Вы находитесь на странице: 1из 9

Proceedings of the ASME 2016 International Manufacturing Science and Engineering Conference

MSEC2016
June 27-July 1, 2016, Blacksburg, Virginia, USA

MSEC2016-8775

NEAR NET SHAPING OF THICK MATERIALS WITH AWJ

Mohamed Hashish
Flow International Corporation
Kent, WA, USA

ABSTRACT metals, composites, stone, and glass. More on AWJ tool can be
This paper presents results on cutting of relatively thick found in (1, 2).
materials such as titanium, steel, and glass with abrasive
waterjets (AWJ). A physical kerf line model was found to
correlate well with experimental trail-back data. Also, a physical
kerf width model is proposed based on waterjet spreading
characteristics which may be used for kerf taper control. The
measurement of two dimensional edge cuts was used to
determine lead and taper angles for improved geometrical
results. It was found that relatively large power jets are needed
to insure cutting faster and with minimal wall waviness. For
example, 200 kW jets were used to cut glass honeycomb shapes
up to 600-mm thick with thin (~ 1-2 mm thick struts) out of a
solid block. Titanium and steel up to 300-mm thick were
precisely cut for jet engine applications. Both taper and
undercutting around corners were either minimized or eliminated
using jet lead and taper angle control.

INTRODUCTION
AWJ Process
Figure 1 shows a schematic of an AWJ nozzle. In the nozzle,
water is pressurized up to 600 MPa and expelled through an
orifice, typically made out of diamond or sapphire, to form a Figure 1. AWJ Nozzle and Parameters
high-velocity waterjet. Typical jet diameters are 0.1 mm. to 0.7
mm. and typical jet velocities are 500 to 1000 m/s. The flow of Thick Material Cutting
the high-velocity waterjet into the concentrically aligned mixing One of the important advantages of abrasive-waterjets (AWJ)
tube creates a vacuum, which is used to transport abrasives from over other beam-like cutting tools such as laser is their ability to
a hopper to the nozzle. A typical abrasive material is garnet, cut relatively thick materials whether hard as steel, glass, and
which has a typical flow rate of about 8 g/s. Medium and fine stone or soft as foam and plastics. There are a few methods by
abrasives (mesh 60 to mesh 220) are most commonly used. The which an AWJ can cut relatively thick (~> 50mm – 600 mm)
abrasives are accelerated and axially oriented (focused) in the materials. These are:
mixing tube made out of a tungsten carbide material in most Single pass cutting: In this case, a jet is used to cut
cases. It has a length-to-diameter ratio from 50 to 150 with through the material in one pass. Typically, the thicker
length up to 100 mm in most commercial applications. Typical the material, the higher the power of the jet that will be
tube diameters range from 0.5 mm to 2 mm or 3 to 4 times the needed to achieve the machining results. This paper is
diameter of the waterjet orifice. Process parameters are shown focused on this approach.
in Figure 1. The AWJ is capable of cutting any material such as

1 Copyright © 2016 by ASME


Multiple pass cutting: In this technique, several passes
are used to groove deeper into the material. The
standoff distance increases with every successive pass
and accordingly the cutting capability will be reduced.
It was observed (3, 4) that there is an optimal
combination of traverse rate and number of passes to
achieve maximum cutting depth or highest productivity.
This topic will not be covered in this paper.
A penetrating tool may be needed to make deep cuts.
However, the kerf must be widened in order to allow
the jet nozzle to enter the kerf. Widening the kerf may
be accomplished by rotating or oscillating a slightly
angled jet while traversing. With this method, over 1-
meter deep cuts were made in nuclear-grade reinforced
concrete using 100 kW AWJ producing 0.5-1.0 m2/hr Figure 3: Trail Back
cut area rate. This was used as a tool for
decommissioning the west valley nuclear services The second phenomenon is that the width of the cut varies along
facility in NY. This method will not be covered in this the cut from top to bottom. This difference in width is typically
paper, but the reader may refer to (5). called the taper of the cut. A taper can be either positive or
negative, that is, the width at the exit of the cut may either be
Figure 2 shows examples of AWJ cuts made in thick materials smaller or larger than the width at the top. Typically, the kerf
using the above methods. In this paper, we will start first by width at the exit side is smaller than that at the entry at practical
discussing the general features of AWJ cuts as pertains more to cutting speeds. Figure 4 shows cuts with different taper.
cutting thick material. This will be followed by discussing some
theoretical aspects of deep cutting. Some cutting observations
will then be presented for glass, steel, and titanium. Conclusions Speed
are then presented at the end.

