Advances in the Physical Fundamentals of Comminution, Advances 
 Mineral Processing, ch. 1, (edited by Somasundaran), SME, Littleton, CO, pp. 19-32. 
 pp. 487-494. 
, pp. 303-317. 
d Solid Mechanics - Lehigh University. 
un,  Z.  and  Ouchterlony,  F.  (1986)  Fracture  Toughness  of  Stripa  Granite  Cores, 
vensson, A. and Steer, J.F. (1990) New Cone Crusher Technology and Developments In 
sis  of  Cracks  Handbook, 
SME Press, New York. 
, Elsevier, New York. 
tp://www.mme.state.va.us/dmr/DOCS/Geol/vageo.html
Tewksbury Symposium, Melbourne, pp. 3.1-3.30. 
 
Schoenert
in
 
Schoenert,  K.  (1991)  Advances  in  Comminution  Fundamentals  and  Impacts  on 
Technology, Aufbereitungs Technik, Vol,. 32, No. 9,
 
Shockey, D., Curran, D., Seaman, L., Rosenberg, J, and Ptersen, C. (1974) Fragmentation 
of  Rock  Under  Load,  International  Journal  of  Rock  Mechanics  Science  and 
Geoemachanics Abstracts, Vol. 11
 
Sih,  G.C.  (1973)  Handbook  of  Stress  Intensity  Factors  for  Researchers  and  Engineers, 
Institute of Fracture an
 
Stansell, B. (2003) Private communication. 
 
S
International Journal of Rock Mechanics, Mining Sciences, and Geomechanics Abstracts, 
Vol. 23, no. 6, pp. 399-409. 
 
S
Comminution Circuits, Minerals Engineering, Vol. 3, No. 1-2, pp. 83-103. 
 
Tada,  H.,  Paris,  P.C.,  and  Irwin,  G.R.  (2000)  The  Stress  Analy
A
 
Tkacova, K. (1989) Mechanical Activation of Minerals
 
Virginia  Department  of  Mines,  Minerals,  and  Energy  (DMME)  (2003)  Division  of 
Mineral Resources, ht . 
Ing Vet Akad, Vol. 
51, pp. 1-39. 
asuring 
article  Breakage,  Proceedings  of  the  1
st
  World  Congress  on  Particle  Technology, 
eiss, N.L. (1985) Jaw Crushers, SME Mineral Processing Handbook, ch. 3B-1, (edited 
by N.L. Weiss), SME/AIME, New York. 
 
Wagner, W.J., (1990) Rating Crushers By Formula, Rock Products, February, pp. 48-51. 
 
Weibull, W. (1939) A Statistical Theory of the Strength of Materials, 
1
 
Weichert,  R.  and  Herbst,  J.A.  (1986)  An  Ultrafast  load  Cell  Device  for  Me
P
Nurnber, pp. 3-14.  
 
W
  155
 
Whiten,  W.J.  (1972)  Simulation  and  Model  Building  for  Mineral  Processing,  PhD 
issertation - University of Queensland, Australia. 
 Methods in the Mineral Industry, Johannesburg, pp. 317-323. 
Control  84 
dited by J.A. Herbst), SME, pp. 217-224. 
hittaker, B.N., Singh, R.N., and Sun, G. (1992) Rock Fracture Mechanics  Principles, 
ills, B.A. (1992) Mineral Processing Technology, Pergamon Press, New York. 
Tohoku University, No. 1, 
p. 116-133. 
D
 
Whiten,  W.J.  (1972a)  The  Simulation  of  Crushing Plants with models Developed Using 
Multiple  Spline  Regression,  The  10
th
  International  Symposium  on  the  Application  of 
Computer
 
Whiten  ,  W.J.  (1984)  Models  and  Control  Techniques  for  Crushing  Plants, 
(e
 
W
Design and Applications, Elsevier, New York. 
 
W
 
Yashima,  S.,  Morohashi,  S.,  and  Saito,  F.  (1979)  Single  Particel  Crushing  Under  Slow 
Rate of Loading, Scientific Reports of the Research Institute  
p
  156
 
 
 
 
 
 
 
 
APPENDIX I 
 
 
 
 
 
 
 
 
 
 
 
END RESULTS: FRACTURE TOUGHNESS 
  157
D t a F
v
F
w
K
Ic
29
BG-W-2   47.447   25.095   21.920   367   367   0.657
BG-W-3   47.473   26.289   21.920   416   416   0.709
BG-W-5   47.498   25.502   22.047   289   289   0.511
23   0.042
9   0.195
BG-W-10   47.447   24.841   22.123   294   293   0.536
AVG  0.601
STDEV  0.083
Did not fracture at notch
Table A1.1 Fracture toughness results for Boscobel granite
mm   mm   mm   N   N
  MPa m
1/2
BG-W-1   47.498   25.375   21.920   73   73   0.1
Specimen
BG-W-4   47.473   24.511   21.920   48   48   0.088
BG-W-6   47.498   24.790   22.123   23
BG-W-7   47.473   25.603   21.971   131   131   0.229
BG-W-8   47.523   26.416   22.123   344   344   0.589
BG-W-9   47.498   25.121   21.996   109   10
  158
Requirement t
Specimen
a D D - a Size of FPZ
mm   mm   mm   mm   mm   mm
BG-W-3   18.871   OK   OK   OK   OK   0.301
0.156
BG-W-6   0.067   OK   OK   OK   OK   0.001
BG-W-8   13.027   OK   OK   OK   OK   0.207
023
BG-W-10   10.772   OK   OK   OK   OK   0.172
Table A1.2 Specimen size requirements for Boscobel granite
BG-W-1   0.628   OK   OK   OK   OK   0.010
BG-W-2   16.157   OK   OK   OK   OK   0.257
BG-W-4   0.289   OK   OK   OK   OK   0.005
BG-W-5   9.792   OK   OK   OK   OK
BG-W-7   1.974   OK   OK   OK   OK   0.031
BG-W-9   1.426   OK   OK   OK   OK   0.
  159
D t a F
v
F
w
K
Ic
2
CG-W-3   47.574   25.527   20.599   862   861   1.404
CG-W-4   47.574   25.044   20.701   805   804   1.343
CG-W-5   47.625   25.248   20.701   1028   1027   1.699
CG-W-6   47.574   25.070   20.599   1044   1043   1.732
CG-W-7   47.650   25.298   20.726   1016   1014   1.676
CG-W-8   47.625   24.943   20.726   762   761   1.277
CG-W-9   47.574   25.146   20.726   935   934   1.556
CG-W-10   47.574   24.232   20.650   799   798   1.375
AVG   1.454
STDEV   0.195
Table A1.3 Fracture toughness results for Culpeper gray siltstone
mm   mm   mm   N   N
  MPa m
1/
Specimen
CG-W-1   47.600   25.121   20.676   749   748   1.243
CG-W-2   47.600   24.790   20.574   736   735   1.231
  160
Requirement t D D - a Size of FPZ
mm   mm   mm   mm   mm   mm
CG-W-1   8.035   OK   OK   OK   1.023
CG-W-2   7.876   OK   OK   OK   OK   1.003
CG-W-3   1.304
CG-W-4   9.380   OK   OK   OK   OK   1.194
CG-W-5   15.008   OK   OK   OK   OK   1.911
CG-W-6   15.587   OK   OK   OK   OK   1.985
CG-W-7   14.602   OK   OK   OK   OK   1.859
CG-W-8   8.474   OK   OK   OK   OK   1.079
CG-W-9   12.582   OK   OK   OK   OK   1.602
CG-W-10   9.824   OK   OK   OK   OK   1.251
Specimen
Table A1.4 Specimen size requirements for Culpeper gray siltstone
a
OK
10.242   OK   OK   OK   OK
  161
  162
 
D t a F
v
F
w
K
Ic
mm   mm   mm   N   N
  MPa m
1/2
CR-W-1   47.625   24.130   22.250   863   862   1.623
CR-W-2   47.600   25.451   22.200   686   685   1.221
CR-W-3   47.600   28.550   22.885   792   791   1.307
CR-W-4   47.574   24.917   22.250   958   957   1.748
CR-W-5   47.549   25.578   22.225   1043   1041   1.853
CR-W-6   47.523   24.460   22.276   823   822   1.535
CR-W-7   47.625   24.943   22.225   921   919   1.673
CR-W-8   47.625   25.451   22.250   558   557   0.995
CR-W-9   47.625   23.063   22.225   734   733   1.442
CR-W-10   47.549   25.121   22.098   897   895   1.611
AVG   1.501
STDEV   0.262
Specimen
Table A1.5 Fracture toughness results for Culpeper red siltstone
  163
 
