Sei sulla pagina 1di 21

Int. J. Miner. Process.

68 (2003) 71 – 91
www.elsevier.com/locate/ijminpro

Application of numerical modelling for prediction


of the influence of power density on
microwave-assisted breakage
D.N. Whittles, S.W. Kingman*, D.J. Reddish
School of Chemical Environmental and Mining Engineering, University of Nottingham,
University Park, Nottingham NG7 2RD, UK
Received 29 May 2001; received in revised form 3 May 2002; accepted 3 May 2002

Abstract

The influence of electric field strength on the microwave treatment of ore is elucidated. The ore
consisted of a microwave-absorbing mineral in a low-absorbing matrix, and the influence of electric
field strength was assessed by numerical simulation. Simulations were undertaken using finite
difference modelling techniques for a theoretical 15  30 mm sample of calcite host rock containing
10 vol.%, 1-mm2 particles of pyrite. The simulations modelled the microwave heating, thermal
conduction, expansion, thermally induced fracturing and strain softening and, finally, uniaxial
compressive strength to predict the effect of microwave heating on the strength of the ore material.
Standard correlations were then used to develop specific comminution energy verses t10 relationships
for the treated and nontreated samples. It is shown that microwave power density is vital to the
fracturing of the rock, and it is suggested that by utilising high power densities, the microwave
fracturing of rock to reduce grinding energy requirements may be economically viable.
D 2003 Elsevier Science B.V. All rights reserved.

Keywords: crushing; energy requirements; fracture mechanics; numerical modelling; microwave treatment

1. Introduction

The mechanical size reduction of solids is an energy-intensive and highly inefficient


process. It has been shown that in the US, 1.5% of all electricity generated is used in size

*
Corresponding author. Fax: +44-115-951-4115.
E-mail address: sam.kingman@nottingham.ac.uk (S.W. Kingman).

0301-7516/03/$ - see front matter D 2003 Elsevier Science B.V. All rights reserved.
PII: S 0 3 0 1 - 7 5 1 6 ( 0 2 ) 0 0 0 4 9 - 2
72 D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91

reduction and that based on the energy required to create new surfaces, grinding is less
than 1% efficient (Rhodes, 1998). It is clear that there is great incentive to improve the
efficiency of size reduction processes. Over several decades, this has promoted significant
amounts of research into size reduction processes. Unfortunately, this has only led to
small, incremental improvements in efficiency. One area that has shown significant
promise for improving the efficiency of comminution processes is thermally assisted
comminution. The application of this technology to aid the mechanical breakage of rocks
and mineral assemblages has been investigated by many workers (Veasey and Fitzgibbon,
1990). These studies have ranged from fundamental investigations of fracture mechanisms
to practical studies of complex mineral systems. Unfortunately, in each case, it was
concluded that the amount of energy required to reduce the mechanical strength of the
minerals was greater than that required for conventional comminution. Several other
benefits of thermally derived reductions in grinding energy were noted:

 Increased mill capacity and reduced wear,


 Better control of mill product size and improved liberation (recovery),
 A reduction in slimes production,
 Alteration of the physicochemical properties of ground products.

Interest in the application of microwave radiation to minerals started in the mid-1980s


when Chen et al. (1984) reported the results of heating 40 minerals individually with
microwave energy. The results showed that most minerals could be divided into two
groups: (1) little or no heat generated, mineral properties remained essentially unchanged;
and (2) heat generated, and minerals were either thermally stable or decomposed/reacted
rapidly into a different product. The test results indicated that most silicates, carbonates
and sulphates reported to Group 1, whilst most sulphides, metal oxides, sulphosalts and
arsenides reported to Group 2. This work was extended some years later by the US Bureau
of Mines who reported test results of microwave heating of a number of minerals and
reagent grade chemicals with similar results being produced to the earlier study
(Walkiwicz et al., 1988). It was suggested that rapid heating of ore minerals in a
microwave-transparent matrix generated thermal stress of sufficient magnitude to create
microcracks along grain boundaries. They suggested that this type of microcracking might
have the potential to improve ore grindability and increase liberation of individual mineral
phases.
Most minerals are brittle materials with a complex three-dimensional structure. The
precise nature of grain boundaries between two mineral phases in rock is not well
understood, but it is suggested to be an area of disorder between two ordered species. If
this were the case, then it would be sensible to assume that grain boundaries are an area of
weakness. However, products of comminution suggest that grain boundaries are areas of
strength (transgranular fracture being common in mineral processing operations) and can
adversely influence liberation of one species from another. If microwave energy can
induce microcracking around grain boundaries, then reductions in required comminution
energy and enhanced liberation of valuable mineral would occur.
The majority of testwork carried out so far concerning microwave treatment of minerals
have utilised standard multimode cavities, similar to that found in a conventional kitchen
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91 73

