You are on page 1of 15

Measurements on 5:1 Scale Abrasive Water Jet Cutting Head Models

P.S. Coray1, B. Jurisevic2, M. Junkar2 and K.C. Heiniger1 Swiss Competence Centre of Water Jet Technology, Laboratory for Thermal and Fluid Engineering, University of Applied Sciences Aargau, Northwestern Switzerland
2 1

Laboratory for Alternative Technologies, Faculty of Mechanical Engineering, University of Ljubljana, Slovenia

ABSTRACT
There is a vast potential market for high precision parts manufactured by abrasive water jet (AWJ) machining, which calls for improvement of the current AWJ cutting methods. A reasonable approach is to use models which adjust machining parameters according to the workpiece properties, however detailed knowledge of the physical behaviour of the cutting tool is required. The measurements presented in this paper intend to extend the current knowledge of abrasive water jets, by determining the kinetic energy distribution of the abrasive particles and the structure of the jet in dependence of the cutting head parameters. Initially 5:1 scale models will be examined by using laser and phase doppler anemometry, force measurements and photographic methods. At a later stage comparative measurements and cutting experiments using real scale cutting heads will be performed. The work is still in progress and mainly intermediate results and measurement methods are discussed in this paper. Keywords: Multiphase-flow measurement, abrasive water jet, kinetic energy distribution.

NOMENCLATURE
AWJ CD Cdrag d E g h L
* foc

Abrasive water jet Discharge coefficient, Drag coefficient, Diameter, mm Energy, J Gravity, g = 9.81 m/s2 Specific enthalpy, J / kg Dimensionless focussing tube length, L* = Lfoc / dfoc Laser Doppler Anemometry Laser Induced Fluorescence Mass, kg Mass flow, kg/s Dimensionless abrasive mass flow, m*abr = mabr/mwater Particle Image Velocimerty Particle Tracking Velocimerty Heat flux, J / s Specific entropy, J / (kgK) Scale factor (e.g. 5 for a 5:1 model)
1/15

LDA LIF m m m
* abr

PIV PTV Q, Q s SF

v V, V W W z Subscripts: abr air air-water air-abr AP CV foc o s w.jet water

Velocity, m/s Volume flow, m3/s "Work flux", J / s Altitude, m Density, kg/m3 Dynamic viscosity, kg/(ms) = Pas Surface tension, kg/s2 = Pam

Abrasive Air Relative to air-water (e.g. relative velocity) Relative to air-abrasive Abrasive particle Control volume Focusing tube Sapphire nozzle orifice Isentropic (e.g. in w.jet.s) Water jet Water

Superscripts: * Symbol for non-dimensional ratios. ' Symbol for a property of a scaled model. Simplified time derivative symbol (e.g. m m )

1 INTRODUCTION
Most of today's commercially used abrasive water jet (AWJ) cutting machines operate in a range where the achievable geometrical tolerances typically are greater than 0.1mm, with some companies like Flow and OMAX already offering machining centres claiming to achieve tolerances up to 0.05mm, an achievable accuracy also mentioned by Hashish in [7]. Improving the AWJ cutting process in a way that parts can be manufactured more accurately and in smaller dimensions would enable AWJ manufacturing to be used in an even broader field of applications and to become more competitive compared to other manufacturing methods like laser cutting and electro-discharge machining. To achieve exacter parts, a machine with precise accuracy of motion, a precisely manufactured tool (cutting head) and optimally set machining parameters like water pressure, abrasive mass flow, motion of the tool relative to the workpiece, etc. is needed. Determining optimal cutting parameters is best achieved by the use of models, which ideally make use of physical relationships between tool and workpiece properties to establish a solution for a favourable cut. However in order to create such models detailed knowledge and understanding of the different processes is required. In this context it is the aim of this work, whose beginnings are presented in this paper, to contribute to the understanding of the output characteristics of the tool in dependence of the tool geometry and input parameters by Laser Doppler, reaction force, high speed imaging and if possible other measurement methods. It is planned to document the results in a non-dimensional way, allowing great flexibility in modelling as the results would be transferable to different cutting head configurations. To simplify the initial measurements and to allow a greater spacial resolution, it was decided to conduct the first set of measurements on cutting head models scale 5:1 and at a later stage make comparative measurements on real scale nozzles. The ultimate goal would be the ability to measure the kinetic energy distribution of all three phases (air, water, abrasive) at the exit of the cutting tool, yet it has to be noted that by using currently available methods a complete understanding is an infeasible aim.
2/15