Requirement t a D D - a
Size of FPZ
mm   mm   mm   mm   mm   mm
CR-W-1   14.373   OK   OK   OK   OK   0.610
CR-W-2   8.131   OK   OK   OK   OK   0.345
CR-W-3   9.318   OK   OK   OK   OK   0.395
CR-W-4   16.679   OK   OK   OK   OK   0.708
CR-W-5   18.726   OK   OK   OK   OK   0.795
CR-W-6   12.853   OK   OK   OK   OK   0.546
CR-W-7   15.271   OK   OK   OK   OK   0.648
CR-W-8   5.398   OK   OK   OK   OK   0.229
CR-W-9   11.349   OK   OK   OK   OK   0.482
CR-W-10   14.156   OK   OK   OK   OK   0.601
Specimen
Table A1.6 Specimen size requirements for Culpeper red siltstone
  164
 
D t a F
v
F
w
K
Ic
mm   mm   mm   N   N
  MPa m
1/2
LD-W-1   47.600   26.060   21.768   588   636   1.081
LD-W-2   47.574   25.527   21.768   650   704   1.222
LD-W-3   47.574   25.908   21.742   754   816   1.395
LD-W-4   47.625   24.841   21.742   682   738   1.313
LD-W-5   47.625   25.730   21.793   667   722   1.243
LD-W-6   47.600   25.959   21.742   753   814   1.387
LD-W-7   47.574   24.968   21.742   604   654   1.159
LD-W-8   47.625   26.441   21.819   679   734   1.233
LD-W-9   47.574   26.238   21.768   724   784   1.324
LD-W-10   47.600   26.086   21.742   647   700   1.186
AVG   1.254
STDEV   0.101
Specimen
Table A1.7 Fracture toughness results for Leesburg Diabase
Requirement t a D D - a Size of FPZ
mm   mm   mm   mm   mm   mm
LD-W-1   9.435   OK   OK   OK   OK   0.400
LD-W-2   12.048   OK   OK   OK   OK   0.511
LD-W-3   15.696   OK   OK   OK   OK   0.666
LD-W-4   13.917   OK   OK   OK   OK   0.591
LD-W-5   12.473   OK   OK   OK   OK   0.529
LD-W-6   15.533   OK   OK   OK   OK   0.659
LD-W-7   10.835   OK   OK   OK   OK   0.460
LD-W-8   12.259   OK   OK   OK   OK   0.520
LD-W-9   14.141   OK   OK   OK   OK   0.600
LD-W-10   11.359   OK   OK   OK   OK   0.482
Specimen
Table A1.8 Specimen size requirements for Leesburg diabase
  165
 
D t a F
v
F
w
K
Ic
mm   mm   mm   N   N
  MPa m
1/2
SMB-W-1   47.447   24.486   21.920   710   709   1.300
SMB-W-2   47.447   24.409   21.946   750   749   1.380
SMB-W-3   47.447   25.197   21.844   1221   1219   2.163
SMB-W-4   47.447   24.105   21.920   1274   1272   2.370
SMB-W-5   47.447   24.917   21.996   729   727   1.317
SMB-W-6   47.396   24.232   21.946   576   575   1.069
SMB-W-7   47.346   24.892   21.946   930   928   1.683
SMB-W-8   47.422   24.917   21.996   812   811   1.469
SMB-W-9   47.422   24.816   21.971   914   912   1.657
SMB-W-10   47.396   24.994   21.971   983   982   1.772
AVG   1.411
STDEV   0.215
Specimen
Failed to meet specimen size reqiurements
Table A1.9 Fracture toughness results for Shadwell metabasalt
  166
 
Requirement t a D D - a Size of FPZ
mm   mm   mm   mm   mm   mm
SMB-W-1   12.166   OK   OK   OK   OK   0.376
SMB-W-2   13.701   OK   OK   OK   OK   0.424
SMB-W-3   33.672 NO NO   OK NO   1.041
SMB-W-4   40.425 NO NO   OK NO   1.250
SMB-W-5   12.481   OK   OK   OK   OK   0.386
SMB-W-6   8.220   OK   OK   OK   OK   0.254
SMB-W-7   20.396   OK   OK   OK   OK   0.631
SMB-W-8   15.539   OK   OK   OK   OK   0.480
SMB-W-9   19.767   OK   OK   OK   OK   0.611
SMB-W-10   22.601   OK NO   OK   OK   0.699
Specimen
Table A1.10 Specimen size requirements for Shadwell metabasalt
  167
 
D t a F
v
F
w
K
Ic
mm   mm   mm   N   N
  MPa m
1/2
SG-W-1   47.473   25.806   21.971   554   553   0.965
SG-W-2   47.447   25.375   21.971   485   484   0.859
SG-W-3   47.447   25.273   21.895   608   608   1.078
SG-W-4   47.473   25.095   22.022   531   530   0.954
SG-W-5   47.498   24.740   21.971   400   400   0.726
SG-W-6   47.473   24.765   22.022   279   279   0.508
SG-W-7   47.371   25.248   21.920   392   392   0.699
SG-W-8   47.498   25.222   22.123   541   540   0.971
SG-W-9   47.473   25.222   21.946   479   478   0.851
SG-W-10   47.523   25.222   22.047   460   459   0.821
AVG   0.843
STDEV   0.165
Specimen
Table A1.11 Fracture toughness results for Spotsylvania granite
  168
  169
 
Requirement t a D D - a
Size of FPZ
mm   mm   mm   mm   mm   mm
SG-W-1   14.533   OK   OK   OK   OK   0.617
SG-W-2   11.525   OK   OK   OK   OK   0.489
SG-W-3   18.144   OK   OK   OK   OK   0.770
SG-W-4   14.194   OK   OK   OK   OK   0.602
SG-W-5   8.235   OK   OK   OK   OK   0.350
SG-W-6   4.032   OK   OK   OK   OK   0.171
SG-W-7   7.621   OK   OK   OK   OK   0.323
SG-W-8   14.714   OK   OK   OK   OK   0.624
SG-W-9   11.314   OK   OK   OK   OK   0.480
SG-W-10   10.528   OK   OK   OK   OK   0.447
Specimen
Table A1.12 Specimen size requirements for Spotsylvania granite
 
D t a F
v
F
w
K
Ic
mm   mm   mm   N   N
  MPa m
1/2
TG-W-1   47.193   26.162   21.285   636   619   1.035
TG-W-2   47.142   25.883   21.311   654   636   1.079
TG-W-3   47.168   25.756   21.234   829   807   1.368
TG-W-4   47.168   25.654   21.234   666   648   1.103
TG-W-5   47.168   25.806   21.234   813   791   1.339
TG-W-6   47.219   24.917   21.184   589   573   1.000
TG-W-7   47.092   24.460   21.184   592   576   1.029
TG-W-8   47.142   25.603   21.234   609   593   1.012
TG-W-9   47.168   25.959   21.260   657   640   1.077
TG-W-10   47.142   25.679   21.209   622   605   1.028
AVG   1.045
STDEV   0.037
Specimen
Failed to meet specimen size reqiurements
Table A1.13 Fracture toughness results for Thornberg granite
  170
Requirement t a D D - a
Size of FPZ
mm   mm   mm   mm   mm   mm
TG-W-1   14.998   OK   OK   OK   OK   0.637
TG-W-2   16.296   OK   OK   OK   OK   0.692
TG-W-3   26.194 NO NO   OK NO   1.112
TG-W-4   17.052   OK   OK   OK   OK   0.724
TG-W-5   25.093   OK NO   OK   OK   1.065
TG-W-6   13.991   OK   OK   OK   OK   0.594
TG-W-7   14.827   OK   OK   OK   OK   0.629
TG-W-8   14.345   OK   OK   OK   OK   0.609
TG-W-9   16.257   OK   OK   OK   OK   0.690
TG-W-10   14.807   OK   OK   OK   OK   0.628
Specimen
Table A1.14 Specimen size requirements for Thornburg granite
  171
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
APPENDIX II 
 
