You are on page 1of 13

Water Jet Cutting Related to Jet & Rock Properties

David A. Summers

Rock Mechanics & Explosives Research Center


University of Missouri at Rolla

Abstract.

A continuous water jet is used to cut slots in seven rock types in an experiment, which
correlates jet and rock properties with jet performance. The jet, operating at pressures from 5,000
to 25,000 psi is traversed across the rock surface at speeds between 15 and 750 ft/min. The rock
properties considered in the analysis are uniaxial compressive strength, Young's modulus, Shore
hardness, Schmidt hammer value, Rock Impact Hardness Number and Rock Fracture Toughness.
Regression analyses are performed on the results, based on the 7,180 measurement tests of the
depth of slot cut.
Introduction

Five years ago Dr. Maurer (1)1 analyzed twenty-three novel drilling techniques to identify
those, which might improve the performance of, rock cutting equipment. At that time there were
very few people who considered water jets as ever being anything but a laboratory project, with
little or no field application. This April over 200 delegates attended the first International
Symposium on Jet Cutting Technology (2) and discussed the growth and use of this new cutting
tool.

Water jets are currently in commercial operation, cleaning civil and industrial installations
ranging from sewers (3) to heat exchangers and chemical plant components filled with solidified
plastics (4). Water jets are also used to debur aluminum engine parts (5) and to cut furniture forms
(6). The jets operate at pressures up to 40,000 psi in continuous daily operation (6) and are
considered a success, the overall savings to industry adding up to several hundreds of thousands
of dollars.

And yet, in one of the areas where water jet research has been continued for the longest time,
the applications have not yet advanced beyond the first tentative field studies. Water was used to
cut rock in the time of the Romans (7), and laboratory research has been carried out at jet pressures
above 10,000 psi since 1962 (8). Rocks ranging in strength from shale to quartzite have been cut
at pressures of less than 30,000 psi, and cutting speeds of up to 1,000 ft/min have been achieved.
Yet there is no widespread acceptance of the method. Perhaps this is because there has been a lack
of published data on the performance of water jets and their relative effectiveness. The impression
may also have been given that it is necessary to use very high pressures in order to cut rock
effectively. This paper describes the results of an experiment in which 7,180 measurements were
made of continuous jet penetration into 7 rocks. Tests were made with a commercially available
pump and very simple test equipment. The object was to find what properties basically effected
how deep and how well a water jet can cut rock and to demonstrate that it is not necessary to use
ultra-high jet pressures in order to do this.

Background.

Two ways of breaking rock with water jets are currently being considered (9). One is to use
the jets to slot the rock surface and to break off the ribs remaining between the slots by a
mechanical means. The second is to use higher pressure pulses of water to shatter the rock on a
large scale. Both ideas have their merits although, while pumps are commercially available to
produce the jets for the first method, equipment to produce the higher pressure pulses is still
experimental, and some doubts have been raised as to its reliability in long term field operation
(10), due to fatigue of the high pressure cylinder and the short life of the high pressure seals.
Nevertheless most of the government funded research has been in the area of the high pressure
pulsed systems. Data from research with pulsed jets has been analyzed and correlations have been
made between jet performance and water jet properties (11) and jet performance and rock
properties (12).

Cooley (11) defined a specific fracture energy of rock as the square of the rock compressive
strength divided by twice the modulus of elasticity. Specific energy of breakage (13) was divided
by this term to give a dimensionless property which defined jet performance in the rock and which
was correlated with several jet properties. The properties considered were defined in a
dimensionless form, and the terms considered were the jet velocity ratio, a jet pressure parameter
and the standoff ratio. The jet velocity ratio is the ratio of the jet velocity to the traverse speed of
the jet nozzle relative to the rock surface. The pressure parameter is defined as the jet pressure
divided by the specific fracture energy of the rock and the standoff ratio is the ratio of the standoff
distance to the nozzle diameter. Cooley found that for the most effective cutting, jet pressure
should exceed the compressive strength of the rock and that the jet velocity ratio should be less
than 1,000.