Divergent Straight Convergent


Cut in 300-mm Cuts in 300- Penetrating AWJ Tool
thick Titanium mm thick in 1-m thick concrete Figure 4: Taper Examples in Steel
glass
Figure 2: Examples of AWJ Cuts made with Different For shape cutting, the trail back and taper phenomena manifest
Techniques themselves in distortions to the geometry of the cut at the exit
side. The sketch in Figure 5 shows an undercut due to the trail-
AWJ KERF GEOMETRY back phenomena. The picture in the same figure shows distorted
As the abrasive waterjet (AWJ) cuts through and separates the square-shaped cuts at the bottom surface of the material due to
material, three phenomena are observed (6-9). The first is that trail-back and taper.
the jet is deflected opposite to the direction of the motion. This
means that the exit of the jet from the material lags behind the
point at the top of the material where the jet enters. The distance
by which the exit lags the entrance is typically called the trail-
back, lag, or drag as shown in Figure 3. Observe that the jet-
material interface is a curved surface.

2 Copyright © 2016 by ASME


are related to the steadiness of the parameters such as
pressure and abrasive flow rate.
Kinematic Issues: Deep cutting, especially when
geometric precision and low surface waviness are
UPPER required, is more sensitive to the quality of the motion
SURFACE
system meeting accurate traverse rates with minimal
JET PATH
deviations. Also, if jet tilting (11) is needed as may be
required for three dimensional cutting or for kerf shape
compensation, then the accuracy and strategy of jet
tilting must be considered. Jet angulation parameters
UNDERCUT can be determined experimentally by measuring the no-
tilt two dimensional trail-back and width of cut profiles
and then mathematically determining the lead and taper
LOWER angles to minimize deviations. Figure 7 shows short
SURFACE
MH009
straight nibble cuts used for this test. Also, short shape
nibble cuts are used to determine the undercut
Figure 5: Undercutting at the Bottom of Cuts parameters.

The third phenomenon is related to the surface finish of the cut.


Due to the transient nature of the jet penetration process and jet
instability, striations (waviness) will form along a cut surface
below an initial rough but waviness-free zone. Figure 6 shows
typical striated (wavy) surface produced by AWJ.

Straight Nibble Cuts in 100-mm Thick Steel

Figure 6: Surface Topography

BASIC STRATEGY
The basic strategy for relatively accurate cutting of thick
materials is related to both process and kinematic issues:
Process Issues: The AWJ process parameters should be
selected to cut the required depth at the required speed
and surface quality. For a given jet power, the speed of
cut can be determined based on the required surface
quality as typically used in current practice. Typically,
deeper cuts require larger jets and thus higher levels of
hydraulic power and abrasive flow rates. A special Shape Nibble Cuts in 100-mm Thick Glass
AWJ nozzle design (10) may be needed to maximize Figure 7: Straight and Corner Nibble Cuts to
the use of the available power. For example, a mixing Characterize Kerf Geometry
tube length of 600-mm was used to precisely cut 300-
mm thick glass using 200 kW jet. Other process issues

3 Copyright © 2016 by ASME


OBSERVATIONS AND THEORETICAL TREATMENT more significant centrifugal abrasion zone. At location “s” along
the kerf, Figure 9 where the radius of curvature is r, the radial
In this section, we present experimental observations and some erosion velocity is𝑟̇ .
simplified models.

Leading Kerf Shape


As mentioned above, the trail-back phenomenon becomes more
significant as depth of cut increases. Figure 8 shows trail-back
data for cutting 300-mm thick titanium 6AL-4V at different
conditions. These conditions are given in table 1 below:

Table 1: Traverse Rate and Abrasive Flow Rates for


Tests in Figure 8
ma u (mm/min)
(g/s) 2.54 3.81 5 6.35
Tests 1 Tests 8 &
22.5 Test 2
&7 14
30.0 Test 9 Test 10
30.4 Test 3
32.0 Test 5 Test 4
Tests 12
37.5 Test 11
& 13