microwave oven. The multimode cavity, whilst mechanically simple, suffers from poor
efficiencies and low electric field strengths, vital to high power adsorption. Whilst the
influence of microwave radiation from this type of cavity has been shown to have a
significant influence on ores and minerals, the inefficiencies of the application method
have led to conclusions that at present, commercial microwave treatment of minerals
(despite the numerous process benefits) is not viable.
Single-mode cavities comprise of a metallic enclosure into which a microwave signal of
the correct electromagnetic field polarisation will undergo multiple reflections. The
superposition of the reflected and incident waves gives rise to a standing wave pattern
that is very well defined in space. The precise knowledge of electromagnetic field
configurations enables the dielectric material to be placed in the position of maximum
electric field strength, allowing maximum heating rates to be achieved at all times. In the
early evolution of microwave heating, such cavities saw little use. This was mainly
because they lacked the versatility offered by multimode cavities. However, the develop-
ment of electronic automatic tuning systems now means that they are finding favour in
industrial situations. They offer extremely rapid heating rates and the ability to heat
materials that appear transparent to microwaves in ordinary multimode cavities.

2. Numerical investigation of the affects of microwave heating

2.1. Numerical methodology

Numerical modelling was undertaken using the geomechanical 2-D finite difference
modelling software application, FLAC V3.3 (Itasca, 1995). The model domain consisted
of an area representing a 15-mm-wide by 30-mm-high section, which was subdivided into
individual square zones of 0.2-mm sides. The positions of the pyrite particles within the
model domain were randomly generated to provide a relatively disseminated ore body
(Fig. 1). This type of dissemination has previously been shown to be responsive to
microwave heating (Kingman, 1998). It is appreciated that the ‘mineralogy’ or texture
used for the modelling may be a simplified version of reality. However, the purpose of the
investigation is to determine the influence of power density on the degree of strength
reduction, not mineralogy. Therefore, as long as the mineralogy or texture is the same for
both tests, the data can be truly comparative. What is important, however, is that the
simulated ore contains species that are both responsive and nonresponsive to microwave
heating.
The finite difference modelling comprised of the five main stages given below and are
more fully described later:

(1) Microwave heating of the two different mineral phases


(2) Transient heat conduction during heating process between minerals
(3) Determination of peak thermally induced stresses and strains
(4) Modelling of thermal damage associated with material failure and strain softening
(5) Simulation of uniaxial compressive strength tests to evaluate the reduction of un-
confined compressive strength due to microwave heating.
74 D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91

Fig. 1. Model of the calcite and pyrite ore sample.

2.1.1. Stage 1. Microwave heating


The amount of thermal energy deposited into a material due to microwave heating
(power absorption density) is dependent on the internal electric field strength, the
frequency of the microwave radiation and the dielectric properties of the material.
The power absorption density per unit volume of the mineral can be approximated by
Eq. (1).

Pd ¼ 2pf eo erWEo2 ð1Þ

where Pd is the power density (W/m3), f is the frequency of the microwave radiation (Hz),
eo is the permittivity of free space (8.854  10  12 F/m), erW is the dielectric loss factor of
the mineral and Eo is the magnitude of the electric field portion of the microwave radiation
(V/m).
Because the microwave absorption factor for calcite is substantially lower than that for
pyrite, no microwave heating of the calcite matrix was assumed during the modelling with
selective heating of the pyrite particles only. The early work of Chen et al. (1984) and
Harrison (1997) shows this assumption to be realistic.
The dielectric loss factor, erW, for pyrite has been found to be dependent on temper-
ature (Salsman et al., 1996). In determining the power density for the pyrite, the
relationship between erW and temperature as shown in Fig. 2 was utilised (Salsman et al.,
1996).
For an initial series of models, the power densities at various temperatures was obtained
for the heating of pyrite within a 2.6-kW, 2.45-GHz multimode microwave cavity. The
calculated power density varied between 3  109 W/m3 at 300 K and 9  109 W/m3 for
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91 75

Fig. 2. Variation of dielecric loss factor of pyrite as function of temperature.

temperatures greater than 600 K (Fig. 3) (Kingman, 1998). The initial temperature of the
ore body sample was taken to be 300 K.