1
Numbers in parentheses refer to references listed at the end of the paper.
Singh and Huck (12) correlated the penetration of a high pressure pulse of water with nine
rock properties measured for 6 rock types. They found that a regression equation could be written
for crater depth based on two rock properties, compressive strength and Schmidt hammer value,
and also based on the water jet pressure. Young's modulus, wave velocity, Poisson's ratio, tensile
strength and Shore scelerescope values were also considered in the study. Results from the
experiment also indicated that an increase in cutting effectiveness is obtained with an increase in
nozzle diameter and that the effectiveness of the jet cut increases with jet pressure to a point where
jet pressure equals 35 times the rock compressive strength.

Experimental Design

Rock is not a homogeneous, isotropic material, and for this reason the experiment was
designed to yield a large amount of data so that variation in rock properties and jet penetration due
to this variation should be kept as small as possible.

Evidence from preliminary studies had shown that it is more effective to cut slots in the rock
surface and remove the rock between the slots mechanically than it is to remove the rock by water
jets alone (151. A reduction in specific energy of breakage from 300 joules/cm3 to 20 joules/cm3
was obtained in one test with Berea sandstone. Further, it had been found that a continuous
traversing jet penetrated further into the rock in the same amount of time than either a pulsed jet or
a stationary, continuous jet (Fig. 1) at the same pressure (14).

Figure 1. Penetration of a jet into red Woolton sandstone at a pressure of 8,000 psi. and a standoff
distance of 2 inches. (after Summers [14]).

The experiment was therefore designed to find the parameters controlling the slot depth cut by the
jet and the relative efficiency of cutting. The initial design was to vary the jet pressure, from
5,000to 25,000 psi. the traverse speed of the jet nozzle over the rock surface, from 15 to 750
ft/min, and the number of passes made by the nozzle over the same area of the rock surface from 1
to 16. Data from this experiment, carried out in Berea sandstone and Indiana limestone, indicated
(Fig. 2) that the jet became increasingly inefficient with increasing pass number (15, 16). At the
same time it is unlikely that, in any machine design, the jet will pass over the surface of the rock
more than once between passes of the mechanical cutters. The design of the experiment was
accordingly changed, and a fourth parameter, nozzle diameter, incorporated into the experiment. A
single pass was made over the rock surface for the remaining tests.

Figure 2. Specific energy of breakage versus number of passes. (After Summers [15]).

Experimental Procedure.
Cubic samples of rock, measuring six inches along each side, were prepared from seven
rocks: four sandstones, one limestone, one marble and one granite (Table 1). Each sample to be
tested was placed in the chuck of a lathe and clamped with the axis of rotation passing through the
center of the specimen. The nozzle and the feed pipe from the supply-pump were attached to the
horizontal feed carriage of the lathe (Fig. 3). The gearing of the lathe allowed the nozzle to be
traversed across the rock surface at a constant feed rate per revolution of the chuck. A slotted steel
plate was placed between the rock sample and the nozzle. The slots in the plate were cut to the
same width as the distance moved by the nozzle in a single revolution of the chuck. The jet was
therefore exposed to the rock surface for one revolution of the chuck as the nozzle moved over the
slot. Five slots were made at half inch intervals and spaced from the center of the axis of rotation.
It had previously been determined that there was no interference between adjacent jet slots when
separated by this distance (14).

Figure 3. Experimental Equipment


Table 1. Properties of Rocks Tested in the Experiment

CS* YM SHOR SCHH RIHN FT


Properties (psi) (106 psi)
Berea Sandstone(l) 5,480 3.37 26.8 35.1 3.85 147
Indiana Limestone(2) 7,990 4.15 20.8 37.5 14.3 428
Barre Granite(3) 24,550 4.76 47.5 77.6 32.6 672
Georgia Marble(4) 23,100 8.30 32.9 38.6 12.9 327
Buff Sandstone(5)** 5,220 2.26 20.5 30.0 1.87 101
Pink Sandstone(6)** 6,750 2.55 22.6 29.5 1.40 131
Red Sandstone(7)*** 8,640 2.79 25.9 25.2 16.75 828
*Properties are symbolized by the abbreviations given in Table 4.
**Supplied by Briar Hill Stone Company, Briar Hill, Ohio
***Supplied by Jacob's Creek Stone Company, Denton, North Carolina

The distance between the nozzle and the rock surface was standardized at two inches.
Previous testing has shown that distance has little effect on jet performance in the range from two
to four inches (17). A soluble oil was added to the water in order to lubricate the pump. Despite
their known advantage (18, 19), no long chain polymers were added to the water. This was
because the equipment included a recirculation system, and the polymer is known to age (20) and
could be expected to denigrate after passing through the pump a number of times. This would have
included an uncontrolled variable into the experiment.