Observe that the trail-back curve is generally circular in nature


and that the radius of curvature is dependent on the AWJ process
parameters. A simple model for the shape of the trail back is a
parabola in the form tb= k x2, where k is a constant for every
Figure 9. Model Parameters
curve and x is the depth. Changes from one curve to another (k
value) depend on the process parameters such as traverse rate, A hypothesis for the erosion velocity 𝑟̇ is that it depends on the
abrasive flow rate, and jet structure. Studies (12-14) have also
normal stress 𝜎 , the coefficient of friction , particle velocity
addressed modeling studies of this kerf shape.
𝑉 and the material resistance 𝜎  as can be expressed below:

30    
p= 345 MPa
sod= 2.5 mm
1
2
𝑟̇ =   (1)
dn= 0.64 mm and 0.51 for test 7 3
25 dm= 305 mm
152 mm for test 7
4
5
The normal force and particle mass, are related to the angular
457 mm for tests 13 & 14
dm= 3.18 mm;
9
10
acceleration 𝑉 /𝑟 through  Newton’s  equation  of  motion.    From  
20 1.78 mm for test 7 11
this, the normal stress can be expressed as:
Trail-Back (mm)

4.06 mm for tests 13 & 14 12


Garnet 8
Mesh 50 13
15 14 ̇    
𝜎 =   (2)
 
10

𝑑 in the above equation represents the width on the kerf which


5
is assumed to be equal to the mixing tube diameter.
Also, the traverse speed u can be related to the radial erosion
0 velocity 𝑟̇ and the horizontal kerf angle as follows:
0 50 100 150 200 250 300 350
Depth (mm)
̇
Figure 8. Trail-back Geometry for Several Cuts in
𝑢=   (3)
Titanium (see Table 1 for test cut parameters)
Where cos 𝜃 is 𝜕y/𝜕s. Using the above relationships, and 𝜕y/𝜕𝑥
Crow and Hashish (15) developed a universal AWJ kerf equation = tan 𝜃,  x can be expressed as the trail back (tb) at depth (y= h)
by dividing the kerf zone into an upper direct impact zone and a as follows:

4 Copyright © 2016 by ASME


tb u hu
ln sec (4) 20

0
15 1 degree
2 degrees
Where: 3 degrees
3.39 degrees
ma Va2
10

Trail-Back (mm)
(5)
f dm 5

The above equation does not include the effect of particle


velocity Va decay as depth increases and accordingly Va in the 0

above equation is the velocity at the exit of the mixing tube 0 50 100 150 200 250 300 350

which is a function of the AWJ parameters as will be described -5


later. The above equation does not also account for with of kerf
variation which was assumed to be fixed at 𝑑 . -10
The above formula was used to normalize the data in Figure 8 Depth (mm)
and to check the results against the universal kerf shape form.
Figure 10 shows that there is an excellent correlation between Figure 11: Effect of Lead Angle on Kerf Shape and
the above equation and experimental data. The coefficient of Trail-Back
friction in the above equation was adjusted to compensate for
errors resulting from ignoring variation of Va with depth. Kerf Width
Hashish and DuPlessis (16), based on a jet-spreading model by
0.04 Yanaida (17), expressed the shape of a cut produced by a pure
Theory waterjet as follows:
1
0.03
2
3 2/3
4 we X X
Non Dimensional Trail-Back

c
0.03
5 0.335 1 (6)
0.02
9
10 dn R Xc 2 P1 X c
11
12
8
0.02 13 Figure 12 shows a plot of the effective jet profiles in terms of the
7
14 non-dimensional numbers of Equation (6):
0.01

0.01

0.00
0.00 0.05 0.10 0.15 0.20 0.25 0.30
Non Dimensional Depth

Figure 10: Normalized Kerf Shape Data and


Comparison to Universal Kerf Shape

In order to minimize the effect of trail back on the cut geometry,


jet angling with a lead angle will be needed. Figure 11 shows
data for 300-mm thick titanium cutting at normal impact angle
and at different lead angles. Observe that at a lead angle of
approximately 3.39 degrees, the maximum trail back occurs near
the middle of the part thickness. The amount of this trail-back is
about 4-mm at this angle versus 17-mm when no lead angle is
used. This dramatic reduction in trail-back is of course due to
the shape of the jet-material interface curve. For shallow cuts,
where this shape is close to a straight line, the trail-back can be
reduced to nearly zero when the appropriate lead angle is used.
This is of critical importance when machining precise parts.
Figure 12: Theoretical Waterjet Kerf Profiles