2 OVERVIEW OF MEASUREMENT METHODS AND PREVIOUSLY CONDUCTED WORK


Himmelreich [1, 2] successfully used Laser-2-Focus (L-2-F) velocimetry to measure the velocity distribution in the cross section of both 5:1 and 1:1 abrasive water jets for different cutting head geometries and input conditions. Similarity laws and important non-dimensional numbers (Re, We, ...) were introduced and the similarity between the 5:1 model and real cutting heads discussed. Possibilities to distinguish between velocity signals from abrasive particles and from the water phase were examined. Different intensities were not distinguishable from the photodetectors, so differences in time of flight were used as a basis for distinguishing velocities of the different phases, a method which however didn't allow to explicitly identify an individual signal as originating from an abrasive particle. Lightsheet photography was also carried out, however with limited results due to lack of intensity and long shutter times. Other researchers using pointwise measurement methods included Chen and Geskin [3] who used a Laser Transit Anemometer (similar to L-2-F) and Neusen et al. [4] who used a Laser Doppler Velocimeter (LDV, also called Laser Doppler Anemometer LDA). Both Chen and Neusen were only able to measure an average velocity in the centre of the jet. Due to the many optical disturbances occurring in the abrasive water jet structure after the mixing process of water with abrasive and air, photographic methods appear to be limited to analysing the surface and outline of the abrasive water jet without being able to depict the structures in the inside of the jet. Nevertheless Sawamura et al. [5] used particle image velocimetry (PIV) to measure the velocity of the water phase and particle tracking velocity (PTV) to measure the abrasive velocity. Their results however appear to be limited by high uncertainties. Claude et al. [6] used high speed photography to measure the propagation velocity of the jet when it's turned on and to analyse the contours of a developed jet. Chen and Geskin [3] made Schlieren images of the shock waves occurring when the surrounding air is accelerated to supersonic velocities. While photographing the jet structure after the mixing process is strongly limited, imaging the mixing and acceleration process of abrasive, water and air inside the mixing chamber allows interesting insight into the interaction of the three phases. Hashish [7] observed the motion of the abrasive inside the mixing chamber and plugging of the abrasive flow when large particles are entrained. Osman et al. [8] made similar measurements inside transparent abrasive cutting heads. Methods based on electrical properties were used by Baar and Riess [9], who used a conductivecorrelative method, and by Hashish [7], Swanson et al. [10], Miller and Archibald [11], Dorle, Tyler and Summers [12] who all used coils and an abrasive substitution that would induce a signal when passing through the coils, thus allowing velocity determination by dividing the coil spacing by the time a particle had to get from one coil to another. Of particular interest is the observation of Summers [12] who explained the temporal characteristics of a signal with the rotation of the particles. As particles appeared to rotate between 1'000 to 5'000'000 rpm this would imply a substantial amount of kinetic energy being stored in the rotation of the abrasive particles. Unfortunately no further research into the rotation of particles is known to the authors of this paper, and it is not clear whether the oscillation can definitely be attributed to the particle's rotation and not to some other oscillation excited in the electric circuit. Other possibilities to determine the cutting tool characteristics include force measurements of which examples can be found in Claude et al. [6] who measured the jet reaction force and Momber A. [13] who measured the impact force on the workpiece. Impact count methods like the ones used by Isobe et al. [14] and Liu et al. [15] allow further insight into the cutting tool properties. The throw distance of the abrasive can also be used as an indirect measure for the kinetic energy of the abrasive as shown by Summers et al. [16]. Finally the use of X-Rays as a method to overcome the limits of an optically dense spray for visible light was used by Neusen et al. [17] to determine the distribution of abrasive across the AWJ.