The result of each cutting run was to leave a concentric set of slots cut in the rock surface
(Fig. 4). The traverse speed of the jet nozzle relative to the rock surface for each slot was a
function of the slot radius and the rotational speed of the chuck. The depth of the slot was
measured at our points around the periphery of the slot, and an average value obtained. This value
was entered in the raw data file. The data was processed using the SPSS computer program (21) to
determine the statistical coefficients between the dependent and independent variables and to derive
regression equations.

Figure 4. Specimen of Buff Sandstone after cutting.

Three dependent variables were considered in the correlation, the depth of the slot cut, the specific
energy of rock breakage, and the specific energy ratio. The specific energy of breakage is the
energy input to the rock per unit volume of rock removed, and the specific energy ratio is the
specific energy of breakage divided by the specific fracture energy of the rock. The energy input to
the rock was assumed to be the jet kinetic energy, where the velocity was calculated from the
experimentally determined equation (22)
V = 10.69 P
Where V is jet velocity in ft/sec and P is jet pressure in psi.

To determine the volume of rock removed it was assumed that the jet cut a slot with a width
3.5 times the nozzle diameter. This value was found to give a better agreement with the
experimental data than the 2.5 value predicted by measurement of the pressure distribution under a
jet (23). The reason for this is probably the high lateral velocity of the jet after impact, which
caused cutting into the side of the slot. In the initial stages of impact, this lateral velocity has been
calculated to be up to thirty times the impacting jet velocity (24).

Water Jet Properties.


The water jet properties considered in the analysis were the jet pressure, the jet velocity, the
nozzle traverse speed and the diameter of the jet. The energy contained in the jet impacting on one
square centimeter of the rock surface and the jet velocity ratio defined by Cooley (11) were also
calculated. The test pump available supplied a constant volume of water to the high pressure
system. Pressure in this system was controlled by a bleed-off circuit through which fluid could be
returned to the pump reservoir. Because of the constant maximum output from the pump, as the
nozzle diameter increased, the maximum jet pressure available was reduced. A comparison was
made between the jet performance at maximum pump output for the three nozzle diameters tested.
Pearson product-moment correlation coefficients were calculated between jet performance and
individual jet properties (Table 2). Multiple regression coefficients were also calculated by
determining the correlation between jet performance and a combination of jet properties (Table 3).
A step-wise regression was performed and only properties which increased the square of the
multiple regression coefficient by more than 0.1 were considered in the regression. By this means
multiple regression coefficients of greater than 0.9 could be realized.

Table 2. Pearson Cross-Moment Correlations Between Jet Performance & Properties

a) Correlation with Slot Depth


Rock (1) (2) (3) (4) (5) (6) (7)
Jet Pressure .557 .574 .273 .403 .483 .537 .601
Nozzle Diameter .027 -.251 -0.005 -0.066 0.030 .095 -
No. of Passes .258 .591 - - - - -
Traverse Speed -.465 -.423 .583 -.464 .432 -.467 -.394
Velocity Ratio .686 .673 .719 .757 .528 .670 .694
Jet Energy .598 .499 .281 .523 .66 .692 .607

b) Correlation with Specific Energy


Jet Pressure .139 .160 -.504 .368 .240 .232 .268
Nozzle Diameter -.354 -.246 .216 0.123 -.123 -.287 -
No. of Passes .653 .618 - - - - -
Traverse Speed -.398 -.474 .172 -.130 -.382 .592 .469
Velocity Ratio .572 .654 .063 .021 .605 .835 .518
Jet Energy -.036 .040 -.428 -.346 .096 .093 -.259

*Rock symbolized by the number given in Table 1.