5 Copyright © 2016 by ASME


These non-dimensional numbers are: applying a tilt angle of one degree, the taper was reduced on one
we side of the cut from about 2-mm to less than 0.5-mm. The
Width number = , accuracy of tilting the jet should be in the order of 0.1 degree in
dn R order to accurately obtain the desired wall straightness. Observe
the other side of the cut which is now more divergent to about 4-
c
Material strength number = , mm deviation.
P1 5.0

X 4.0
Standoff distance number =
Xc 3.0
P= 379 MPa, dn= 0.4 mm, dm= 1 mm,
lm= 76 mm, ma= 3.75 g/s, u= 0.008 mm/s

It has been observed that the AWJ-produced kerf profiles are 2.0

similar to those shown in Figure 12. For example, Figure 13

Kerf Width (mm)


1.0
shows kerf width data for 300-mm thick titanium at the same
0.0
conditions shown in Figure 8 above. Observe that most of the -20 0 20 40 60 80 100 120
kerf profiles are converging as the traverse rate was relatively -1.0

high to produce a divergent shape profile. -2.0

2.5 -3.0

Kerf Profile Before Rotation


2.0 -4.0
Kerf Profile After Rotation by 1 Degree

-5.0
1.5
Depth (mm)
1.0 1
2
Width Profile (mm)

0.5 3
Figure 14: Kerf Width Profile Rotation in 100-mm
4 Thick Steel
0.0 5
0 50 100 150 200 250 300 7 350
-0.5 8 Abrasive Acceleration and Nozzle Design
-1.0
9
10
The need for a relatively large AWJ nozzle for precision cutting
11 of thick materials has been identified. This need is because
-1.5 12 relatively large abrasive flow rates (up to 70 g/s) are needed and
13
-2.0 14 thus demanding larger size mixing chamber and mixing tube
-2.5
diameter. A simple analysis for the abrasive particle velocity has
Depth (mm) been addressed (1, 10). Based on the momentum and continuity
equations of the liquid and solid flow, the following equation
Figure 13: Kerf Width Profile in 300-mm Thick results for the particle velocity Va at a distance x inside the
Titanium mixing tube:

Based on this similarity, it is proposed here that the above 1 1


equation be applied to AWJs with some modifications. The jet x ln (8)
hydrodynamic force can be replaced by the abrasives rate of K 1 1
change of momentum, maVa, and the material strength, c, can Where
be expressed using the material erosion resistance f . With this 2
adaptation, the following equation results:
3 CD 1 r Va Va
K , (9)
4 Sa d p Vamax Vj 1 r
2/3
we X d m2 f X
0.335 1 (7)
dm R Xc 8m a. V a2 Xc Where r= ma m w is the abrasive loading ratio. Table 2 shows
the mixing tube length for different cases for a mixing efficiency
of 90%, i.e., = 0.9.
The Characteristic length Xc is to be determined experimentally.
Using the above equation for predicting kerf shape and
Observe that larger particles require longer mixing tubes; note
comparing results to experimental results is a subject of current
also that as the abrasive flow rate increases, shorter tubes can be
investigation.
used to attain the maximum velocity. For the commonly used
100 mesh abrasives, for example, a mixing tube length of only
Similar to rotating the kerf shape using a lead angle, the kerf
33-mm is required for an abrasive loading ratio of 0.12. The
width profile can be rotated to minimize taper on one side of the
typical mixing tube length used in industry for this case,
cut. Figure 14 shows a typical kerf width shape for 100-mm
however, is about 76-mm. The additional length is used to
thick steel at low traverse rate where the cut diverges. By

6 Copyright © 2016 by ASME


collimate the jet and to raise the value of to about 0.95, as can
be calculated from the above equations. Note that the additional
43-mm of mixing tube length contributes only 5% to the
maximum possible velocity.

Table 2. Mixing Tube Length for = 0.9


r=
Mesh m a mw 0.10 0.12 0.15 0.2 0.25
No.
dp (mm) Mixing tube length, lm (mm)
16 1.65 439 423 401 369 340
36 0.76 202 195 185 170 157 Figure 16: Titanium Blisk Roughing with AWJ
60 0.38 101 98 93 85 78
80 0.25 67 65 62 57 52 A concept of rapid roughing using multi-axis manipulator such
100 0.13 34 33 31 28 26 as a 6-axis robots is shown in Figure 17. A robot dripped may
hold a part and articulate it under a stationary AWJ to near net
Figure 15 shows AWJ nozzles used for deep cutting. The mixing shape it within the limits of the robot arm and process accuracies.
tube lengths for these nozzles are 300-mm and 450-mm. Again, the advantage in this case to minimize chip cutting.
Observe that two abrasive feed lines were used to transport
abrasives to the cutting head as one line is not large enough for
feeding the required flow of abrasives.