HECT RESULTS: SPECIFIC COMMINUTION ENERGY 
  172
D t Mass F E T P
mm   mm   g   kN   pts   MPa   kWh/t
BG-E-1-1   47.473   25.654   115.3   10.898   42.665   5.697   0.103
BG-E-1-2   47.473   25.603   116.3   14.887   67.242   7.798   0.161
BG-E-1-3   47.473   25.984   123.4   35.993   8.889   18.576   0.020
BG-E-1-4   47.422   25.984   123.4   41.936   87.393   21.666   0.197
BG-E-1-5   47.473   25.705   115.3   18.975   39.197   9.899   0.094
BG-E-1-6   47.498   26.035   123.4   43.693   7.809   22.493   0.018
BG-E-1-7   47.473   25.832   117.4   24.909   0.463   12.931   0.001
BG-E-1-8   47.549   25.552   117.4   19.073   3.205   9.994   0.008
BG-E-1-9   47.447   25.806   122.4   51.281   70.323   26.662   0.160
BG-E-1-10   47.574   25.324   113.3   15.764   46.524   8.330   0.114
BG-E-1-11   47.473   25.502   119.4   17.516   59.564   9.211   0.139
BG-E-1-12   47.549   25.070   111.3   13.233   18.058   7.067   0.045
BG-E-1-13   47.549   25.781   117.4   17.125   59.472   8.893   0.141
BG-E-1-14   47.473   25.171   115.3   24.326   82.403   12.960   0.198
BG-E-1-15   47.498   25.730   113.3   12.067   24.360   6.286   0.060
BG-E-1-16   47.473   25.654   114.3   21.604   85.755   11.293   0.208
BG-E-1-17   47.498   25.375   115.3   21.697   61.704   11.461   0.149
BG-E-1-18   47.498   25.806   118.4   19.269   46.797   10.008   0.110
BG-E-1-19   47.498   25.756   113.3   16.738   32.339   8.710   0.079
BG-E-1-20   47.473   26.035   114.3   15.666   38.749   8.069   0.094
AVG 11.900 0.105
STDEV 5.840 0.065
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  284
THROW =   50.80 mm
SET PT =   410
CSS =   31.75 mm
P = Mass specific crushing energy
Table AII.1 HECT results for Boscobel granite at reduction ratio 1
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Specimen
D = Specimen diameter
t = Specimen thickness
  173
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
BG-E-2-1   47.523   25.654   110.3   11.676   42.032   6.097   0.106
BG-E-2-2   47.295   26.543   115.3   14.207   88.610   7.205   0.213
BG-E-2-3   47.371   26.441   113.3   9.634   13.547   4.897   0.033
BG-E-2-4   47.320   25.756   111.3   17.418   92.112   9.098   0.230
BG-E-2-5   47.447   26.162   120.4   23.379   155.383   11.990   0.359
BG-E-2-6   47.498   26.187   120.4   24.615   133.220   12.598   0.307
BG-E-2-7   47.473   25.705   116.3   15.470   63.534   8.071   0.152
BG-E-2-8   47.422   26.060   117.4   12.748   135.602   6.567   0.321
BG-E-2-9   47.523   25.502   115.3   12.748   57.332   6.696   0.138
BG-E-2-10   47.523   26.111   119.4   23.641   39.300   12.129   0.091
BG-E-2-11   47.473   25.705   117.4   20.238   75.158   10.558   0.178
BG-E-2-12   47.523   26.416   120.4   17.027   145.752   8.635   0.336
BG-E-2-13   47.523   25.781   119.4   34.957   72.519   18.164   0.169
BG-E-2-14   47.523   25.908   118.4   25.594   72.192   13.233   0.169
BG-E-2-15   47.523   26.162   120.4   27.151   135.790   13.902   0.313
BG-E-2-16   47.498   26.543   120.4   25.496   196.590   12.874   0.454
BG-E-2-17   47.447   26.416   121.4   21.697   165.771   11.021   0.379
BG-E-2-18   47.498   25.933   119.4   23.739   91.050   12.269   0.212
BG-E-2-19   47.473   26.314   120.4   19.073   122.132   9.720   0.282
BG-E-2-20   47.396   25.527   113.3   9.243   60.464   4.863   0.148
AVG 10.029 0.230
STDEV 3.456 0.110
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  224
THROW =   50.80 mm
SET PT =   406
CSS =   16.00 mm
P = Mass specific crushing energy
Table AII.2 HECT results for Boscobel granite at reduction ratio 2
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Specimen
D = Specimen diameter
t = Specimen thickness
  174
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
CG-E-1-1   47.625   26.822   129.8   65.875   146.230   32.830   0.313
CG-E-1-2   47.727   26.492   128.8   52.740   NA   26.555   NA
CG-E-1-3   47.574   26.721   128.8   57.117   157.110   28.604   0.339
CG-E-1-4   47.600   25.832   122.7   50.792   93.486   26.298   0.212
CG-E-1-5   47.600   26.111   123.7   70.741   174.160   36.234   0.391
CG-E-1-6   47.650   26.695   127.8   53.349   136.560   26.700   0.297
CG-E-1-7   47.600   26.949   128.8   46.028   147.640   22.843   0.318
CG-E-1-8   47.625   26.695   127.8   53.518   153.410   26.799   0.334
CG-E-1-9   47.600   26.518   127.8   53.714   104.140   27.091   0.226
CG-E-1-10   47.625   26.695   129.8   47.389   137.093   23.729   0.293
CG-E-1-11   47.625   26.899   130.8   65.292   42.735   32.447   0.091
CG-E-1-12   47.625   26.314   126.8   57.802   52.096   29.362   0.114
CG-E-1-13   47.625   26.848   130.8   52.157   80.812   25.969   0.172
CG-E-1-14   47.625   27.229   131.8   53.714   236.690   26.370   0.499
CG-E-1-15   47.625   26.594   129.8   55.560   NA   27.927   NA
CG-E-1-16   47.600   26.924   128.8   52.059   219.860   25.860   0.474
CG-E-1-17   47.650   26.721   128.8   55.075   198.612   27.537   0.428
CG-E-1-18   47.650   26.721   129.8   54.879   123.510   27.439   0.264
CG-E-1-19   47.625   26.137   126.8   50.787   182.050   25.975   0.399
CG-E-1-20   47.600   26.111   125.8   46.219   41.576   23.674   0.092
AVG 27.512 0.292
STDEV 3.219 0.123
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  284
THROW =   50.80 mm
SET PT =   410
CSS =   31.75 mm
P = Mass specific crushing energy
Specimen
D = Specimen diameter
t = Specimen thickness
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Table AII.3 HECT results for Culpeper gray siltstone at reduction ratio 1
  175
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
CG-E-2-1   47.650   24.892   121.4   45.152   177.472   24.234   0.406
CG-E-2-2   47.600   25.933   126.4   45.152   105.540   23.286   0.232
CG-E-2-3   47.650   25.425   122.4   45.632   205.113   23.978   0.466
CG-E-2-4   47.650   25.578   124.4   65.479   146.622   34.202   0.327
CG-E-2-5   47.650   25.705   124.4   46.415   170.931   24.124   0.382
CG-E-2-6   47.600   24.587   119.4   50.009   196.981   27.203   0.458
CG-E-2-7   47.625   25.730   124.4   53.323   188.210   27.702   0.420
CG-E-2-8   47.600   25.730   125.4   46.708   140.070   24.279   0.310
CG-E-2-9   47.625   25.197   122.4   49.235   244.950   26.120   0.556
CG-E-2-10   47.600   24.511   118.4   47.776   131.061   26.069   0.308
CG-E-2-11   47.625   25.552   123.4   51.183   102.832   26.776   0.232
CG-E-2-12   47.574   25.375   123.4   50.890   116.780   26.837   0.263
CG-E-2-13   47.600   26.162   227.050   26.016   0.499
CG-E-2-14   47.600   25.781   125.4   43.497   142.160   22.565   0.315
CG-E-2-15   25.375   123.4   .554   158.700   25.578   0.357
CG-E-2-16   0.552
CG-E-2-17   47.625   25.400   123.4   52.936   46.233   27.859   0.104
CG-E-2-18   47.650   27.559   132.4   60.230   153.372   29.199   0.322
CG-E-2-19   47.625   26.060   126.4   61.200   139.650   31.392   0.307
CG-E-2-20   47.650   25.171   122.4   44.275   81.963   23.500   0.186
AVG 26.286 0.350
STDEV 2.854 0.119
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  224
THROW =   50.80 mm
SET PT =   406
CSS =   16.00 mm
P = Mass specific crushing energy
Specimen
D = Specimen diameter
t = Specimen thickness
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Table AII.4 HECT results for Culpeper gray siltstone at reduction ratio 2
126.4   50.890
47.625   48
47.600   25.781   125.4   47.825   249.290   24.810
  176
  177
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
CR-E-1-1   47.574   26.848   131.4   55.707   183.590   27.766   0.388
CR-E-1-2   47.600   25.908   125.4   55.756   142.040   28.783   0.315
CR-E-1-3   47.625   26.289   127.4   61.204   147.412   31.121   0.321
CR-E-1-4   47.625   26.187   127.4   60.146   90.996   30.701   0.198
CR-E-1-5   47.625   26.238   127.4   71.911   92.529   36.636   0.202
CR-E-1-6   47.650   26.721   128.4   56.632   33.229   28.316   0.072
CR-E-1-7   47.549   27.051   131.4   54.395   92.853   26.922   0.196
CR-E-1-8   47.650   26.035   125.4   52.446   34.285   26.914   0.076