2.1.2. Stage 2. Modelling of transient heat conduction during microwave heating


The transient conduction of the microwave thermal energy during heating was
modelled using an explicit finite difference method written as an algorithm using FLAC’s
compiler language known as FISH.
The basic concept in the thermal conduction modelling was that a thermal energy flux
may occur between a zone and its four immediately adjacent zones. The direction, i.e. into
or out of the zone, and the magnitude of the thermal energy flux was dependent on the
temperature gradient that existed between the zones and the conductivity of the zone. The
boundary conditions were such that no thermal energy was lost from the material, i.e. the
material was assumed to be fully insulated.

Fig. 3. Variation of microwave power density of pyrite in a 2.6-kW, 2.45-GHz cavity as a function of temperature.
76 D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91

The basic law that was used to determine the thermal energy flow between the zones
was Fourier’s law (Duncan, 1985), which has been given as Eq. (2):
q ¼ kTdiff ð2Þ

where q is the heat flux vector (J/s/m), k is the thermal conductivity tensor (W/m C) and
Tdiff is the temperature difference (jC).
Thus, the change in stored energy per time increment, Dt, is given by Eq. (3)
Db ¼ Dtq ð3Þ

where Db is the change in stored energy (J).


Expressing this in an explicit finite difference form for a square zone i,j with side length l:
Dbði;jÞ ¼ Dtkði;jÞ l½ðTði;jÞ  Tði;j1Þ Þ þ ðTði;jÞ  Tði;jþ1Þ Þ þ ðTði;jÞ  Tðiþ1;jÞ Þ

þ ðTði;jÞ  Tði1;jÞ Þ ð4Þ

where k(i,j) is the thermal conductivity of zone i,j, Dt is the time increment in seconds, l is the
length of the sides of the zones and T(i,j) is the temperature of zone i,j.
The relationship between thermal energy in joules and temperature in degrees kelvin for
a given time increment, Dt, is given by Eq. (5):
Dbði;jÞ
DTði;jÞ ¼ ð5Þ
mði;jÞ Cði;jÞ

where DT(i,j) is the temperature change in zone i,j (K), m(i,j) is the mass of zone i,j (kg)
and C(i,j) is the specific heat of zone i,j (J/kg K).
Thus, at the end of each time increment, the new temperatures of each zone due to
thermal conduction and microwave heating are determined using Eq. (6)
Tði;jÞ ð1Þ ¼ 300K Tði;jÞ ðn þ 1Þ ¼ Tði;jÞ ðnÞ þ DTði;jÞ þ Pdði;jÞ =ðCði;jÞ DtÞ ð6Þ

where T(i,j)(n) is the temperature of zone i,j at time increment n and Pd(i,j) is the power
density of zone i,j.
The microwave heating and thermal conduction for a specified heating time, ht, was
simulated by recursively iterating Eqs. (4) –(6) until Eq. (7) was satisfied.
ht ¼ nDt ð7Þ

where n is the time increment number, Dt is the time increment in seconds and ht is the
heating time in seconds.

Table 1
Specific heat capacity as a function of temperature (Salsman et al., 1996)
Mineral Specific heat capacity (J/kg K)
298 K 500 K 1000 K
Calcite 819 1051 1238
Pyrite 517 600 684
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91 77

Table 2
Thermal conductivity as a function of temperature (Salsman et al., 1996)
Mineral Thermal conductivity (W/m K)
273 K 373 K 500 K
Calcite 4.02 3.01 2.55
Pyrite 37.90 20.50 17.00

The time increment, Dt, was restricted to 2.5  10  4 s to ensure numerical stability,
which itself corresponds to a measure of the characteristic time needed for the thermal
diffusion front to propagate through a zone.
The thermal conductivity and specific heat properties of calcite and pyrite vary with
temperature (Harrison, 1997) and have been included as references in Tables 1 and 2.