3 THE 5:1 MODEL


Fig. 1 gives a schematic overview of the 5:1 test stand and the measured input variables, which are temperature, pressure and mass flow of water and air; the sand mass flow, air pressure in the mixing chamber inlet and the jet reaction force. The test stand consists of an abrasive feeder unit with a screw conveyor, a collimation tube, a variable cutting head consisting of a mixing chamber and focusing tube and a Danfoss plunger pump capable of reaching pressures up to 15 MPa. For reaction force measurements the usually firmly fixed cutting head and collimation tube hang loosely on a rope only guided by two PTFE bearings.
3/15

Rope

Wa te r

Abra s ive Fe e de r T f

E+H

Pump

Sa nd

p Collima tion tube m

Air
p T

E+H

Nozzle Mixingcha mbe r


Foc using tube

Le ge nd: m ma s s flow T te mpe ra ture p pre s s ure f fre que ncy (m) F force

Fig. 1 Test stand scheme A detailed sketch of one of the cutting heads used is shown in Fig. 2. Most geometrical dimensions are variable, with only the specially manufactured sapphire nozzle kept at a constant nozzle diameter do of 0.75mm. The focusing tubes have diameters dfoc=1.5, 2.3, 3 and 4.3mm with a maximum length of 350mm (200 for dfoc=1.5). The focusing tube can be tilted to correct slight misalignment errors of nozzle and focusing tube.

Fig. 2 Cutting head The model is usually run at a pressure of (14.30.2) MPa, which yields a water mass flow of (2.430.03) kg/min and an isentropic jet velocity (Chapter 4.3) of (1691) m/s.

4 SIMILARITY AND CUTTING HEAD SYSTEM DEFINITION


4.1 Introduction A very thorough overview about the important dimensionless numbers describing an abrasive water jet and a comparison of actually measured geometrically similar 1:1 (called prototype) and 5:1 (called model) cutting heads is given in Himmelreichs PhD work [1]. Nevertheless an overview of the most important parameters is given at this place. The basic idea behind the laws of similarity is that similar models will have similar dimensionless geometric parameters, so the relationship between dependent and independent parameters would only have to be determined once and then could be applied regardless of scale as a perfectly similar system would have similar dimensionless parameters. In real world applications however, it usually is not possible to fulfil the criteria of complete similarity so some sort of compromise has to be achieved. 4.2 System definition and overview of the relevant dimensionless parameters Fig. 3 shows a simplified overview of the system for which dimensionless numbers will be introduced.
4/15

The boundaries of the problem are the outlet of the jet, the exit of the focussing tube and the abrasive inlet including the abrasive conveyor tube and inlet, as the tube diameter and length affects the mass flow and velocity of air, which both are crucial for the transportation of abrasive.
nozzle

do

air + abrasive inlet mixing chamber focusing tube Cutting Head Control Volume
df

Fig. 3 Sketch of the system To describe the problem in a non dimensional way three independent variables covering the fundamental units length (m), mass (kg) and time (s) are chosen from the driving force of the system in Fig. 3, which is the water flowing into the control volume: The initial velocity of the waterjet vw.jet, which is assumed to be identical to the isentropic jet velocity vw.jet.s described in subchapter 4.3. The density of the water in the mixing chamber water. The diameter of the waterjet dw.jet as described in equation 1:

d w.jet d o CD (1)
with CD being the discharge coefficient defined as
CD m water do
2