CR-E-1-9   47.650   26.670   128.4   45.828   72.848   22.957   0.158
CR-E-1-10   47.650   26.848   130.4   59.843   56.112   29.780   0.120
CR-E-1-11   47.650   24.740   121.4   55.164   154.000   29.790   0.352
CR-E-1-12   47.625   26.949   130.4   70.149   68.577   34.795   0.146
CR-E-1-13   47.676   25.908   126.4   52.446   142.793   27.031   0.314
CR-E-1-14   47.574   27.305   130.4   57.215   302.560   28.040   0.645
CR-E-1-15   47.650   25.883   125.4   48.070   192.811   24.813   0.427
CR-E-1-16   47.701   25.298   123.4   66.751   201.000   35.214   0.453
CR-E-1-17   47.600   26.619   130.4   48.848   172.720   24.543   0.368
CR-E-1-18   47.650   25.730   125.4   58.869   NA   30.567   NA
CR-E-1-19   47.650   26.822   130.4   58.994   240.730   29.385   0.513
CR-E-1-20   47.650   26.035   125.4   46.219   180.122   23.718   0.399
AVG 28.890 0.298
STDEV 3.691 0.156
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  284
THROW =   50.80 mm
SET PT =   410
CSS =   31.75 mm
P = Mass specific crushing energy
Specimen
D = Specimen diameter
t = Specimen thickness
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Table AII.5 HECT results for Culpeper red siltstone at reduction ratio 1
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
CR-E-2-1   47.600   25.629   123.4   63.055   61.943   32.906   0.139
CR-E-2-2   47.600   25.171   122.4   57.606   65.419   30.608   0.149
CR-E-2-3   47.523   26.416   127.4   52.157   169.870   26.450   0.370
CR-E-2-4   47.600   26.162   126.4   24.424   158.400   12.486   0.348
CR-E-2-5   47.574   25.654   124.4   54.782   19.440   28.575   0.043
CR-E-2-6   47.625   25.832   125.4   62.374   NA   32.277   NA
CR-E-2-7   47.676   25.959   128.4   38.924   142.820   20.023   0.309
CR-E-2-8   47.574   26.340   127.4   63.344   214.040   32.181   0.467
CR-E-2-9   47.600   25.121   121.4   56.534   43.105   30.099   0.099
CR-E-2-10   47.625   25.629   126.4   46.121   189.552   24.056   0.417
CR-E-2-11   47.701   25.527   125.4   49.044   214.000   25.641   0.474
CR-E-2-12   47.625   24.994   123.4   51.570   88.018   27.581   0.198
CR-E-2-13   47.549   25.883   126.4   47.287   305.770   24.461   0.672
CR-E-2-14   47.574   25.400   122.4   46.415   195.310   24.453   0.443
CR-E-2-15   47.625   25.679   126.4   43.884   NA   22.844   NA
CR-E-2-16   47.625   25.883   128.4   55.756   76.973   28.796   0.167
CR-E-2-17   47.523   26.340   127.4   51.179   317.941   26.029   0.693
CR-E-2-18   47.650   25.095   124.4   46.415   34.331   24.710   0.077
CR-E-2-19   47.574   26.060   125.4   61.787   174.622   31.727   0.387
CR-E-2-20   47.650   26.035   127.4   43.497   144.390   22.321   0.315
AVG 26.411 0.320
STDEV 4.932 0.192
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  224
THROW =   50.80 mm
SET PT =   406
CSS =   16.00 mm
P = Mass specific crushing energy
Specimen
D = Specimen diameter
t = Specimen thickness
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Table AII.6 HECT results for Culpeper red siltstone at reduction ratio 2
  178
  179
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
LD-E-1-1   47.752   25.781   138.4   50.987   173.160   26.366   0.348
LD-E-1-2   47.803   27.229   145.4   48.750   220.864   23.844   0.422
LD-E-1-3   47.549   27.203   145.4   51.668   170.274   25.430   0.325
LD-E-1-4   47.574   26.975   144.4   57.290   238.110   28.420   0.458
LD-E-1-5   47.574   27.381   145.4   46.219   150.700   22.588   0.288
LD-E-1-6   47.574   26.594   140.4   53.518   183.330   26.930   0.363
LD-E-1-7   47.549   26.746   142.4   61.302   300.290   30.687   0.586
LD-E-1-8   47.625   27.000   144.4   59.794   NA   29.603   NA
LD-E-1-9   47.625   26.568   141.4   59.794   NA   30.084   NA
LD-E-1-10   47.625   26.899   143.4   45.272   NA   22.498   NA
LD-E-1-11   47.600   26.873   142.4   50.111   118.630   24.940   0.231
LD-E-1-12   47.574   27.381   146.4   49.920   213.943   24.397   0.406
LD-E-1-13   47.650   26.949   144.4   59.648   190.881   29.570   0.367
LD-E-1-14   47.600   27.280   146.4   62.957   341.770   30.866   0.648
LD-E-1-15   47.523   26.772   143.4   47.776   15.401   23.906   0.030
LD-E-1-16   47.549   27.127   144.4   68.308   343.380   33.714   0.660
LD-E-1-17   47.625   26.975   145.4   53.420   332.040   26.472   0.634
LD-E-1-18   47.600   26.975   144.4   55.271   122.160   27.404   0.235
LD-E-1-19   47.600   27.229   145.4   58.189   233.731   28.582   0.446
LD-E-1-20   47.625   27.178   144.4   45.734   14.111   22.494   0.027
LD-E-1-21   47.193   23.089   107.3   26.274   93.826   15.351   0.243
LD-E-1-22   47.523   22.454   121.4   46.317   141.230   27.633   0.323
AVG 26.445 0.371
STDEV 3.929 0.182
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  284
THROW =   50.80 mm
SET PT =   410
CSS =   31.75 mm
P = Mass specific crushing energy
Table AII.7 HECT results for Leesburg diabase at reduction ratio 1
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Specimen
D = Specimen diameter
t = Specimen thickness
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
LD-E-2-1   47.600   25.527   137.4   48.946   67.404   25.644   0.136
LD-E-2-2   47.625   26.137   139.4   46.820   NA   23.946   NA
LD-E-2-3   47.625   26.035   140.4   NA   NA   NA   NA
LD-E-2-4   47.549   25.959   138.4   48.167   219.020   24.843   0.440
LD-E-2-5   47.625   26.391   141.4   52.082   NA   26.380   NA
LD-E-2-6   47.600   26.340   140.4   37.826   NA   19.207   NA
LD-E-2-7   47.574   26.187   139.4   47.193   290.021   24.116   0.578
LD-E-2-8   47.625   26.035   138.4   49.528   304.533   25.430   0.611
LD-E-2-9   47.625   26.162   139.4   65.973   326.283   33.708   0.650
LD-E-2-10   47.549   26.238   140.4   55.075   240.430   28.104   0.476
LD-E-2-11   47.625   26.365   141.4   63.735   135.390   32.314   0.266
LD-E-2-12   47.600   26.264   140.4   52.740   211.564   26.857   0.419
LD-E-2-13   47.625   25.857   138.4   50.013   251.110   25.855   0.504
LD-E-2-14   47.650   26.060   139.4   47.487   76.479   24.345   0.152
LD-E-2-15   47.574   26.137   139.4   47.585   253.400   24.363   0.505
LD-E-2-16   47.676   25.375   135.4   48.265   203.254   25.399   0.417
LD-E-2-17   47.625   26.518   141.4   53.714   299.070   27.077   0.587
LD-E-2-18   47.625   25.603   136.4   40.188   236.360   20.982   0.481
LD-E-2-19   47.625   25.857   137.4   47.678   344.290   24.648   0.696
LD-E-2-20   47.549   25.705   137.4   49.044   355.170   25.545   0.718
AVG 25.724 0.477
STDEV 3.275 0.174
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  224
THROW =   50.80 mm
SET PT =   406
CSS =   16.00 mm
P = Mass specific crushing energy
Table AII.8 HECT results for Leesburg diabase at reduction ratio 2
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Specimen
D = Specimen diameter
t = Specimen thickness
  180
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
SMB-E-1-1   47.447   26.289   132.4   52.059   217.710   26.570   0.457
SMB-E-1-2   47.473   25.629   122.4   46.900   96.723   24.540   0.220
SMB-E-1-3   47.549   25.883   124.4   44.652   201.453   23.098   0.450
SMB-E-1-4   47.574   25.705   121.4   48.542   198.755   25.270   0.455
SMB-E-1-5   47.422   26.010   131.4   46.121   200.700   23.805   0.424
SMB-E-1-6   47.371   26.314   131.4   33.389   275.846   17.052   0.583
SMB-E-1-7   47.473   25.756   124.4   53.445   226.807   27.827   0.507
SMB-E-1-8   47.498   26.086   132.4   45.948   231.300   23.608   0.485
SMB-E-1-9   47.523   25.832   124.4   42.224   215.783   21.897   0.482
SMB-E-1-10   47.447   26.060   131.4   45.152   187.250   23.247   0.396
SMB-E-1-11   47.498   25.578   121.4   44.641   145.765   23.392   0.334
SMB-E-1-12   47.523   25.781   122.4   45.126   187.956   23.448   0.427
SMB-E-1-13   47.473   25.654   120.4   40.068   200.047   20.945   0.462
SMB-E-1-14   47.549   25.425   120.4   30.358   120.210   15.986   0.277
SMB-E-1-15   47.473   25.984   123.4   41.353   129.331   21.342   0.291
SMB-E-1-16   47.473   26.035   132.4   46.117   216.691   23.754   0.455