Table 3. Multiple Regression Correlation Coefficients for the Most Influential Independent
Variables

Correl. Correl.
Rock Depth Coef. Specific Energy Coef .
Berea Sst. Vel rto*, Press, Pasno0.85 Velrto, Diam, Pasno 0.84
Indiana Llst. Velrto, Press, Pasno 0.90 Velrto, Diam, Pasno 0.84
Red Sst. Velrto, Press, Tips 0.91 Velrto, Depth, Tips 0.82
Pink Sst. Velrto, Press, Diam .86 Velrto, Depth, Diam .92
Buff Sst. Velrto, Press, Diam .85 Ins, Tips, Depth .78
Georgia Mbl. Velrto, Diam .91 Press , Depth .40
Barre Gnt. Velrto .71 Press , Depth .61
Velrto = Jet velocity/nozzle traverse speed Diam = Nozzle diameter (ins)
Press = Jet pressure (psi) Tips = (1/traverse speed)2 (min/ft)2
Pasno = Number of passes made Ins = 1/traverse speed (min/ft)
Depth = depth of hole (ins)
From the analysis the jet velocity ratio was found to have the highest correlation with the depth of
slot cut for all rock types. Analysis of the specific energy data indicated that the jet velocity ratio
had the highest correlation coefficient for the sedimentary rock, but that the jet pressure had a
higher correlation with specific energy for the crystalline rock. This would suggest a possible
difference in the cutting mechanism of the jet in the two rock groups. Such a difference has been
observed in the test specimens. Granular material is eroded by the jet, with the particles of the
material remaining substantially intact. Crystalline rock suffers intra- and intercrystalline failure
under water jet impact and the particles removed are fragments of the original crystalline structure.
This conclusion was confirmed in subsidiary tests carried out on specimens of Missouri Granite
and Georgia Granite.

The slot depth was reduced with an increase in the nozzle traverse speed (Fig. 5), but the specific
energy of rock removal was also decreased (Fig. 6). Thus, although less depth was achieved, a
greater volume of rock was removed. The fact that the faster the nozzle is traversed, the more
efficiently the jet cuts, makes it somewhat rare among cutting devices.

Figure 5. Depth of slot cut versus nozzle speed.

An increase in the nozzle diameter increased the depth of the slot cut and reduced the specific
energy required to cut the sedimentary rock. Because a higher initial pressure, around 15,000 psi,
was required to cut the crystalline rock, there was not enough data to draw conclusions on the
effect of nozzle diameter.

A comparison of jet performance at maximum pump output (Figs. 7 & 8) for the three nozzles
showed that, for the sedimentary rock, a larger nozzle will cut more effectively than a smaller one
but the depth of the slot achieved will not always be greater. For the three nozzles tested the 0.03
inch diameter proved to give the best results despite the fact that the maximum jet pressure obtained
was 18,000 psi relative to the 25,000 psi obtained through the 0.023 inch diameter nozzle.

Rock Properties.
Six rock properties were measured for each rock tested (Table 1). Uniaxial compressive
strength and Young's modulus at 50% strength were measured using an LVDT strain jacket. The
specific fracture energy, as defined by Cooley (11) was also calculated. Shore scelerescope and
Schmidt hammer values were determined as non-destructive measurements of the rock hardness,
and the Rock Impact Hardness Number (RIHN) was measured as a destructive hardness value.
The Fracture Toughness (FT) of the rock was also measured.

Figure 6. Specific Energy of breakage versus nozzle traverse speed. (Buff sandstone).

Figure 7. Depth of slot cut versus nozzle diameter (Pink sandstone).


Figure 8. Specific energy of breakage versus nozzle diameter (Pink sandstone).

The measurements of the hardness made with the Schmidt hammer were carried out on six
inch cubes of rock, three times the size recommended by the manufacturers (25). Nevertheless a
noticeable edge effect on the values obtained was noted (Fig. 9). Approximately 250
measurements of hardness were made for each rock. Measurements of the Shore scelerescope
hardness were made with two inch cubes of rock with 360 readings taken for each rock type. The
readings were made over three mutually perpendicular surfaces of the rock and an average result
obtained.