2.2. Thermal/mechanical coupling

2.2.1. Stage 3. Thermally generated strains and stresses


At the end of the heating interval, the thermally induced strains within a zone, assuming
perfect restrainment by the surrounding zones and isotropic expansion, is given by Eq. (8).

eði;jÞ ¼ aði;jÞ ðTnði;jÞ  T1ði;jÞ Þ ð8Þ

where e(i,j) is the strain in zone i,j, a(i,j) is the thermal expansion coefficient (1/K) of zone
i,j, Tn(i,j) is the final temperature of zone i,j and T1(i,j) is the initial temperature of zone i,j.
The thermal expansion coefficient for pyrite and calcite has also been found to be
temperature dependent (Harrison, 1997). Table 3 outlines the thermal expansion coef-
ficient at various temperatures for calcite and pyrite as assumed and implemented within
the modelling.
The calculated thermally induced stress within a zone can then be determined using
Hoek’s law for isotropic elastic behaviour (Jaeger and Cook, 1979) (Eq. (9)).

eði;jÞ Eði;jÞ
rði;jÞ ¼ ð9Þ
ð1  2tði;jÞ Þ

where r(i,j) is the isotropic thermally induced stress within zone i,j, assuming perfect
restrainment, E(i,j) is the Young’s modulus of zone i,j and t(i,j) is the Poisson’s ratio of zone
i,j.

Table 3
Thermal expansion coefficient as a function of temperature (Salsman et al., 1996)
Mineral Thermal expansion coefficient (1/K)
373 K 473 K 673 K 873 K
6 6 6
Calcite 13.1  10 15.8  10 20.1  10 24.0  10  6
Pyrite 27.3  10  6 29.3  10  6 33.9  10  6 –
78 D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91

2.3. Redistribution of thermally induced stresses

To obtain a state of static mechanical equilibrium throughout the domain of the


material, a redistribution of the thermally induced stresses and strains was necessary. To
obtain the equilibrium distribution, the model was stepped in FLAC’s default calculation
mode for static mechanical analysis. This default mode performs an explicit time-marching
finite difference calculation utilising Newton’s law of motion to relate nodal strain rates,
velocities and forces (Itasca, 1995). The material was assumed to behave as a linear
isotropic elastic medium with mechanical properties determined by the Young’s modulus,
Poisson’s ratio and density (Table 4).

2.3.1. Stage 4. Modelling of thermal damage associated with material failure and strain
softening
When static equilibrium was obtained, modelling of brittle fracture and subsequent post
failure strain softening, which is characteristic of the stress – strain relationship of a
crystalline limestone (Hoek and Brown, 1980), was undertaken by simulating the
constitutive behaviour of the ore material as elasto-plastic with plastic strain softening.
The strength of the ore was approximated as a very strong brittle crystalline limestone with
an unconfined compressive strength of 125 MPa and a shear strength related by a linear
Mohr – Coulomb strength criterion (Eq. (10)).
s ¼ rn tan/ þ c ð10Þ
where s is the shear strength, rn is the normal stress acting normal to the shear plane, / is
the friction angle of the material and c is the cohesive strength of the material.
Upon failure, the ore was assumed to behave as a brittle linear strain softening medium
undergoing plastic deformation with a final residual strength being obtained after 1% strain
again characteristic of a strong crystalline limestone (Hoek and Brown, 1980) (Table 4).

2.3.2. Stage 5. Simulations of the unconfined compressive strength tests on the thermally
damaged samples
The effect of thermal heating on the unconfined compressive strength and fracture
development within the modelled ore was predicted by the simulation of the uniaxial
compressive strength test on the thermally damaged models (Fig. 4).
The simulation was undertaken as a plane strain analysis with the material being
considered as continuous in the out of plane direction. The simulation was undertaken by
applying a constant velocity to the grid points positioned at the top and base of the model
domain whilst the left and right boundaries were unstrained. This is analogous to a

Table 4
Mechanical properties of the minerals (Salman et al., 1996)
Mineral Density Young’s Poisson’s Peak strength Residual strength (after 1% strain)
(kg/m3) modulus ratio / (j) c (MPa) T (MPa) /r (j) cr (MPa) Tr (MPa)
(GPa)
Pyrite 5018 292 0.16 54 25 15 54 0.1 0
Calcite 2680 797 0.32 54 25 15 54 0.1 0
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91 79

Fig. 4. Direction of simulated loading during the modelling of the uniaxial compression test.

displacement-controlled uniaxial compressive strength test. To monitor the load –defor-


mation relationship within the samples during testing, history files were generated from
the average stress conditions at the top and bottom boundaries. The models were run until
approximately 0.2% axial strain of the sample whereupon the models predicted failure
strength and some strain softening details of the samples was obtained.