SMB-E-1-17   47.422   26.111   131.4   45.632   196.410   23.461   0.415
SMB-E-1-18   47.473   25.781   121.4   24.037   54.397   12.503   0.125
SMB-E-1-19   47.473   26.162   124.4   43.688   112.980   22.394   0.252
SMB-E-1-20   47.396   26.010   131.4   51.962   225.980   26.834   0.478
SMB-E-1-21   47.498   25.171   127.4   33.280   161.384   17.721   0.352
SMB-E-1-22   47.473   26.035   132.4   58.674   271.380   30.222   0.569
SMB-E-1-23   47.473   26.010   132.4   54.003   223.073   27.843   0.468
AVG 22.816 0.407
STDEV 4.102 0.112
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  284
THROW =   50.80 mm
SET PT =   410
CSS =   31.75 mm
P = Mass specific crushing energy
Specimen
D = Specimen diameter
t = Specimen thickness
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Table AII.9 HECT results for Shadwell metabasalt at reduction ratio 1
  181
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
SMB-E-2-1   47.498   26.238   124.4   32.301   157.010   16.500   0.351
SMB-E-2-2   47.549   25.933   125.4   22.378   108.620   11.553   0.241
SMB-E-2-3   47.422   25.883   131.4   51.468   381.332   26.695   0.806
SMB-E-2-4   47.523   25.781   124.4   45.343   411.740   23.560   0.920
SMB-E-2-5   47.447   26.264   133.4   43.977   178.293   22.467   0.371
SMB-E-2-6   47.447   25.908   124.4   49.235   243.370   25.498   0.544
SMB-E-2-7   47.447   26.060   125.4   47.967   174.430   24.696   0.386
SMB-E-2-8   47.523   26.314   131.4   46.806   77.250   23.828   0.163
SMB-E-2-9   47.574   26.060   129.4   29.388   143.030   15.090   0.307
SMB-E-2-10   47.447   25.908   123.4   54.879   320.174   28.421   0.721
SMB-E-2-11   47.473   26.162   127.4   79.695   690.190   40.850   1.505
SMB-E-2-12   47.523   25.959   126.4   57.606   147.130   29.727   0.323
SMB-E-2-13   47.447   26.213   133.4   48.065   129.262   24.603   0.269
SMB-E-2-14   47.473   26.086   132.4   63.242   329.730   32.511   0.692
SMB-E-2-15   47.473   26.264   132.4   43.693   280.570   22.310   0.589
SMB-E-2-16   47.473   26.111   132.4   47.967   449.760   24.635   0.944
SMB-E-2-17   47.447   26.238   126.4   72.102   331.470   36.871   0.729
SMB-E-2-18   47.498   26.162   128.4   37.559   187.190   19.242   0.405
SMB-E-2-19   47.473   26.365   133.4   41.936   216.980   21.330   0.452
SMB-E-2-20   47.473   26.264   131.4   50.111   138.500   25.587   0.293
AVG 24.799 0.550
STDEV 6.885 0.323
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  224
THROW =   50.80 mm
SET PT =   406
CSS =   16.00 mm
P = Mass specific crushing energy
Specimen
D = Specimen diameter
t = Specimen thickness
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Table AII.10 HECT results for Shadwell metabasalt  at reduction ratio 2
  182
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
SG-E-1-1   47.447   26.848   131.4   31.430   167.524   15.707   0.354
SG-E-1-2   47.549   27.127   133.4   25.785   137.333   12.726   0.286
SG-E-1-3   47.473   26.670   127.4   8.562   NA   4.305   NA
SG-E-1-4   47.422   27.178   127.4   42.514   159.812   21.000   0.348
SG-E-1-5   47.422   26.492   124.4   23.641   150.570   11.980   0.336
SG-E-1-6   47.422   26.543   124.4   33.858   118.760   17.124   0.265
SG-E-1-7   47.447   24.028   114.3   24.726   96.719   13.807   0.235
SG-E-1-8   47.473   27.483   135.4   NA   NA   NA   NA
SG-E-1-9   47.473   27.051   132.4   27.827   227.321   13.795   0.477
SG-E-1-10   47.422   26.899   133.4   18.099   95.680   9.033   0.199
SG-E-1-11   47.422   25.933   120.4   25.687   33.569   13.297   0.077
SG-E-1-12   47.447   27.229   133.4   22.769   130.452   11.220   0.272
SG-E-1-13   47.447   27.203   131.4   35.900   31.810   17.707   0.067
SG-E-1-14   47.473   27.076   132.4   34.441   93.259   17.058   0.196
SG-E-1-15   47.473   26.822   127.4   12.744   30.655   6.371   0.067
SG-E-1-16   47.498   26.797   130.4   26.755   62.013   13.382   0.132
SG-E-1-17   47.447   25.375   124.4   27.827   113.524   14.714   0.254
SG-E-1-18   47.447   27.076   133.4   35.513   118.510   17.598   0.247
SG-E-1-19   47.422   27.000   126.4   21.017   85.165   10.450   0.187
SG-E-1-20   47.447   26.619   123.4   33.271   89.760   16.770   0.202
AVG 13.581 0.233
STDEV 4.126 0.108
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  284
THROW =   50.80 mm
SET PT =   410
CSS =   31.75 mm
P = Mass specific crushing energy
Specimen
D = Specimen diameter
t = Specimen thickness
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Table AII.11 HECT results for Spotsylvania granite at reduction ratio 1
  183
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
SG-E-2-1   47.422   26.340   122.4   33.596   160.440   17.123   0.364
SG-E-2-2   47.473   27.051   126.4   45.539   191.473   22.575   0.421
SG-E-2-3   47.498   26.086   129.4   26.657   169.610   13.697   0.364
SG-E-2-4   47.396   25.781   119.4   43.813   219.350   22.826   0.510
SG-E-2-5   47.447   26.391   130.4   24.037   148.560   12.221   0.316
SG-E-2-6   47.396   25.629   121.4   33.765   183.120   17.696   0.419
SG-E-2-7   47.422   26.187   128.4   28.316   184.861   14.516   0.400
SG-E-2-8   47.371   25.654   119.4   31.043   195.880   16.262   0.456
SG-E-2-9   47.396   25.883   122.4   25.105   87.930   13.028   0.200
SG-E-2-10   47.422   26.492   132.4   43.786   218.544   22.188   0.459
SG-E-2-11   47.447   25.984   124.4   29.188   122.543   15.072   0.274
SG-E-2-12   47.447   26.162   130.4   23.161   208.560   11.878   0.444
SG-E-2-13   47.396   25.248   125.4   21.893   167.030   11.647   0.370
SG-E-2-14   47.422   26.213   121.4   34.058   74.480   17.443   0.170
SG-E-2-15   47.422   25.832   121.4   25.687   126.260   13.349   0.289
SG-E-2-16   47.473   26.619   129.4   29.775   244.413   15.000   0.525
SG-E-2-17   47.473   25.629   124.4   20.238   33.833   10.590   0.076
AVG 15.712 0.356
STDEV 3.855 0.123
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  224
THROW =   50.80 mm
SET PT =   406
CSS =   16.00 mm
P = Mass specific crushing energy
Specimen
D = Specimen diameter
t = Specimen thickness
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Table AII.12 HECT results for Spotsylvania granite at reduction ratio 2
  184
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
TG-E-1-1   47.168   27.051   127.4   43.742   76.249   21.825   0.166
TG-E-1-2   47.244   26.746   125.4   32.986   111.243   16.619   0.246
TG-E-1-3   47.168   26.873   126.4   45.734   NA   22.970   NA
TG-E-1-4   47.193   26.695   124.4   24.010   NA   12.133   NA
TG-E-1-5   47.168   27.102   126.4   34.543   134.870   17.203   0.296
TG-E-1-6   47.193   27.229   128.4   46.806   89.711   23.189   0.194
TG-E-1-7   47.168   26.568   122.4   34.445   NA   17.498   NA
TG-E-1-8   47.066   27.153   127.4   29.290   150.820   14.591   0.329
TG-E-1-9   47.168   26.568   123.4   31.136   215.930   15.817   0.486
TG-E-1-10   47.142   26.416   123.4   21.213   77.514   10.844   0.175
TG-E-1-11   47.142   27.153   126.4   33.280   148.490   16.552   0.326
TG-E-1-12   47.219   26.949   126.4   27.733   76.644   13.875   0.168
TG-E-1-13   47.193   27.254   128.4   31.136   72.996   15.411   0.158
TG-E-1-14   47.219   26.949   125.4   43.203   109.430   21.614   0.242
TG-E-1-15   47.193   22.962   107.3   45.343   270.472   26.638   0.700
TG-E-1-16   47.142   27.229   127.4   48.265   192.031   23.937   0.419
TG-E-1-17   47.168   27.229   128.4   17.321   30.161   8.586   0.065
TG-E-1-18   47.168   26.899   126.4   34.156   106.990   17.139   0.235
TG-E-1-19   47.168   26.772   125.4   40.966   125.354   20.653   0.278
TG-E-1-20   47.193   26.772   126.4   29.481   65.774   14.855   0.145
AVG 17.597 0.272
STDEV 4.718 0.152
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  284
THROW =   50.80 mm
SET PT =   410
CSS =   31.75 mm
P = Mass specific crushing energy
Specimen
D = Specimen diameter
t = Specimen thickness
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Table AII.13 HECT results for Thornburg granite at reduction ratio 1
  185
 
