Sie sind auf Seite 1von 18

THE ROLE OF SOME BASIC ROCK PROPERTIES IN

ASSESSING CUTTABILITY
by

Frank F. Roxborough
The University of New South Wales, Sydney, Australia

1. INTRODUCTION

Assessing the cuttability of a rock formation in respect to the performance potential


of a mining or tunnelling machine is an imprecise area of endeavour.

Since the force required to break rock is governed by its strength, it is reasonable to
expect that the forces and energies involved in mechanical cutting should likewise be
determined by the strength of the rock. However, the magnitude of cutting forces
and energies occurring in machine mining is governed not only by the strength of the
rock being cut but also by the design and method of operation of the machine’s
cutting system.

Each of the elementary strength parameters of a rock (compressive, tensile and shear
strengths) is known to have an effect on the forces and energy required to cut it. But
their respective influences on cutting force (and thereby energy) vary, according to
both their absolute and relative values and depend also on the shape and disposition
of the cutting elements on the machine. For instance, cutting theory indicates that the
force generated by one particular pick shape is influenced by the tensile strength and
also by the ratio of tensile to compressive strength of the rock. For a different shape,
its performance will be affected either by tensile or shear strength depending on its
angle of presentation to the rock.

Such complex interaction between rock strength parameters and cutting system
geometry begins to reveal the difficulties, in practice, of relating rock properties to
cutting performance. It further underlines that “cuttability” can not be fully defined
in terms of rock properties alone.

Given ideal materials these difficulties could be dealt with quite readily, but rocks
are not ideal materials. They are notoriously heterogeneous, with a coefficient of
variation of typically 20 per cent and their properties in the mass are invariably
affected by structural defects such as joints, bedding and cleavage. The mineralogy of
a rock, most particularly its quartz content, is often a matter of crucial significance to
cutting. It has a major bearing on the rate at which a machine’s cutting elements wear
and thereby on the longer term average performance of the machine. Blunting of
2

cutters can be rapid in abrasive rocks and this leads to a corresponding fall in cutting
performance and loss of production. Such wide ranging factors apply to cutting
performance in a single rock formation. But the rock section falling within the path of
a mining machine commonly comprises interbedded strata or plies of very different
strength, character and thickness, the interactive effects on cutting performance of
which adds a further dimension to the difficulties involved in assessing cuttability.

2. CUTTABILITY ASSESSMENT — THE MAIN CRITERIA

The factors that affect the performance of mechanical rock cutting systems can be
grouped under the headings
• Cutter head design
• Cutter head operation
• Rock mass properties

Matters relating to cutter head design (cutter type and shape, spacing, head
diameter, torque etc.) and its operation (depth of cut, rotational speed etc.) can be set
aside in this paper. Although crucial to excavation performance, the fact is that the
type of machine to be used for excavating a given formation is open to selection.
Conversely, the rock formation in which the excavation is required is not a matter of
choice and, apart from the possibility of shifting the site, is beyond the control of the
engineer.

The question of cuttability, therefore, reduces to one of having to accept the rock
mass properties at the site but with the opportunity to choose the type of machine
and adapt it to the site conditions. In considering the relevance of rock mass
properties, some basis is required on which to link machine potential to the
measured properties of the rock. For a given machine type and configuration the
performance parameters that are, thereafter, governed substantially if not wholly by
the properties of the rock are
• Specific cutting force
• Specific normal force
• Specific energy
• Pick wear rate

These performance parameters refer to machines that deploy picks. There are
equivalent parameters for machines that deploy free rolling cutters. They are defined
and explained through Figure 1.
3

Figure 1 — Orthogonal forces for a cutter pick

2.1 Specific Cutting Force (FC/d)

This is the mean force (FC) per unit depth of cut acting on a pick in its direction of
cutting. The units are usually kN/mm. Cutting force for a pick, increases linearly
with depth of cut and so the torque and power required to cut rock at a specified rate
can be estimated from this parameter.

2.2 Specific Normal Force (FN/d)

This is the mean force (FN) per unit depth of cut acting on a pick normal to its
direction of cutting. Normal force also increases linearly with depth. It is the force
that must be provided by a machine’s thrusting system to achieve and maintain the
required depth of cut.