Figure 9. Contour map of Schmidt hammer values measured over a 6 inch cube of Pink sandstone.

The RIHN value was determined as the number of blows required to reduce a two inch long,
one inch diameter core to give 25% of the crushed weight less than 40 mesh (26). The blows were
delivered by a 2.4 kg. hammer falling 64 cms. onto the core, which had been laid on its side in a
recessed anvil.

The FT measurements were made using pre-notched beams of rock broken in 3-point bending
(27). The values obtained by this method gave a much more detailed measure of the local strength
of the rock, the value obtained being for the mineral at the point of the notch only. Fifty values
were obtained for each rock. The tests were made on beams prepared from each side of the cubic
samples of rock as received from the quarry. The value used in the analysis was an average of the
values obtained for the three directions.

In the granite specimens, spallation and total rib removal occurred between adjacent slots
under certain test conditions. Although this phenomenon was consistent above certain test levels,
for example Missouri granite spalled around the cut when the depth exceeded one third of an inch,
this enlarged crater was not considered in the analysis. The reason for this was that spallation may
have been increased by the presence of an adjacent free surface, and also that the parameters
controlling the spallation were not completely included in the analytical calculations.

Depth, specific energy and specific energy ratio, the ratio of specific energy of rock breakage
to fracture energy, values were correlated with the rock properties both directly and with the
inverse of the rock properties (Table 4). A correlation matrix was also obtained between the
various rock properties (Table 5).

The depth of the slot cut had the highest correlation with the inverse of the rock compressive
strength, followed by the Fracture Toughness of the rock. Specific Energy correlated best with the
Rock Impact Hardness Number and the inverse of the compressive strength. The specific energy
ratio correlated best with the Young's modulus and the inverse of the Rock Impact Hardness
Number.

Table 4. Pearson Correlation Coefficients between Water Jet Cutting Parameters and Rock
Properties
Specific
Specific Energy
Rock Property Correlation Coefficient with Depth Energy Ratio
Compressive Strength (CS) -.249 .470 .030
Young's Modulus (YM) -.066 .479 .519
Shore Hardness (SHOR) -.013 .286 -.060
Schmidt Hardness (SCHH) -.079 .434 .081
Rock Impact Hardness Number (RIHN) -.279 .597 .278
Fracture Toughness (FT) -.312 .494 .242
Specific Fracture Energy (SFE) -.223 .379 -.099
Inverse Compressive Strength .352 -.547 -.223
Inverse Young's Modulus -.004 -.496 -.495
Inverse Shore Hardness -.078 -.193 .063
Inverse Schmidt Hardness -.028 -.412 -.221
Inverse RIHN .052 -.505 -.446

Table 5. Correlation Matrix between Rock Properties

Properties* CS YM SHOR SCHH RIHN FT SFE


CS 1 .45 .82 .90 .85 .61 .98
YM .45 1 .43 .55 .62 .39 .28
SHOR .82 .43 1 .82 .59 .36 .84
SCHH .90 .55 .82 1 .73 .34 .87
RIHN .85 .62 .59 .73 1 .88 .75
FT .61 .39 .36 .34 .88 1 .51
SFE .98 .28 .84 .87 .75 .51 1

Properties are symbolized by the abbreviations given in Table 4.

Regression Analysis.
The calculation of the Pearson product-moment correlation coefficients provided a measure of
the accuracy of prediction of the penetration parameters based on individual independent variables.
However, although the properties for the various rocks were initially considered as independent of
each other, this is not necessarily the case (Table 5). Because there is a high correlation between
some of the properties considered, a step-wise regression was performed. This regression was
performed for each of the three dependent variables in turn to determine which jet and rock
properties in combination provided the best correlation with the dependent variable.

In a step-wise regression (21), a linear regression equation is calculated between the dependent
variable and the independent variable which is found to be the best predictor. The correlation
coefficients are then recalculated to determine which independent variable gives the best prediction
when combined in the equation with the first variable. This variable is added to the regression
equation and the procedure is repeated, adding a single independent variable in each step of the
regression. The program used also gives a value of the tolerance. A low value for tolerance
indicates that the variable considered is a linear combination of variables already present in the
regression, and should therefore be disregarded.