Figure 15: Examples of AWJ Nozzles for Cutting


Thick (over 300-mm) Materials

SELECTED APPLICATIONS
Tests were performed to determine the feasibility of roughing
thick (150-mm) Titanium blisk. Figure 16 shows a segment used
for testing. To obtain optimum geometrical accuracy and surface Figure 17: Roughing of Inconel Injector Nozzle
finish and minimum cycle time, a sufficiently accurate 5-axis
gantry style manipulator and 600 MPa pump pressure were used. In another example, Figure 18, a large glass block 300-mm deep
Robotic manipulators were also used in actual production. It was was light-weighted for a space telescope mirror core by cutting
found that a sacrificial cut must be made to unlock the parts cut out triangular shapes, leaving about of 2-mm thick struts. Jet
between any two blades. Standard milling machines are used to lead and taper angles were determined to minimize undercutting
finish the roughed-out blisks. A rough economic analysis and strut thickness variation to less than 0.5-mm. This means
suggests that AWJ roughing is highly competitive due to the fact that the wall straightness should be within 0.25-mm. One of the
that no chips are produced and that the produced slugs have challenging tasks for cutting fragile thick materials is handling
residual value. the cut-outs. Due to their relatively large weight, a weight
transfer method must be used. This can be achieved by either

7 Copyright © 2016 by ASME


clamping the cutout using a simple clamp or using frictional CONCLUSIONS
inserts inside the kerf.
The following conclusions can be drawn from this work:
Abrasive-waterjets have been demonstrated for cutting
of thick materials such as titanium, glass and steel.
A certain level of jet power needs to first be determined
to cut at the required speeds and surface finish. Jet
angulation is then used to overcome kerf taper and
undercutting.
Jet lead angle is the key parameter for reducing the trail-
back and the associated undercutting anomalies around
corners and tight curves.
Kerf wall taper can be significantly improved using jet
taper angles. For thick materials, this taper angle needs
Glass Block and a Large AWJ Nozzle to be accurately controlled to within 0.1 degree
A simple physical model has been developed and
shown to correlate well with experimental data of kerf
shape.
A kerf width model has been proposed in this paper.
This model will be tested against experimental kerf
width data in a future study.

NOMENCLATURE

CD Drag coefficient
Hexagonal Honeycomb Residual pieces of glass removed dm Mixing tube diameter
Shape Cut in Glass from block dp Particle diameter
Figure 18: Light-weighting of Thick Glass for Mirror dn Nozzle diameter
Cores K Defined in equation (4)
9
lm Mixing tube length
The capability of the abrasive-waterjets for cutting relatively
thick sections of steel offers a great potential for rapid fabrication mp Particle mass
of automotive tooling molds. This process can deliver near net ma Abrasive flow rate
shape tooling in fraction of the time currently being used to cast mw Water flow rate
or rough a mold cavity. This process involves the assembly of
P Pressure
an array of uniquely profiled sections of steel to form the mold
core or cavity. Also, this method allows for the incorporation of P1 Jet dynamic pressure at nozzle exit
conformal cooling (or heating) passages. The profiling and r Abrasive loading ratio
beveling of each steel section is accomplished with an abrasive
R Ratio of Xc to dn
waterjet. Figure 19 shows cuts in 100-mm thick aluminum to
form a 3-D laminated part to illustrate the concept. Sa Abrasive particle specific gravity
sod Standoff distance
tb Trail back
u Traverse rate
U Water velocity
Abrasive-solid friction coefficient
Defined in equation (2)
5
Va Abrasive particle velocity
Vj Waterjet velocity
Vmax Maximum abrasive velocity
Figure 19: AWJ Cuts for 3-D Laminated Object we Effective jet diameter /kerf width