2.3 Specific Energy (SE)

This is the energy or work required to cut unit volume of the rock. The units are
usually MJ/m3. Multiplying cutting force (FC) by the distance the pick travels (L) gives
the energy expended in cutting. Dividing this by the volume of rock cut in the
process gives the specific energy. In a given rock it is used as a measure of the
efficiency of a cutting system with lower values indicating higher efficiency. But
more importantly, in the context of cuttability assessment, it can be used to both
compare the cuttability of different rocks and to indicate approximately the potential
excavation rate for a particular machine type in a given rock.

2.4 Pick wear rate

This is the rate at which a cutter pick wears in a given rock. It is measured as the
weight loss of pick material (usually tungsten carbide) per unit cutting distance
(units mg/m). Pick deterioration arises from two main sources, as abrasion due to
quartz or other abrasive constituents and as mechanical damage due to high rock
strength.
4

Of the four parameters defined above, specific energy and pick wear rate provide the
most useful basis on which to relate machine potential to rock properties.

3. DIRECT CUTTABILITY TESTING

A special cuttability test has been developed for the purpose of making direct
determinations of the four machine performance parameters for a given rock. Since
the test involves cutting the rock with a pick, it parallels (albeit in a simplistic way
and at smaller scale) the basic action of a mining machine. The procedure, which is
specifically designed for testing core samples, is standardised so the results exclude
machine design influences, permitting a direct comparison between the cuttability of
different rocks.

The test consists of cutting a groove 127 mm wide and 5 mm deep along the surface
length of a rock core sample parallel to its axis. The core is then rotated through 180°
to enable a similar parallel cut to be repeated. If the core survives this procedure,
similar cut replications are made at 90° and 270° rotations. With a typical core test
length of 250 mm, this provides a total cutting distance of up to 1.0 m. The test
arrangement is illustrated by Figure 2.

Figure 2 – Core cuttability test apparatus and arrangement.

Each set of up to four cuts is made using a new tungsten carbide cutter pick insert.
The insert, which is of simple chisel shape and 12.7 mm wide, has a front rake angle
5

of 0° and a back clearance angle of 5°. The tungsten carbide is of a standard quality
with a nominal grain size of 3-3.5 μm and cobalt content of 9-10%.

The principal forces acting on the pick are measured by means of a triaxial
dynamometer to which the pick is attached. The forces are thereby measured
continuously throughout the cut, and integrated on-line by computer. The amount of
rock cut from each groove is measured and used to calculate specific energy. The
tungsten carbide is weighed before and after the set of four cuts and its weight loss is
used to determine cutting wear.

Figure 3 — Abrasive wear testing arrangement.

Because, as mentioned in Section 2.4, all cutter wear might not be attributable to
abrasion, a supplementary abrasive wear test is sometimes undertaken in order to
identify the principal source of pick wear. This also uses a new tungsten carbide
insert which is made to take a continuous cut of nominally 25 m length. This is done,
as illustrated in Figure 3, by rotating a fresh piece of core in a lathe at nominally
50 r.p.m. and feeding the pick insert, which is disposed at an angle of about 45°,
axially along the outer surface of the core. The forward feed rate is controlled to
between 0.1-0.2 mm per revolution which ensures that virtually all wear is the result
of abrasion. This result, when compared to the cutting wear rate, gives an indication
of whether pick deterioration is likely to arise from high rock strength or rock
abrasion.

All four performance parameters, as discussed previously, are affected by cutting


system design and operational factors such as pick shape, depth of cut, cutting speed
and pick material, but the effects of these factors are reasonably well understood and
can, therefore, be isolated from the effects of rock properties. So, based on cutting
forces, specific energy and pick wear, all measured using a standard procedure, the
cuttability of a rock sample can be defined and, albeit arbitrarily, standardised.
6

4. INTERPRETATION OF CORE CUTTING TEST DATA

McFeat-Smith and Fowell (1977) have. related specific energy and cutter wear values
from the core cuttability test to the on-site performance of several roadheading
machines. The results are shown in Figure 4. It is significant to note that the shape of
the curves in Figure 4A conforms well with the trend indicated by theory.

Figure 4 – Relationship between core cuttability test results and (A) cutting
rates and (B) pick consumption for roadheaders.
(after McFeat-Smith and Fowell 1977)

It appears from these curves that specific energy and pick wear rates, derived from
the core cuttability test, can provide a reasonably good guide to the performance of
roadheading machines in that rock.