Although the regression was continued until no improvement in the regression was achieved,
only the first four independent variables are included for the first two equations below. The
addition of further terms to the regression equation did not improve the multiple regression by
more than 2 percentage points. Six variables are combined in the third equation for the same
reason.

The equations obtained were:


i) DEPTH = 05.72 + 24.8 x 10-4 (VR)2 + l6xlO 3 (CS)-1+ 3 x 10-5 (P) - 7.4 x 1O -2 (SHOR)

The regression coefficient units should be chosen to give depth values in inches.

ii) SPECIFIC ENERGY = -99.4 x10 3 + 8.19x103 (RIHN) + 84.3xlO 7 (CS)-1+1.97P-38.2(SFE)-1

The regression coefficient units should be chosen to give specific energy values in joules/cm3

iii) SPECIFIC ENERGY RATIO =-1228 - 1.56(VR) + 7.4x106 (YM)-1 - 0.248(CS) + 177(RIHN)
+ 6.1x10-4 (YM) - 2. O(FT) - 42 .4xlO3 (SHOR)-1

The regression coefficient units should be chosen to give dimensionless units.

These equations are written with the variables listed in the order obtained, apart from the
grouping of pressure terms in equation (ii). The equations are empirical descriptions based on the
test results obtained and should not be considered as absolute equations, but rather as a guide. The
statistical data for the equations are:

Multiple Regression Standard Error


Coefficient of Estimate F Value
Depth .72 .35 ins. 413
Specific Energy .89 4978joules/cm 3 1531
Specific Energy Ratio .91 122 1133

Concluding Remarks.
Three regression equations have been developed, based on experimental data, to describe the
water jet cutting of rock. The equations are constructed as a combination of water jet and rock
properties combined to predict the depth of the slot cut in the rock, and the specific energy of
breakage of the rock. The third equation describes the properties which control a dimensionless
measure of jet performance. The specific energy and dimensionless ratio equations had a multiple
regression coefficient of around 0.9, indicating that it should be possible to predict jet performance
with a previous knowledge of the rock and jet properties. The analysis did not, however, include
any data on spallation. Because spallation has been found to occur consistently under certain test
conditions, the results predicted in the equation are likely to give a high estimate of the energy
required to cut rock.

2
Abbreviations as in Tables 3 and 5.
The specific energy equation indicates that, where rock properties are held constant, the
energy required to remove a given volume of rock will increase with increasing jet pressure.
Specific energies below 200 joules/cm3 were obtained in cutting sandstone at high traverse speeds
and low jet pressures. This low energy value would be further reduced to below 50 joules/cm3 if
mechanical means were used to excavate the rock between adjacent jet slots. Data from tests on
crystalline rock is not yet sufficiently comprehensive to enable a prediction as to the best pressure
range for the water jet to operate in to cut these rocks.

Acknowledgments.

This research is part of a current program being funded by the Department of Defense Contract
No. DACA-45-69-C-0087, to which agency thanks is given. The experimental work was carried
out with the assistance of Mr. R. L. Henry, Mr. T. G. Thompson, and Mr. D. Heisler. Mr. D.
Gaitros gave much helpful advice on the SPSS program. The test equipment was constructed with
the help of Mr. B. Hale and Mr. C. Rapier. The author wishes to express his sincere appreciation
to these gentlemen.