Lfocus

Lmix

dmix

d2 w.jet do
2

dIn

v w.jet

v w.jet.s

v w.jet.s

Strictly spoken all dimensionless numbers would now have to be made with the three parameters vw.jet, water and dw.jet. However, using parameters only occurring in the subsystem makes more sense in cases where the influence of changing a subsystem (like the focussing tube for example) is investigated. 4.2.1 Reynolds-similarity The Reynolds-number, which describes the ratio of the inertia to the viscosity forces, is chosen as the most important number because it affects the drag and the turbulent mixing of the different phases. Operating the model at a higher Reynolds-number would have the consequence of increased dissipation and higher loss of flow momentum compared to the 1:1 prototype. d w.jet v w.jet water d AP v AP v air-water AP Re w.jet (2) Re AP (3)
water air-water

Equation (2) shows the Reynolds-number of the waterjet as it enters the control volume. It can be treated as an independent parameter, as it mainly is a function of parameters outside the control volume. For Reynolds-similar models similar Reynolds-numbers would have to be achieved, which for the case of equation (2) implies:
Re w.jet Re ' w.jet d w.jet v w.jet
water water

d ' w.jet v ' w.jet ' water ' water


v w.jet . SF

d w.jet SF

v w.jet SF ' water

' water

For the case of identical fluid properties and the term d w.jet v w.jet remains constant for similar models. Out of d ' w.jet d w.jet SF follows v ' w.jet

5/15

Equation (3), the Reynolds-number of the abrasive particles, is directly correlated with the drag acting on the abrasive. 4.2.2 Weber-similarity The Weber-number (equation (4)) denotes the ratio of the inertia forces to the surface tension and is of particular importance for the breakup of the water jet and the formation of water droplets. More water droplets tend to increase the mass flow of air and the exchange of momentum with the abrasive, but also tend to dissipate their kinetic energy much more than a compact water jet due to the increased surface friction and the energy used for surface formation. Hence a jet which atomizes too early won't have enough energy left for cutting.
We w.jet d w.jet v 2 w.jet
water water

(4)
2

The condition for a similar Weber-number would be as follows:


d w.jet v 2 w.jet
water

We w.jet We ' w.jet

water

d ' w.jet v ' 2 ' water w.jet ' water


v w.jet

d w.jet SF

v w.jet SF ' water

' water

, which is not possible for ReynoldsSF similar conditions except if the fluid properties and could be changed. An enlarged model operated at Reynolds-similar conditions will thus have a lower Weber-number than the original. Hence one can expect the droplets to be larger than implied by the scale factor as there is less kinetic energy left for atomization. This in turn alters the exchange of momentum. We make the simplifying assumption that the deviation of flow conditions between the enlarged model and the original is neglectable if both model and original are operated in the same atomization regime, with the same disintegration mode, as shown in Fig. g4 with an x for the original and o for 5:1 model.
10

The condition d ' w.jet d w.jet SF would imply v ' w.jet

Ohnesorge-number

0.1
Raleigh Mechanism

Secondary Atomization

III II

0.01

I
3 4 5 100 1 .10 1 .10 1 .10 Reynolds-number of the Liquid

1 .10

10

1 .10

Border between I (Raleigh) and II (Air influence) Border between II and III Border to complete atomization Real 1:1 nozzle 5:1 Nozzle

Fig. 4 Modes of disintegration. More information about the data in Fig. 4 can be found in [18]. It has to be noted that Fig. 4 strictly isn't valid for the type of supercavitated nozzles used. Eventually validation and comparison measurements on original 1:1 cutting heads will be unavoidable. 4.2.3 Froude-similarity The ability of the airflow to "drag along" abrasive particles to the inlet of the mixing chamber against the force of gravity can be described by the Froude-number, which is the ratio of inertia to gravity forces. For similar abrasive transportation capability the drag coefficient would have to remain similar as in the following equation:
C drag
air AP

d3 AP

which in a simplified form results in the Froude number: 1 2 2 v air-water d AP 2 4