2.4. Results of the numerical modelling

2.4.1. Microwave heating times


To determine the effect of microwave heating on the strength of the calcite and pyrite
ore, numerical modelling was undertaken for an unheated sample and for samples with
microwave heating times of 1, 5, 15 and 30 s. It was assumed that the samples were treated
in a multimode microwave cavity with a power density that varied from 3  109 W/m3 at
300 K to 9  109 W/m3 for temperatures greater than 600 K.

2.4.2. Temperature distributions


The modelled temperature distributions for each of the four time intervals is shown in
Fig. 5. It can be seen in the figure that the highest temperatures and temperature gradients
were generated where the pyrite particles were clustered. Table 5 summarises the
temperature distributions within the modelled samples for each temperature increment.
80
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91
Fig. 5. Modelled temperature distributions for a 2.45-GHz, 2.6-kW microwave cavity.
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91 81

Table 5
Modelled temperatures and unconfined compressive strengths for various microwave heating times (2.6-kW,
2.45-GHz microwave cavity)
Heating Maximum Minimum Unconfined compressive
time (s) temperature (K) temperature (K) strength (MPa)
0 300 300 126
1 350 300 126
5 460 320 123
15 700 400 97
30 900 600 79

Due to the length of time required to heat the pyrite particles within the 2.6-kW
microwave cavity, conduction of the deposited thermal energy from the pyrite into the
surrounding calcite host was predicted to occur. After 30 s of microwave heating time, the
calcite host had been heated to greater than 600 K. This conduction can be seen to reduce
the temperature gradient generated within the ore sample and thus reduce the thermally
generated stresses within the sample.

2.4.3. Effect of microwave heating on the unconfined compressive strength


The effect of the microwave treatment on the unconfined compressive strength of the
ore sample has been illustrated in Fig. 6 and summarised in Table 5. Fig. 7 shows the
unconfined compressive strength of the ore material plotted against microwave heating
time and indicates that the heating intervals of 1 and 5 s had little affect on the
unconfined compressive strength of the material. A more noticeable reduction in strength
was predicted with microwave heating times of 15 and 30 s. This observation may be
attributed to the fact that the rate of heating was insufficient to induce localised thermal
gradients of a magnitude that would generate thermal stresses that exceed the strength of
the ore material. Thus, the modelled reduction in strength of the ore body may be
attributed to the differential expansion of the pyrite and calcite material due to different
thermal expansion coefficients, generating stresses that exceed the strength of the
sample.

2.4.4. Pattern of shear planes


Also of interest was the pattern of the simulated shear planes developed within the
modelled samples after the unconfined compressive tests. These patterns have been shown
in Fig. 8 for the samples with microwave heating times of 1, 5, 15 and 30 s. The fracture
patterns developed within the microwave-heated samples were similar to the fracture
patterns displayed by the unheated sample, i.e. consisting mainly of continuous shear
planes inclined at approximately 25j to the direction of loading.

2.5. Effect of increasing the microwave power density

2.5.1. Power density and heating time intervals


To assess the effect of increasing the microwave power density on the temperature
distribution, unconfined compressive strength and shear plane development within the ore
82 D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91

Fig. 6. Effect of varying heating times on the numerically modelled stress – strain curves for the theoretical calcite
and pyrite sample (heated in a 2.6-kW, 2.45-GHz microwave cavity).

Fig. 7. Effect of microwave heating time on the predicted unconfined compressive strength of the theoretical
calcite and pyrite sample (2.6-kW, 2.45-GHz cavity).
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91
Fig. 8. Modelled shear plane development during unconfined compressive tests for a 2.45-GHz, 2.6-kW microwave cavity.