The test does, however, have a serious drawback. Because it involves a lengthy and
complicated testing procedure it is not suited to the rapid evaluation of a large
number of rock samples. Furthermore, it uses equipment and instrumentation that is
available in only a few geotechnical laboratories specializing in rock cutting
technology and research.

So although the test may provide reliable cuttability data, they are of value only if the
rock samples tested are truly representative of the rock mass to be excavated. Given
the heterogeneity of rock, it is necessary to test a large number of samples to
determine both reliable average cuttability values and their variability.
7

5. ROCK VARIABILITY

A fairly extensive scrutiny of the geomechanics literature has provided data on the
coefficients of variation1 (V) of the unconfined compressive strengths for 112 different
sedimentary rocks. The results which cover a range of sandstones, shales, limestones,
evaporites, coals and so on, are shown in Figure 5. Regrettably, many fewer data
were found for tensile and shear strengths and for the strength values for igneous
and metamorphic rocks. However, the data available showed the following average
coefficients of variability.
Variability in compressive strength
all, sedimentary rocks (112) = 21.7% (Figure 5)
sandstones (40) = 19.8%
limestones (16) = 22.0%
coals (18) = 27.2%
evaporites (8) = 14.2%
all igneous rocks (15) = 19.5%
all metamorphic rocks (10) = 15.3%
Variability in tensile strength
all sedimentary rocks (38) = 28.8%
sandstones (17) = 26.8%
Variability in shear strength
all sedimentary rocks (20) = 37.1%
sandstones (8) = 48.2%

These data indicate that rocks, in general, exhibit a coefficient of variability in


compressive strength of typically 20 per cent and a somewhat higher variability in
tensile and shear strengths. It is reasonable to expect from this that the variation in
cuttability within a given rock formation will be of similar order. Sufficient test work
must be undertaken, therefore, to provide not only reliable mean cuttability data but
also a measure of the variation about the mean that is likely to be encountered within
the rock formation.

6. ROCK STRENGTH AS A MEASURE OF CUTTABILITY

Whatever test method is used for assessing cuttability, it must be conducive to the
respective testing of quite large numbers of rock samples. The need is for a simple,
quick and reliable test.

1Coefficient of variation is defined as (S/X) x 100% where S is the standard deviation for a population
mean of X.
8

Several test methods, that qualify as simple and quick, are available. In addition to
the directly determined strength properties, there are a number of other tests
involving the use simple devices or procedures including:
Schmidt rebound number Bar wave velocity
Shore hardness Protodyakonov number
N.C.B. Cone indentor hardness Impact strength index
Point load index Penetrometer resistance

By and large these, along with some others, have been developed as indirect means
for assessing rock strength, their particular virtues being simplicity, portability or
non-destruction of the test sample. They cannot, therefore, be expected to provide
any better indication of cuttability than direct strength determinations. Of the simple
direct strength values for a rock, unconfined compressive strength is the easiest to
measure. It is less affected by test method than either tensile or she strengths.
Understandably, it is the most widely used measure of rock strength and,
furthermore, engineers have a ‘feel’ for it.

6.1 Rock property data base

An extensive data base has been established covering multiple core cuttability test
results, rock strengths and other properties for more than fifty different rock
formations. Each set of data refers to major formations occurring in different mining
and tunnelling projects, mainly in Australia and the United Kingdom.

The data base includes the following information

For all formations For at least 70%of the formations For at least 40%of the formations

Specific cutting force Abrasive wear Shore hardness


Specific normal force Tensile strength N.C.B. Cone indentor
Cutting wear Shear. strength- Schmidt rebound number
Compressive wear Young’s modulus Impact strength index
Compressive strength Quartz content
Bulk density
(Rock moisture content)

The data base has been used to relate the various rock properties to the standardized
machine performance parameters as defined by the core cuttability test. Linearity
between the rock strengths and both pick force and specific energy has been assumed
and tested for, on the grounds that is both expected from theory and can be
demonstrated by controlled experiment. It is unrealistic to present graphically all of
the relationships that are available from the data base. Suffice to say that unconfined
compressive strength provides a marginally better correlation with specific cutting
9

force and specific energy than either tensile or shear strength and a significantly
better correlation than modulus or any of the listed indirect strength test methods.