8 Copyright © 2016 by ASME


x Distance along mixing tube length International Conference on Water Jetting, BHR, Aix-en-
Province, France: October 16-18, pp. 211-224
Particle velocity ratio
13. Henning,  A,  Goce,  E.,  and  Westkamper,  (2002)  “Analysis  
X Distance along the jet and Control of Striations Structure at the Cutting Edge of
Xc Length of initial jet region Abrasive   Waterjet   Cutting,”   Proceedings   of   the   16 th
International Conference on Water Jetting, BHR, Aix-en-
f Density of AWJ fluid
Province, France: 16-18 October, pp. 173-191.
c Material strength 14. Henning,  A.,  Wetkamper,  E.,  (2000),  “Modeling  of  Contour
f Material erosion (flow) strength Generation   in   Abrasive   Waterjet   Cutting,”   Proceedings   of  
the 15th International Conference on Water Jetting, BHR,
REFERENCES Ronneby, Sweden, September 6-8, pp.309-320.
1. Hashish,  M.  (2003),  “Inside  AWJ  Nozzles,”  Proceedings  of   15. Crow,  S.,  and  Hashish,  M.  (1989)  “Mechanics  of  Abrasive  
the 2003 WaterJet Conference, WJTA, Houston, TX, Jet   Cutting,”   Flow   Research   Presentation, Flow Research
August. Company, Kent, WA, March 31.
2. Hashish,   M.,   (2015)”Waterjet   Machining   Processes”,   16. Hashish,   M.,   and   DuPlessis,   M.   P.   (1979)   “Prediction  
Springer-Verlag London 2015, A.Y.C. Nee (ed.), Handbook Equations Relating High Velocity Jet Cutting Performance
of Manufacturing Engineering and Technology, Chapter 34, to Stand Off Distance and Multi-passes,”   ASME
pp.1651-1686 Transactions, Journal of Engineering for Industry, Vol.
3. Hashish, M., (1982) "The Application of Abrasive Waterjets 101, No. 3, pp. 311-318.
to Concrete Cutting," Proceedings of the Sixth International 17. Yanaida,   K.   (1974)   “Flow   Characteristics   of   Waterjets,”  
Symposium on Jet Cutting Technology, BHRA, England, pp. Proceedings of the Second International Symposium on Jet
447-464. Cutting Technology, BHR Group, The Fluid Engineering
4. Agus, M., et., al., (2002) "Multipass Abrasive Waterjet Centre, Canfield, England.
Cutting Strategy," Proceedings of the 16th International
Symposium on Jet Cutting Technology, BHR Group,
England, pp. 243-257.
5. Hashish, M., Echert, D., (1989) "Abrasive-Waterjet Deep
Kerfing and Waterjet Surface Cleaning for Nuclear
Facilities," ASME Transactions, Journal of Engineering for
Industry. Volume 111, No. 3, pp. 269-281.
6. Hashish,  M.  (1988)  “Visualization  of  the  Abrasive-Waterjet
Cutting  Process,”  Journal  of   Experimental  Mechanics,  pp.  
69-79.
7. Matsui, S., et., al., (1991), “High  Precision  Cutting  Method  
for Metallic Materials by Abrasive-Waterjet,”  Proceedings  
of the 6th American Waterjet Conference, Houston, Texas,
August 2-4, pp127-137.
8. Henning,  A.,  and  Anders,  S.,  (1998),  “Cutting  Edge  Quality  
Improvement Through Geometrical Modeling, Proceedings
of the 14th International Conference on Water Jetting, BHR,
Brugge, Belgium, pp. 321-328.
9. Zheng, X., Chen, F., Siores, E., Steele, P., Biskup, C., and
Pude,  F.  (2002),  “Two  Dimensional  and  Three  Dimensional  
Abrasive Waterjet Machining of Aerospace Composite
Components,”   Proceedings   of   the   16th International
Conference on Water Jetting, BHR, Aix-en-Province,
France: October 16-18, pp. 211-224
10. Hashish,  M.,  (2009)  “Special  AWJ  Nozzles,  ”  Proceedings
of the 2009 American Water Jet Conference, Houston, TX,
paper 4-B
11. Hashish,  M.,  (2007)  “Benefits  of  Dynamic  Waterjet  Angle  
Compensations,”  Proceedings of the 2007 American Water
Jet Conference, Houston, TX, paper 1-H
12. Zheng, X., et., al., (2002),   “Two   Dimensional   and   Three  
Dimensional Abrasive Waterjet Machining of Aerospace
Composite   Components,”   Proceedings   of   the   16 th

9 Copyright © 2016 by ASME

Вам также может понравиться