D t W F E T P
mm   mm   g   kN   pts   MPa   kWh/t
TG-E-2-1   47.193   26.187   121.4   23.730   177.611   12.224   0.407
TG-E-2-2   47.142   26.289   122.4   23.521   222.123   12.082   0.504
TG-E-2-3   47.168   26.416   121.4   42.425   365.982   21.677   0.838
TG-E-2-4   47.168   26.264   118.4   36.976   NA   19.002   NA
TG-E-2-5   47.193   25.984   121.4   34.058   195.280   17.681   0.447
TG-E-2-6   47.168   26.010   118.4   32.502   126.730   16.866   0.297
TG-E-2-7   47.168   26.010   119.4   31.234   92.050   16.208   0.214
TG-E-2-8   47.219   25.756   117.4   29.290   108.204   15.333   0.256
TG-E-2-9   47.142   26.035   120.4   36.100   194.480   18.725   0.449
TG-E-2-10   47.193   26.365   121.4   32.399   41.661   16.577   0.095
TG-E-2-11   47.142   25.883   118.4   28.707   147.312   14.978   0.346
TG-E-2-12   47.193   26.010   120.4   24.130   NA   12.515   NA
TG-E-2-13   47.193   25.781   118.4   22.089   44.743   11.558   0.105
TG-E-2-14   47.269   26.416   121.4   26.759   103.990   13.643   0.238
TG-E-2-15   47.219   25.883   119.4   35.322   174.910   18.399   0.407
TG-E-2-16   47.219   25.705   117.4   26.857   144.524   14.087   0.342
TG-E-2-17   47.219   25.908   120.4   33.280   278.404   17.319   0.643
TG-E-2-18   47.193   26.035   120.4   36.002   139.790   18.654   0.323
TG-E-2-19   47.219   26.289   121.4   39.507   326.520   20.261   0.747
TG-E-2-20   47.219   26.060   118.4   29.775   109.230   15.404   0.256
AVG 16.160 0.384
STDEV 2.878 0.201
1 pts =   1 watt sec
FREQ =   3.8 Hz
SPAN =  224
THROW =   50.80 mm
SET PT =   406
CSS =   16.00 mm
P = Mass specific crushing energy
Specimen
D = Specimen diameter
t = Specimen thickness
M = Specimen mass
F = Peak crushing force
E = Total crushing energy
T = Tensile strength
Table AII.14 HECT results for Thornburg granite at reduction ratio 2
  186
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
APPENDIX III 
 