83
84
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91
Fig. 9. Modelled temperature distributions for a microwave cavity with a power density of 1  1011 W/m3.
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91 85

samples, a microwave power density of 1  1011 W/m3 was assumed for the pyrite
material. This power density was approximately 10 – 15 times greater than the power
density generated by using the 2.6-kW, 2.45-GHz microwave cavity, although still easily
within the range that can be achieved by microwave heating of pyrite in a single-mode
cavity (Salsman et al., 1996). It is assumed that this power density is achieved by a single-
mode cavity supplied with microwave energy at a power level of 15 kW at 2.45 GHz. The
calcite host material was considered to be unheated by the microwave energy. Due to the
higher power density, much shorter heating times of 0.05, 0.25, 0.5 and 1 s were
considered.

2.5.2. Temperature distributions


The modelled temperature distributions within the ore samples for each of the four time
intervals are shown in Fig. 9. The figure illustrates that significantly greater temperatures
were generated within the pyrite particles. The shorter heating times compared to the 2.6-
kW microwave cavity reduced the degree of thermal conduction, thus reducing the amount
of heating of the calcite matrix. This generated temperature gradients of a significantly
higher magnitude within the ore samples. The temperatures within the ore samples
obtained by the modelling have been summarised in Table 6.

2.5.3. Effect of microwave heating on the unconfined compressive strength


The effect of the microwave heating on the unconfined compressive strength of the ore
samples is illustrated in Fig. 10. Compared to the reduction in strength within the 2.6-kW
cavity, it can be seen in Fig. 11 that the higher power density generates a considerably
larger reduction in strength, with the majority of the strength reduction occurring very
quickly (within 0.05 s of microwave heating). The results of the modelling have been
summarised in Table 6.

2.5.4. Pattern of shear planes


The pattern of shear planes developed within the ore samples after the simulated
uniaxial compression test for the 0.05, 0.25, 0.5 and 1 s heating intervals are shown in Fig.
12. The figure indicates, unlike the unheated and 2.6-kW cavity heated samples, that the
shear planes are irregular and concentrated along the grain boundaries between the pyrite
and calcite. This may be attributed to the high thermally induced stress that develop along

Table 6
Modelled temperatures and unconfined compressive strengths for various microwave heating times (microwave
cavity with a power density of 1  1011 W/m3)
Heating Maximum Minimum Unconfined compressive
time (s) temperature (K) temperature (K) strength (MPa)
0 300 300 126
0.05 1200 300 57
0.25 1700 300 29
0.5 1900 300 26
1 1900 300 25
86 D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91

Fig. 10. Effect of varying heating times on the numerically modelled stress – strain curves for the theoretical
calcite and pyrite sample (heated microwave cavity with a power density of 1  1011 W/m3).

Fig. 11. Effect of microwave heating time on the unconfined compressive strength of the theoretical calcite and
pyrite sample (power density 1  1011 W/m3).
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91
Fig. 12. Modelled shear plane development during unconfined compressive tests for a microwave cavity with a power density of 1  1011 W/m3.

87
88 D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91

these boundaries due to the rapid localised heating and expansion of the pyrite particles
within the relatively unheated calcite matrix.

3. Discussion

The influence of microwave power density on a theoretical ore has been demonstrated.
The numerical simulation has shown very clearly that if the preferential dielectric material
can be made to absorb the majority of the applied energy, significant reductions in
compressive strength can be achieved. To further illustrate this in the context of
comminution, the extremely well-known relationships developed by Broch and Franklin
(1972) and Bieniawski (1975) were used to calculate the point load index (Is(50)) from the
modelled UCS data. The equation used was:

Is ð50Þ ¼ UCS=Kf ð11Þ

where Is(50) is the point load strength corrected to 50-mm core, Kf is an empirical constant
equal to 24 and UCS is the uniaxial compressive strength.
The results of this analysis are shown in Figs. 7 and 11. Fig. 7 shows the influence of
microwave heating time versus point load index for the lower power density. It can clearly
be seen that as microwave exposure time is increased, the point load index decreases
significantly. This is also true in Fig. 11, which shows microwave heating time versus
point load index for the ore exposed to the higher density. As for the UCS tests in Figs. 7
and 11, the reductions in point load index are particularly significant at the higher power
density with a reduction from 5.25 for nontreated to 1.25 after just 0.2 s.
Point load index is of particular interest to the mineral processing engineer because it
allows rapid prediction of the relationships between ECS (specific comminution energy,
kW h/t) and t10 (t10 is the percentage passing 1/10th of the initial mean particle size)
(Bearman et al., 1997). The parameter t10 can be interpreted as a fineness index with larger
values of t10 indicating a finer product. However, in practice, the value of t10 can be used
to reconstruct the size distribution of the broken ore. The t10 value is related to the specific
comminution energy by the following equation (Napier-Munn et al., 1996):

t10 ¼ A½1  eðbECSÞ ð12Þ

where A and b are material-specific breakage parameters. A is the theoretical limiting


factor of t10, and b is the slope of the ECS versus t10 plot. Determination of A and b for a
specific material can lead to the calculation of a specific size distribution for a specific
energy input.
It has previously been shown that point load index is intimately related to Mode 1
fracture toughness (Bearman, 1999). Bearman showed that

KIC ¼ 0:209Is ð50Þ ð13Þ

where KIC is the Mode 1 fracture toughness (MN/m3/2).


D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91 89

Table 7
Breakage parameters for the 2.6-kW multimode cavity microwave treatment
Time (s) Is(50) KIC b Ab A
0 5.25 1.097 1.91 107.61 56.03
10 4.45 0.93 2.54 145.16 57.14
30 3.4 0.7106 4.22 238.56 56.63

Mode 1 fracture toughness has also been shown to have highly significant correlation
with the breakage parameters A and b (Bearman et al., 1997).
It was shown that

1:6986
b ¼ 2:2465  KIC ð14Þ

1:8463
Ab ¼ 126:96  KIC ð15Þ

Table 7 shows the calculation of the breakage parameters for the theoretical ore exposed to
the 2.6-kW microwave radiation for 0, 10 and 30 s. Table 8 shows the calculation of
breakage parameters for the same ore treated at the higher power density. This data was
used in conjunction with Eq. (11) to calculate the influence of ECS on t10. Energy inputs of
0, 0.25, 1 and 2.5 kW h/t were used for the calculation. For clarity, data are only presented
for the nontreated and the most extreme treatment times, i.e. 30 s and 0.02 s. Fig. 13 shows
the influence of power density on the ECS versus t10 graph. It can be seen that as power
density is increased, the slope of the plot increases significantly and the theoretical limiting
value of t10 is reached for a much lower energy input. Put simply, this means that,
theoretically, ore treated at the lower power density produces a much coarser product for a
set specific comminution energy input than that treated at the higher power density. If it is
assumed that the mass of material heated is 1 kg, the sample energy input for each case is
for the 2.6-kW-treated sample heated for 30 s in the multimode cavity:

2:6  0:5=60  1000=1 ¼ 125 kW h=t

and for the 15-kW-treated sample heated in the single-mode cavity for 0.2 s:

15  3:33  103 =60  1000=1 ¼ 0:8325 kW h=t:

This clearly shows the influence of power density on the comminution of ores.

Table 8
Breakage parameters for the 15-kW, 2.45-GHz single-mode microwave cavity treated ore
Time Is(50) KIC b Ab A
0 5.25 1.097 1.91 107.01 56.03
0.1 1.8 0.376 11.83 772.67 65.31
0.2 1.25 0.2615 21.96 1513.41 68.91
90 D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91

Fig. 13. Plot of ECS versus t10 for nontreated and microwaved samples.

The purpose of this paper has been to illustrate the influence of power density (or
electric field strength) on the comminution of minerals. It is appreciated that the texture
used for the modelling stage is not exactly like a ‘real’ ore. However, the ore has
behaved in a similar manner to real ores previously tested (Kingman et al., 2000). In
addition, the values obtained for the breakage parameter A are similar to those expected
for a typical hard rock ore (Napier-Munn, 1996). It is the intention of the authors to
develop the technique so that the mineralogy is more typical of actual ores. When this is
done, quick estimations of the influence of variables such as grain size and location,
material properties and host rock can have on microwave-assisted comminution can be
undertaken. This technique then coupled with steady-state simulation packages, like
JKSimMet, will offer a powerful design tool for microwave-assisted comminution
circuits.