References.
1. Maurer, W. C., Novel Drilling Techniques, Pergamon Press, 1967, 117 pp.
2. Proceedings, First Int. Symp. Jet Cutting Technology, held in Coventry U.K., April 5 - 7,
1972, to be published by BHRA, Cranfield, Bedford, U.K., Summer 1972, 600 pp.
3. Torpey, P., "Some Experiences in the Manufacture and Application of High Pressure Water
Cleaning Equipment," paper Dl, Proc. 1st Int. Symp. Jet Cutting Technology, Coventry,
1972.
4. Ward, G. M., "Safety Considerations Arising from Operational Experience with High Pressure
Water Jet Cleaning," paper Fl, Proc. 1st Int. Symp. Jet Cutting Technology, Coventry, 1972.
5. Wick, C. H., "Deburring with High Pressure Water Jets," Manufacturing Engng &
Management, August 1971.
6. Walstead, 0. M. & Noecker, P. W Development of High Pressure Pumps and Associated
Equipment for Fluid Jet Cutting," paper C3. Proc. 1st Int. Symp. Jet Cutting Technology,
Coventry, 1972.
7. Pliny, XXXIII, p. 21.
8. Farmer, I. W. & Attewell, P. B., "Rock Penetration by High Velocity Water Jet: A Review of
the General Problem & An Experimental Study," Int. J. Rock Mech. & Min. Sci_ 2, pp.
135-1531, 1964.
9. Bresee, J. C., Cristy, G. A.9 & McClain, W. C., "Some Comparisons of Continuous and
Pulsed Jets for Excavation," paper B9. Proc. 1st Int. Symp. Jet Cutting Technology,
Coventry, 1972.
10. Crossland, B. & Logan, J. G., "Development of Equipment for Jet Cutting," Proc. 1st Int.
Symp. Jet Cutting Technology, Coventry, 1972.
11. Cooley, W. C., "Correlation of Data on Erosion and Breakage of Rock by High Pressure
Water Jets," Proc. 12th Symp. Rock Mechanics, Rolla, 1970.
12. Singh, M. M. & Huck, P. J., "Correlation of Rock Properties to Damage Effected by Water
Jets," Proc. 12th Symp. Rock Mechanics, Rolla, 1970.
13. Teale, R., "The Concept of Specific Energy in Rock Drilling," Int. J. Rock Mech. & Min.
Sci., 2, p. 57-74, 1965.
14. Summers, D. A., Disintegration of Rock ~y High Pressure Jets, Ph.D. Thesis, University of
Leeds, England, 1968.
15. Summers, D. A.& Henry, R. L., "Water Jet Cutting of Rock With and Without Mechanical
Assistance," SPE 3533, Fall Meeting of Soc. of Petr. Engrs. 1971.
16. Summers, D. A. & Henry. R. L., "The Effect of Change in Energy and Momentum Levels on
the Rock Removal Rate in Indiana Limestone," paper B4 Proc. 1st Int. Symp. Jet Cutting
Technology, Coventry, 1972.
17. Henry, R. L. The Penetration of Continuous High Pressure Water Jets into Rock, M.S.,
Thesis, University of Missouri-Rolla, May 1972.
18. Franz, N. C., "Fluid Additives for Improving High Velocity Jet Cutting," paper A7 Proc. 1st
Int. Symp. Jet Cutting Technology, Coventry, 1972.
19. Brook, N. & Summers, D. A., "The Penetration of Rock by High Speed Water Jets," Int. J.
Rock Mech. & Min. Sci., 6, pp. 249-258s 1969.
20. White, A., "Studies on the Flow Characteristics of Dilute Polyox Solutions," Symp. on
Turbulent Suppression, with Reference to Polyox Type Additives, Bradford, 1968.
21. Nie, N. H., Bent, D. H., & Hull, C. H., Statistical Package for the Social Sciences,
McGraw-Hill Book Company, 1970, pp. 343.
22. Brook, N., Personal Communication, 1971.
23. Leach, S. J. & Walker, G. L., "The Application of High Speed Liquid Jets to Cutting. Some
Aspects of Rock Cutting by High Speed Water Jets," Phil. Trans. Roy. Soc. (Lond), A260,
pp. 295-308, 1966.
24. Rochester, M. C. & Brunton, J. H., "High Speed Impact of Liquid Jets on Solids," paper Al,
Proc. 1st int. Symp. Jet Cutting Technology, Coventry, 1972.
25. Soiltest, Inc., "The Concrete Test Hammer," reprint from Concrete Constr. Mag.
26. Brook, N. & Misra, B., "A Critical Analysis of the Stamp Mill Method of Determining
Protodyakonov Rock Strength and the Development of a Method of Determining a Rock
Impact Hardness Number," Proc. 12th Symp. Rock Mechanics, Rolla, 1,970.
27. Morris, W. J. & Summers, D. A., "Influence of Initial Fracture Angle on the Apparent
Fracture Toughness of Blunt Notched Specimens," in preparation.

You might also like