HECT RESULTS: BREAKAGE SIZE DISTRIBUTION 
  187
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.492   0.00   1.0000
38.100   0.00   1.0000
25.400   651.0   1713.5   1062.5   0.5466
19.100   665.0   1154.5   489.5   0.3377
12.700   632.5   1009.5   377.0   0.1769
6.350   615.5   810.5   195.0   0.0937
4.750   543.5   589.5   46.0   0.0740
3.360   506.5   545.0   38.5   0.0576
2.360   489.5   515.0   25.5   0.0467
1.700   475.0   494.5   19.5   0.0384
1.190   427.0   442.5   15.5   0.0318
0.850   433.5   446.5   13.0   0.0262
0.595   449.0   459.0   10.0   0.0220
0.425   434.5   442.5   8.0   0.0186
0.300   426.0   432.0   6.0   0.0160
0.210   450.0   456.5   6.5   0.0132
0.150   409.5   418.0   8.5   0.0096
0.105   384.5   386.0   1.5   0.0090
0.075   391.0   398.5   7.5   0.0058
pan   366.0   379.5   13.5   0.0000
Table AIII.1 Breakage distribution results for Boscobel granite at RR1
  188
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.464   0.00   1.0000
38.100   0.00   1.0000
25.400   651.0   1026.0   375.0   0.8401
19.100   665.0   903.0   238.0   0.7385
12.700   632.5   1252.5   620.0   0.4741
6.350   615.5   1209.5   594.0   0.2207
4.750   543.5   662.0   118.5   0.1702
3.360   506.5   612.0   105.5   0.1252
2.360   489.5   547.5   58.0   0.1004
1.700   475.0   521.0   46.0   0.0808
1.190   427.0   463.0   36.0   0.0655
0.850   433.5   462.5   29.0   0.0531
0.595   449.0   471.5   22.5   0.0435
0.425   434.5   453.0   18.5   0.0356
0.300   426.0   439.5   13.5   0.0299
0.210   450.0   462.5   12.5   0.0245
0.150   409.5   429.5   20.0   0.0160
0.105   384.5   388.0   3.5   0.0145
0.075   391.0   404.0   13.0   0.0090
pan   361.0   382.0   21.0   0.0000
Table AIII.2 Breakage distribution results for Boscobel granite at RR2
  189
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.624   0.00   1.0000
38.100   0.00   1.0000
25.400   650.5   1694.3   1043.8   0.5915
19.100   663.9   1077.0   413.1   0.4298
12.700   631.7   982.1   350.4   0.2927
6.350   614.3   982.0   367.7   0.1488
4.750   542.3   625.2   82.9   0.1163
3.360   505.6   585.2   79.6   0.0852
2.360   488.6   538.8   50.2   0.0655
1.700   474.4   514.4   40.1   0.0498
1.190   426.2   456.9   30.8   0.0378
0.850   432.9   458.0   25.2   0.0279
0.595   448.6   466.1   17.5   0.0211
0.425   433.6   447.5   14.0   0.0156
0.300   425.2   435.3   10.1   0.0117
0.210   449.4   457.3   7.8   0.0086
0.150   408.9   418.7   9.9   0.0048
0.105   384.0   385.4   1.4   0.0042
0.075   390.3   394.5   4.3   0.0025
pan   365.2   371.6   6.5   0.0000
Table AIII.3 Breakage distribution results for Culpeper gray siltstone at RR1
  190
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.621   0.00   1.0000
38.100   0.00   1.0000
25.400   651.0   1093.0   442.0   0.8220
19.100   665.0   594.5   0.5826
12.700   632.5   1083.0   450.5   0.4012
6.350   0.2039
4.750   3.5   634.0   90.5   0.1675
3.360   506.5   603.5   97.0   0.1284
2.360   489.5   568.5   79.0   0.0966
1.700   475.0   533.0   58.0   0.0733
1.190   427.0   474.5   47.5   0.0542
0.850   433.5   469.0   35.5   0.0399
0.595   449.0   474.5   25.5   0.0296
0.425   434.5   455.0   20.5   0.0213
0.300   426.0   441.0   15.0   0.0153
0.210   450.0   461.5   11.5   0.0107
0.150   409.5   422.0   12.5   0.0056
0.105   384.5   387.0   2.5   0.0046
0.075   391.0   397.5   6.5   0.0020
pan   365.5   370.5   5.0   0.0000
Table AIII.4 Breakage distribution results for Culpeper gray siltstone at RR2
1259.5
615.5   1105.5   490.0
54
  191
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.631   0.00   1.0000
38.100   0.00   1.0000
25.400   650.5   1710.2   1059.7   0.5845
19.100   663.9   1025.6   361.7   0.4426
12.700   631.7   908.7   277.0   0.3340
6.350   614.3   1052.1   437.8   0.1623
4.750   542.3   626.9   84.6   0.1292
3.360   505.7   593.7   88.0   0.0947
2.360   488.5   543.1   54.6   0.0733
1.700   474.2   518.1   43.9   0.0560
1.190   426.2   461.5   35.3   0.0422
0.850   432.9   460.2   27.3   0.0315
0.595   448.5   468.2   19.7   0.0238
0.425   433.7   449.8   16.1   0.0174
0.300   425.3   436.3   11.0   0.0131
0.210   449.5   457.6   8.1   0.0099
0.150   408.8   420.8   12.0   0.0052
0.105   384.0   384.5   0.6   0.0050
0.075   390.3   395.7   5.4   0.0029
pan   360.7   368.1   7.4   0.0000
Table AIII.5 Breakage distribution results for Culpeper red siltstone at RR1
  192
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.605   0.00   1.0000
38.100   0.00   1.0000
25.400   651.0   1346.5   695.5   0.7230
19.100   665.0   1183.5   518.5   0.5165
12.700   632.5   1075.0   442.5   0.3403
6.350   615.5   1062.0   446.5   0.1625
4.750   543.5   629.5   86.0   0.1282
3.360   506.5   596.5   90.0   0.0924
2.360   489.5   543.0   53.5   0.0711
1.700   475.0   518.0   43.0   0.0540
1.190   427.0   458.5   31.5   0.0414
0.850   433.5   459.0   25.5   0.0313
0.595   449.0   468.0   19.0   0.0237
0.425   434.5   448.5   14.0   0.0181
0.300   426.0   437.0   11.0   0.0137
0.210   450.0   459.0   9.0   0.0102
0.150   409.5   421.0   11.5   0.0056
0.105   384.5   386.5   2.0   0.0048
0.075   391.0   401.5   10.5   0.0006
pan   365.5   367.0   1.5   0.0000
Table AIII.6 Breakage distribution results for Culpeper red siltstone at RR2
  193
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.587   0.00   1.0000
38.100   0.00   1.0000
25.400   650.5   2135.6   1485.1   0.5213
19.100   664.0   1374.8   710.8   0.2922
12.700   631.8   984.7   352.9   0.1784
6.350   614.3   850.6   236.3   0.1023
4.750   542.4   588.9   46.5   0.0873
3.360   505.7   563.9   58.2   0.0685
2.360   488.7   523.4   34.7   0.0573
1.700   474.3   504.1   29.8   0.0477
1.190   426.3   448.1   21.8   0.0407
0.850   432.9   453.6   20.6   0.0341
0.595   448.5   463.6   15.1   0.0292
0.425   433.6   446.0   12.4   0.0252
0.300   425.2   434.9   9.7   0.0221
0.210   449.5   459.3   9.8   0.0189
0.150   408.9   419.4   10.6   0.0155
0.105   384.0   390.5   6.5   0.0134
0.075   390.3   401.1   10.8   0.0099
pan   365.2   395.9   30.7   0.0000
Table AIII.7 Breakage distribution results for Leesburg diabase at RR1
  194
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.608   0.00   1.0000
38.100   0.00   1.0000
25.400   651.0   1551.5   900.5   0.6763
19.100   665.0   1162.5   497.5   0.4974
12.700   632.5   1088.0   455.5   0.3336
6.350   615.5   1009.5   394.0   0.1920
4.750   543.5   639.5   96.0   0.1575
3.360   506.5   609.0   102.5   0.1206
2.360   489.5   550.5   61.0   0.0987
1.700   475.0   523.0   48.0   0.0814
1.190   427.0   465.5   38.5   0.0676
0.850   433.5   464.0   30.5   0.0566
0.595   449.0   473.0   24.0   0.0480
0.425   434.5   454.5   20.0   0.0408
0.300   426.0   441.0   15.0   0.0354
0.210   450.0   464.5   14.5   0.0302
0.150   409.5   429.0   19.5   0.0232
0.105   384.5   385.0   0.5   0.0230
0.075   391.0   416.5   25.5   0.0138
pan   361.5   400.0   38.5   0.0000
Table AIII.8 Breakage distribution results for Leesburg diabase at RR2
  195
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.477   0.00   1.0000
38.100   0.00   1.0000
25.400   651.0   1962.5   1311.5   0.4801
19.100   665.0   1148.0   483.0   0.2886
12.700   632.5   862.0   229.5   0.1976
6.350   615.5   854.0   238.5   0.1031
4.750   543.5   599.5   56.0   0.0809
3.360   506.5   562.0   55.5   0.0589
2.360   489.5   521.0   31.5   0.0464
1.700   475.0   501.0   26.0   0.0361
1.190   427.0   447.0   20.0   0.0281
0.850   433.5   449.0   15.5   0.0220
0.595   449.0   460.0   11.0   0.0176
0.425   434.5   443.0   8.5   0.0143
0.300   426.0   432.0   6.0   0.0119
0.210   450.0   455.5   5.5   0.0097
0.150   409.5   419.5   10.0   0.0057
0.105   384.5   388.5   4.0   0.0042
0.075   391.0   395.0   4.0   0.0026
pan   361.5   368.0   6.5   0.0000
Table AIII.9 Breakage distribution results for Shadwell metabasalt at RR1
  196
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.481   0.00   1.0000
38.100   0.00   1.0000
25.400   651.0   1202.0   551.0   0.7853
19.100   665.0   1144.5   479.5   0.5985
12.700   632.5   1189.5   557.0   0.3815
6.350   615.5   1116.5   501.0   0.1862
4.750   543.5   636.5   93.0   0.1500
3.360   506.5   602.0   95.5   0.1128
2.360   489.5   554.0   64.5   0.0877
1.700   475.0   523.0   48.0   0.0690
1.190   427.0   465.0   38.0   0.0542
0.850   433.5   462.5   29.0   0.0429
0.595   449.0   470.0   21.0   0.0347
0.425   434.5   451.5   17.0   0.0281
0.300   426.0   440.0   14.0   0.0226
0.210   450.0   468.5   18.5   0.0154
0.150   409.5   422.5   13.0   0.0103
0.105   384.5   389.0   4.5   0.0086
0.075   391.0   407.0   16.0   0.0023
pan   365.5   371.5   6.0   0.0000
Table AIII.10 Breakage distribution results for Shadwell metabasalt at RR2
  197
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.454   0.00   1.0000
38.100   0.00   1.0000
25.400   650.6   2179.0   1528.4   0.4035
19.100   663.9   1073.0   409.1   0.2438
12.700   631.6   859.0   227.4   0.1551
6.350   614.2   757.0   142.8   0.0993
4.750   542.2   579.0   36.8   0.0850
3.360   505.6   542.5   36.9   0.0706
2.360   488.9   511.0   22.1   0.0620
1.700   474.2   492.5   18.3   0.0548
1.190   426.0   440.9   14.9   0.0490
0.850   432.7   446.2   13.4   0.0438
0.595   448.4   460.9   12.5   0.0389
0.425   433.3   446.6   13.3   0.0337
0.300   425.3   439.8   14.6   0.0280
0.210   449.4   463.3   14.0   0.0226
0.150   408.8   425.2   16.4   0.0162
0.105   383.9   391.9   8.0   0.0131
0.075   390.2   401.9   11.6   0.0085
pan   365.1   386.9   21.8   0.0000
Table AIII.11 Breakage distribution results for Spotsylvania granite at RR1
  198
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.431   0.00   1.0000
38.100   0.00   1.0000
25.400   651.0   931.0   280.0   0.8686
19.100   665.0   1078.5   413.5   0.6745
12.700   632.5   1130.0   497.5   0.4410
6.350   615.5   1037.0   421.5   0.2431
4.750   543.5   626.5   83.0   0.2042
3.360   506.5   584.0   77.5   0.1678
2.360   489.5   535.5   46.0   0.1462
1.700   475.0   515.0   40.0   0.1274
1.190   427.0   459.0   32.0   0.1124
0.850   433.5   460.5   27.0   0.0997
0.595   449.0   473.5   24.5   0.0882
0.425   434.5   459.5   25.0   0.0765
0.300   426.0   450.5   24.5   0.0650
0.210   450.0   477.0   27.0   0.0523
0.150   409.5   442.5   33.0   0.0368
0.105   384.5   395.0   10.5   0.0319
0.075   391.0   417.5   26.5   0.0195
pan   365.5   407.0   41.5   0.0000
Table AIII.12 Breakage distribution results for Spotsylvania granite at RR2
  199
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.174   0.00   1.0000
38.100   0.00   1.0000
25.400   650.3   2173.4   1523.1   0.3908
19.100   663.9   948.5   284.6   0.2770
12.700   631.7   863.1   231.4   0.1845
6.350   614.3   811.1   196.8   0.1058
4.750   542.4   588.8   46.4   0.0872
3.360   505.8   547.8   42.0   0.0704
2.360   488.6   516.4   27.8   0.0593
1.700   474.3   496.5   22.2   0.0504
1.190   426.2   442.5   16.3   0.0439
0.850   433.0   446.6   13.6   0.0384
0.595   448.5   459.5   11.1   0.0340
0.425   433.6   443.5   9.9   0.0301
0.300   425.2   434.5   9.3   0.0264
0.210   449.5   459.4   9.9   0.0224
0.150   408.9   422.0   13.1   0.0171
0.105   384.0   391.5   7.5   0.0141
0.075   390.3   401.0   10.7   0.0099
pan   360.7   385.4   24.7   0.0000
Table AIII.13 Breakage distribution results for Thornburg granite at RR1
  200
 