4. Future work

This numerical modelling programme has shown power density to have a significant
effect on stress formation in microwaved ore samples. The next stage of the work will be
to validate the predictions against ‘real’ ore samples. Ores of different texture will be
studied, so quantification of the underlying relationships between grain size, chemical and
physical properties and distribution can be developed.
D.N. Whittles et al. / Int. J. Miner. Process. 68 (2003) 71–91 91

However, the numerical modelling has been shown to be a useful comparative tool that
allows quick qualification of microwave heating variables.

5. Conclusions

Microwave power density has been shown to have a significant influence on the
thermo-mechanical failure of an ore. It has been shown that by increasing the power
density, significantly greater stresses are created for much lower energy inputs. This has
significant ramifications for the development of microwave-assisted comminution flow-
sheets. It is concluded that the use of high power density cavities may make the microwave
treatment of minerals economic especially when coupled with the additional benefits of
thermally assisted comminution.

References

Bearman, R.A., 1999. The use of the point load test for the rapid estimation of Mode I fracture toughness. Int. J.
Rock Mech. Min. Sci. 36 (2), 257 – 263.
Bearman, R.A., Briggs, C.A., Kojovic, T., 1997. The application of rock mechanics parameters to the prediction
of comminution behaviour. Miner. Eng. 10 (3), 255 – 264.
Bieniawski, Z.T., 1975. The point load test in engineering practice. Eng. Geol. 9, 1 – 11.
Broch, E., Franklin, J.A., 1972. The point load strength test. Int. J. Rock Mech. Min. Sci. 9, 669 – 697.
Chen, T.T., Dutrizac, J.E., Haque, K.E., Wyslouzil, W., Kashyap, S., 1984. The relative transparency of minerals
to microwave radiation. Can. Metall. Q. 23 (1), 349 – 351.
Duncan, T., 1985. Advanced Physics: Material and Mechanics, second edn. John Murray, London. ISBN 0-7195-
3854-8.
Harrison, P.C., 1997. A fundamental study of the heating effect of 2.45 GHz microwave radiation on minerals.
PhD thesis, University of Birmingham.
Hoek, E., Brown, E.T., 1980. Underground Investigations in Rock, Institution of Mining and Metallurgy, London
(text book).
Itasca, E.T., 1995. Fast Langrangrian Analysis of Continua, Version 3.3 Itasca Consulting Group, Minneapolis,
MN, USA.
Jaeger, J.C., Cook, N.G.W., 1979. Fundamentals of Rock Mechanics, third edn. Chapman & Hall, London.
Itasca, 1995. Fast Langrangrian Analysis of Continua, Version 3.3. Itasca Consulting Group, Minneapolis, MN,
USA.
Jaeger, J.C., Cook, N.G.W., 1979. Fundamentals of Rock Mechanics, third edn. Chapman & Hall, London.
Kingman, S.W., 1998. The effect of microwave radiation on the comminution and beneficiation of minerals and
ores. PhD thesis, University of Birmingham.
Kingman, S.W., Vorster, W., Rowson, N.A., 2000. The influence of mineralogy on microwave assisted grinding.
Miner. Eng. 13 (3), 313 – 327.
Napier-Munn, T.J., Morell, S., Morrison, R.D., Kojovic, T., 1996. Mineral comminution circuits. Their operation
and optimisation. JKMRC Monograph Series in Mining and Mineral Processing 2. JKMRC, Queensland,
Australia, pp. 81 – 83.
Rhodes, M., 1998. Introduction to Particle Technology. Wiley, Chichester, UK.
Salsman, J.B., Williamson, R.L., Tolley, W.K., Rice, D.A., 1996. Short pulse microwave treatment of dissemi-
nated sulphide ores. Miner. Eng. 9 (1), 43 – 54.
Veasey, T.J., Fitzgibbon, K.E., 1990. Thermally assisted liberation of minerals—a review. Miner. Eng. 3 (1/2),
181 – 185.
Walkiwicz, J.W., Kazonich, G., McGill, S.L., 1988. Microwave heating characteristics of selected minerals and
compounds. Miner. Metall. Process. 5 (1), 39 – 42.

Potrebbero piacerti anche