6.2 Compressive strength v cuttability parameters

The relation between compressive strength and specific energy for all rocks in the
data base is shown in Figure 6. Those for compressive strength versus cutting force
and versus normal force are much the same as that for specific energy, in terms of
both linear correlation coefficient (r) and scatter.

Figure 6 shows wide prediction limits allowing only a tentative prediction of specific
energy from compressive strength and involving a large degree of variability about
the predicted mean. For example, a rock strength of 150MPa would suggest a specific
energy of 23 MJ/m3 but with a 1 in 40 chance that it would be greater than 38 MJ/m3
and a 1 in 40 chance less than 8 MJ/m3.

Figure 5 – Coefficient of variation for some sedimentary rocks.


10

Figure 6 – Relationship between unconfined compressive strength and


specific energy (all rock types).

Bearing in mind that machines deploying picks, such as roadheaders, are mostly
used in sedimentary rocks where the compressive strength is less than about
100 MPa, corresponding data has been abstracted from the data base to provide
Figure 7.

Figure 7 – Relationship between compressive strength and specific energy


for sedimentary rocks.

This shows a much better correlation than Figure 6 and with substantially less
variability. In this case, for example, a compressive strength of 40 MPa would
indicate a specific energy of about 10 MJ/m3 with a 1 in 40 chance of it being more
than 13 or less than 7 MJ/m3. Similarly, as indicated by the 50 per cent prediction
limit, there is a 75 per cent chance that it would not exceed 11 MJ/m3.
11

Two of the sixty formations making up the data base provide sufficient information
on which to examine the relation between compressive strength and specific energy
that occurs within a single formation. These are shown in Figure 8.

Figure 8 – Relationship between compressive strength and specific energy


within two formations.

Figure 8A is for the Bunter Sandstone (U.K.) and Figure 8B for the Hawkesbury
Sandstone (Australia) each of which is, or has been, host to several tunnelling and
mining projects for which a substantial volume of rock cuttability and strength data
have been accumulated. It is interesting now to compare Figures 7, 8A and 8B which
provide the following equations.
for all sedimentary rocks SE = 0.25σ + 0.11 ..............................................1)
for Bunter sandstone SE = 0.24σ + 0.02 ..............................................2)
for Hawkesbury sandstone SE = 0.23σ + 2.65 ..............................................3)

where SE = specific energy (MJ/m3)


σ = U.C.S. (MPa)

These equations are substantial1y the same. The only significant discrepancy is the
2.65 constant for Hawkesbury Sandstone, which in each case should be virtually
zero.

With Figures 7, 8A and 8B showing essentially the same trend, it is reasonable to


conclude that compressive strength can provide a good indication of specific energy.
In this context the following data are relevant.
12

r V

All sedimentary rocks 0.81 21.7%


Bunter Sandstone 0.94 4.7%
Hawkesbury Sandstone 0.76 35.5%

(for which r is the correlation coefficient for the equation and V is the measured
coefficient of variability of the corresponding rock).

It can be inferred from this that the scatter of the data points in Figures 7, 8A and 8B
is attributable to rock strength variance and given a “perfect” rock (i.e. V = 0%) the
relation between strength and specific energy would likewise be perfect (i.e. r = 1.0).

This now brings the issue of rock variability into sharper focus. The strength of rock
at a particular stratigraphic and geographic location in a formation is variable, as
illustrated previously in Figure 5. Several test replications are needed, therefore, in
order to establish not only a reliable mean but also a measure of strength variability
at that location.

The suggested minimum number of strength test replications in order to provide an


acceptable mean is shown in Figure 9 (IBRM 1964).

Figure 9 – Test replication requirements.

Rock strength and related properties, however, can also change very markedly
throughout the formation. The sort of geographic variation that can occur is implicit
in Figures 8A and 8B which show, for example, that the mean strength for Bunter
Sandstone at 29 different locations ranged from 10 to 60 MPa. So assessing machine
potential for a mining or tunnelling project must necessarily involve a large amount
of test work, the extent of which will be governed by the geographic and
stratigraphic span of the project.
13

However, the relationship that has been found to exist between unconfined
compressive strength and specific energy provides the opportunity to base cuttability
assessment for sedimentary rocks largely on simple strength determinations. This
would involve establishing reliable cuttability data using the core cutting or similar
test at one location and thereafter assessing variability throughout the rock mass on
the basis of compressive strength, using the equation

S.E. = 0.25 σ + C................................................4)

An assessment based exclusively on compressive strength would be much less


reliable because:
i. it requires the assumption C = 0 (or some preferred value) in equation (4), and
ii. there is no relationship between compressive strength and cutter pick wear
rate (see Figure 10).