Size Sieve Weight
Sieve + Sample 
Weight
Weight Retained
Fraction 
Passing
mm   g   g   g
47.192   0.00   1.0000
38.100   0.00   1.0000
25.400   651.0   1487.5   836.5   0.6509
19.100   665.0   979.0   314.0   0.5198
12.700   632.5   1087.5   455.0   0.3299
6.350   615.5   946.0   330.5   0.1920
4.750   543.5   625.0   81.5   0.1580
3.360   506.5   588.5   82.0   0.1237
2.360   489.5   536.0   46.5   0.1043
1.700   475.0   512.5   37.5   0.0887
1.190   427.0   457.0   30.0   0.0762
0.850   433.5   460.5   27.0   0.0649
0.595   449.0   471.0   22.0   0.0557
0.425   434.5   454.5   20.0   0.0474
0.300   426.0   444.0   18.0   0.0399
0.210   450.0   468.5   18.5   0.0321
0.150   409.5   443.0   33.5   0.0182
0.105   384.5   387.0   2.5   0.0171
0.075   391.0   407.0   16.0   0.0104
pan   361.5   386.5   25.0   0.0000
Table AIII.14 Breakage distribution results for Thornburg granite at RR2
  201
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
APPENDIX IV 
 
LABORATORY JAW CRUSHING RESULTS 
  202
 
Size
mm SMB BG CR SG
35.92   1.0000   1.0000   1.0000   1.0000
25.40   0.8947   0.8030   1.0000   0.6028
19.10   0.1776   0.2833   0.5093   0.0489
12.70   0.0170   0.0385   0.0157   0.0075
4.75   0.0147   0.0090   0.0076   0.0056
0.00   0.0000   0.0000   0.0000   0.0000
Fraction passing
Table AIV.1 Feed size data for lab jaw crusher
  203
 
Bulk 
density
Mass  Volume Time
Vol. 
capacity
Avg. Feed 
Size
K1
Predicted 
Capacity
kg/m
3
kg
  m
3
sec
  m
3
/hr
  m
  m
3
/hr
Culpeper siltstone   1702   38.80   0.02280   160   0.5130   0.0157   0.8389   0.8515
Shadwell metabasalt   1837   45.97   0.02502   195   0.4618   0.0185   0.8333   0.8457
Spotsylvania granite   1691   42.97   0.02541   169   0.5413   0.0182   0.8340   0.8464
Boscobel granite   1670   47.14   0.02823   172   0.5909   0.0212   0.8266   0.8389
Closed side set (m)   0.0079
Throw (m)   0.0064
Gape (m)   0.0953
Jaw width (m)   0.1461
Jaw length (m)   0.2096
RPM   300
K2   0.7079
Rock Type
Table AIV.2 Volumetric capacity results from lab jaw crusher tests
  204
 
Size Sieve Weight
Sieve + 
Sample 
Weight
Weight 
Retained
Fraction 
Passing
mm   g   g   g
19.10   579.5   579.5   0.0   1.0000
12.70   552.5   910.0   357.5   0.9292
6.35   599.5   3243.5   2644.0   0.4059
4.75   533.5   1083.5   550.0   0.2971
3.36   480.5   898.5   418.0   0.2143
2.36   489.5   754.0   264.5   0.1620
pan   366.0   1184.5   818.5   0.0000
Table AIV.3 Measured product size of Boscobel granite
  205
 
Size Sieve Weight
Sieve + 
Weight
Weight 
Retained
Fraction 
Passing
mm   g   g   g
19.10   1.0000
12.70   552.5   984.0   431.5   0.9211
6.35   599.5   3364.0   2764.5   0.4154
4.75   533.5   1061.0   527.5   0.3189
3.36   480.5   958.0   477.5   0.2316
2.36   489.5   779.5   290.0   0.1785
pan   366.0   1342.0   976.0   0.0000
Table AIV.4 Measured product size of Culpeper siltstone
Sample 
579.5   579.5   0.0
  206
 
Size Sieve Weight
Sieve + 
Sample 
Weight
Weight 
Retained
Fraction 
Passing
mm   g   g   g
19.10   579.5   640.0   60.5   0.9884
12.70   552.5   1443.0   890.5   0.8184
6.35   599.5   3235.5   2636.0   0.3150
4.75   533.5   954.0   420.5   0.2347
3.36   480.5   839.0   358.5   0.1662
2.36   489.5   694.0   204.5   0.1272
pan   366.0   1032.0   666.0   0.0000
Table AIV.5 Measured product size of Shadwell metabasalt
  207
 
Size Sieve Weight
Sieve + 
Sample 
Weight
Weight 
Retained
Fraction 
Passing
mm   g   g   g
19.10   579.5   600.5   21.0   0.9952
12.70   552.5   926.0   373.5   0.9103
6.35   599.5   3181.5   2582.0   0.3230
4.75   533.5   883.0   349.5   0.2435
3.36   480.5   746.5   266.0   0.1830
2.36   489.5   633.5   144.0   0.1502
pan   366.0   1026.5   660.5   0.0000
Table AIV.6 Measured product size of Spotsylvania granite
  208
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
APPENDIX V 
 
EQUILIBRIUM ANALYSIS OF WEDGE FORCES 
  209
  210
 
 
P
sp
 
P
P
v
Notch
P
(   )   (   )
(   )   (   )
(   )   (   )
(   )   (   )
Summing forces in the x and y directions:
: cos sin
2 2
:                         [EQ. 1]
cos sin
2 2
: sin cos
2 2
2
:                 [EQ. 2]
2sin 2 cos
2 2
Setting Equation 1 
x sp
sp
x
v
y
v
y
F P P P
P
F P
P
F P P
P
F P
   
=   
=
=   +
=
+
(   )   (   )
(   )   (   )
(   )
(   )
(   )
  (   )
(   )
equal to Equation 2:
cos sin
2 2
2sin 2 cos
2 2
1 tan
2
2 1
1 tan
2
2tan 1 cot
2 2
v
sp
v
sp
v
sp
P
P
P
P
P
P
   
   
=
+
   (
   
=
+
=   
+
 
VITA 
 
James  Donovan  was  born  on  June  24,  1975  in  Manalapan,  NJ.    He  graduated  from  St. 
John Vianney High School in 1993, where upon he enrolled at Virginia Tech.  He earned 
his Bachelor of Science degree in Mining Engineering in 1997.  During his undergraduate 
years he spent his summers working for New Hope Crushed Stone.  In 1999 he earned his 
Master  of  Science  degree  in  Mining  Engineering  from  Virginia  Tech  and  completed  a 
thesis entitled The Effects of Backfilling on Ground Control and Recovery in Thin-Seam 
Coal  Mining.    He  stayed  on  at  Virginia  Tech  as  a  doctoral  research  and  teaching 
assistant,  and  completed  his  PhD  requirements  in  the  summer  of  2003.    He  currently 
resides  in  Christiansburg,  Virginia  with  his  beautiful  wife,  Erin,  and  his  glorious  son, 
Quinn.     
  211