6.3 Pick Wear

The data base shows, ostensibly, that cutter pick wear is unaffected by the
compressive strength of a rock. However, by separating pick wear into its main
components, it is found that whereas abrasive wear is unaffected (Figure 10A) the
amount of chipping of the carbide increases with rock strength (Figure l0B).

There is a wide scatter of data in Figure l0B but the underlying trend of an increase
with strength is statistically significant. Such excessive scatter is attributed in large
part to random variations in the size of the carbide fragments breaking off the pick,
exacerbated by the cubed relation between fragment mass and its linear size.

Figure 10 – Effect of compressive strength on pick wear.


14

The dominant factor in abrasive wear is the quartz content of the rock as shown in
Figure 11A. This refers to a medium grade tungsten carbide in sedimentary rocks.

Figure 11 –Effect of quartz content on pick wear rate.

Similar abrasive wear data produced by Schimazek and Knatz (1976) is shown in
Figure 11B. In this case, however, the picks were made of steel and were pointed in
shape.

Figure 12 – Pick wear as a function of the square of quartz content.

The underlying trend for both graphs appears to be the same. Grouping the data and
taking statistical mean values produces Figure 12, which suggests that,
15

i. abrasive wear rate, defined in terms of pick weight loss, increases as the
square of the quartz content of the rock.

ii. the rate of increase in wear is substantially greater with steel than with
tungsten carbide, reflecting the difference in hardness of the two materials.

Figure 13 – Comparison of equation (4) and transposed (AM75) roadheader


prediction data.

Making allowance for the difference in density between steel (7.8 g/cc) and tungsten
carbide (14.0 g/cc), it is found that the wear rate for steel is about four times greater
than for tungsten carbide. This corresponds roughly with the relative hardnesses of
the two materials.

Although wear rate is dominated by quartz content it is also affected to a significant


extent by
• quartz grain size
• quartz grain angularity, and
• degree of cementing

Regrettably these effects have not all been well defined. Schimazek (1976) and
Larson-Basse (1974) have both shown, however, that cutter wear increases linearly
with unidimensional grain size.

Concerning angularity and degree of cementing, it can be postulated that more


angular quartz and also a stronger matrix will produce greater wear.

Pick wear has a big effect on machine performance. The South African Chamber of
Mines Research Organization (Baker et al 1981) has classified pick wear into six
progressive categories from unworn to worn out. Studies on the performance of
worn picks show that wear can occur very rapidly in abrasive rocks, and for most of
their operating life, in such rocks, the picks are in a typically class III condition.
16

Compared to the class I condition, this involves


i. a 2-2.5 fold increase in cutting force
ii. a 2-2.5 fold increase in specific energy
iii. a 4-6 fold increase in normal force.

So the reduction in machine performance due to pick wear can be very substantial
and this must be taken fully into account when interpreting cuttability data.

It is worth noting that the most deleterious effect of is to cause a disproportionately


higher increase in pick normal force. This puts an excessive demand on the thrust
capabilities of a machine such that it may not be able to penetrate the face sufficiently
for the picks to interact. This is often reflected in practice where machines are found
to be thrust rather than torque limited.

7. OTHER FACTORS IN CUTTABILITY ASSESSMENT

Some other features of a rock mass that have the potential to influence cutting
machine performance include jointing and other structural defects, the state of stress
in the rock mass, and its moisture content. Some brief comment is offered on each.

7.1 Discontinuities

The presence of discontinuities in the rock mass generally has a favourable effect on
cutting. A higher frequency, length and degree of openness of joints and a greater
prominence of other structural features, such as bedding, cleats and slips, can assist
the cutting process. However the effect becomes significant only when the joints are
open and when the spacing between them is less than about five times the picks
intended depth of cut. An adverse effect of jointing should be noted. Its geometry
may be conducive to large blocks of rock being dislodged from the face by the
machine, with the prospect of downstream handling problems.

7.2 Moisture content

The mechanical strength of a rock commonly reduces as its moisture content


increases. It does not follow, however, that when saturated a rock is easier to cut than
when dry. In fact where a saturated rock remains competent, the forces required to
cut it can be somewhat higher than for dry rock, even though its mechanical strength
is significantly reduced. The phenomenon has been observed. in rocks of quite
different character and, although not universally found nor fully understood, there is
a plausible explanation for it. This lies in the dissipation of high local stress
concentrations at the pick-rock interface by pore water. One benefit of rock moisture
is that it usually results in a lower pick wear rate.
17

7.3 In-situ stress conditions

Where a formation is joint free the state of stress in the rock mass, originating
principally from gravity forces, usually has little effect on cutting performance.
Conversely, where structural defects exist, in-situ stress often has the effect of
increasing cutting resistance up to a critical stress level beyond which cutting
becomes easier. Near surface formations are unlikely to be affected to any discernible
extent, since the gravity forces are generally low.

8. CONCLUDING REMARKS

Rock cuttability is neither an absolute nor a precisely definable quantity. This is


because machine designs, and what is expected by way of production from them, are
just as variable as the properties of the rock in which they are required to work..

Setting aside for the moment the question of machine design, it is possible to provide
rock property data on which the relative cuttability between and within different
rock formations can be based. Thereafter any assessment of machine potential,
benefits from a knowledge of the historical performance of the same or similar
machines in other known situations. In the final analysis the decision making process
on machine applications must be leavened with sound engineering judgement.

The link between laboratory test data and machine performance is rather tenuous.
There have been too few instances where the loop has been closed; where rock
assessment data and performance predictions have been matched against the
measured performance of a machine.

In this paper, examination of the role of rock strength in cuttability assessment has
been based wholly on laboratory determinations. The only direct link with machine
mining is through Figures 4A and 4B produced by McFeat-Smith and Fowell (1977).

An indirect link, or at least grounds for supporting, the proposition, is provided by


Voest-Alpine who, in their equipment brochure, publish relationships between rock
compressive strength and the average cutting rates for each of their main types of
roadheaders. The data for one of these machines, the AM75, has been transposed to
produce curve A in Figure 13. This assumes that 100kW of the available 160kW cutter
motor power reaches the pick point.

The relationship S.E. = 0.25 σ, as established earlier in the paper, superimposed as


curve B. The following and final comments are appropriate.

i. There is close agreement between curves A and B for compressive strengths


below 30-40 MPa, but with significant divergence at higher strengths.
18

ii. The difference in shape of the curves is not incongruous and, in fact, conforms
with cutting theory. Specific energy for curve B is for a constant depth of cut
and is therefore linearly related to strength. The depth of cut associated with
curve A will be reducing progressively as rock strength increases involving a
corresponding increase in specific energy.

iii. The difference between curves A and B gives an indication of the effect
machine design and operation factors have on specific energy and hence
cuttability assessment.

iv. The agreement between curves A and B for rock strengths up to 30-40 MPa
implies that the AM75 machine can operate at or close to its maximum depth
of cut over this strength range.

v. Curves for other types of machine would produce different valued but
similarly shaped curves to A. Likewise, if cuttability testing was based on a
different depth of cut, the relationship between specific energy and rock
strength would still be linear but with a different gradient. The numerical
agreement between curves A and B at the lower strengths is, therefore,
coincidental between the AM75 and the cuttability test specification.

9. REFERENCES

Baker, R.H., Jones, G.J. and Hardman, D.R. (1981). An Assessment of Pick Wear and
effect on Continuous Miner Performance. Sth African Chamber of Mines Res. Org.
Report No. 10/81.

Larson-Basse, J., Shishido, C.M., and Tanouye, P.O. (1974). Some Features of Abraded
WC—Co Surfaces. J. of Austr. Inst. of Metals, Vol. 19, No. 4, December.

McFeat-Smith, I. and Fowell, R.J. (1977). Correlation of Rock Properties and the
Cutting Performance of Tunnelling Machines. Proc. of Conf. on Rock Engineering,
University of Newcastle upon Tyne, England, 4-7 April.

Schimazek, J. and Knatz, H. (1976). Die Beurteilung der Bearbeitbarkeit von


Gesteinén durch Schneidund Rollenbohr-werkzeuge. Erzmetall 29.3

Original version of this paper was presented at Seminar on Tunnels – Wholly


Engineered Structures, 1987, April, AFCC (IEAust: Canberra), pp1-21.

Das könnte Ihnen auch gefallen