6/15

Fr

v2 air-water (5) - Froude number g d AP

Again Reynolds- and Froude-similarity cannot be achieved at the same time. For this reason, the 5:1 model used is built in a way that the transport of abrasive to the mixing chamber makes increased use of the gravitational potential energy to compensate for the reduced capability of the air to drag along the abrasive against the force of gravity. 4.2.4 Other dimensionless numbers
m abr m air * * and m air m as they influence the Of further importance are the mass flow ratios m abr m water water exchange of momentum between water and abrasive. The characteristic density number * used by Himmelreich [1] is useful as a measure for similar momentum exchange in situations where the abrasive densities between model and prototype are different. * is the ratio of the abrasive density to the average fluid density in the focusing tube and is calculated as follows: * AP water V water air V air with air-water V air V water air-water

Further dimensionless numbers like the Mach- and Galilei-number are currently considered less important (though not necessarily absolutely neglectable) and are not further discussed in this paper. 4.2.5 Varied parameters Among the most important parameters to be varied are:

L foc * the dimensionless length of the focussing tube L foc d foc

d foc d AP * * the diameter ratios d foc d and d AP d ,which to reduce the number of measurements can w.jet w.jet d AP * with denominator of d* being a measure for the gap between be combined to d 0.5 d foc d w.jet the focussing tube wall and the water jet. m abr * the mass flow ratio m abr m w.jet

and geometric parameters of the mixing chamber which in turn affect the dimensionless mass flow m air * of air m air m w.jet

4.3 Isentropic velocity The conditions to be able to use the Bernoulli-Equation (equation 6) to calculate the velocity of the water jet are: Steady state conditions, incompressible flow, no friction, neglectable gravity forces and flow along streamlines, no heat transfer.
v w.jet.Bernoulli
2

p
water

(6) - Bernoulli equation

For high pressure water-jets the condition of incompressibility is not fulfilled as at 400MPa the water density increases by 13% compared to water at ambient pressure. A thermodynamical approach to solve for the ideal expansion velocity of the water is to use the energy balance (equation 7) set over a control volume as in Fig. 5.
p1, T1, v1 Control Volume p2

Fig. 5 Control volume over a simplified nozzle

7/15

dE CV W CV Q CV m dt

h1

1 2 v g z1 2 1

h2

1 2 v g z2 2 2

(7) - Energy balance

Assuming steady state conditions, adiabatic walls, constant entropy, isentropic expansion, neglecting the potential and kinetic energy at the inlet simplifies equation (7) to equation (8), the isentropic velocity. (8) - Isentropic velocity The enthalpies can be determined from a database of thermodynamic properties of water. The isentropic temperature T2.s can be determined from the condition of constant entropy s1=s2. Details about the equations applied above can be found in thermodynamics literature like [19]. Comparing the calculated isentropic velocities for an initial temperature of T1=25C with an equation for the waterjet velocity given by Hashish in [7] yields excellent agreement (Fig. 6).
1000

v 2. s p 1 ,T 1 , p 2

2 h1 p 1 ,T 1

h2 p 2 ,T 2.s

800 Velocity m/s

600

400

200

100

200 Pressure MPa

300

400

Jet velocity according to M. Hashish Isentropic velocity Bernoulli velocity

Fig. 6 Jet velocity comparison

5 LDA / PDA MEASUREMENTS


5.1 Setup and principle A detailed overview of the principle underlying LDA / PDA (also called LDV / PDPA) measurements is given in [20]. Basically the method utilises two coherent laser beams which cross in the focal point of a lens and there define a control volume of which the frequency of light from the two laser beams reflected from a particle passing through the control volume is directly correlated with the velocity of the particle. For "perfectly" spherical particles (i.e. small water droplets) the phase shift of the reflected light measured from two different angles contains information about the size of the particles. The currently used system utilises a 2D Dantec Dynamics Fibre PDA and BSA P80 system with 60mm transmitter and receiver probes and a Coherent Argon-Ion laser. The receiver probe is aligned in a scattering angle of 70 to the transmitter probe and the size of the control volume is (0.25 x 0.092 x 0.092) mm. Two velocities, one along the axis of the jet and one perpendicular to both the jet axis and the optical axis of the transmitter are measured simultaneously. 5.2 Limitations / Difficulties In order to get good, clearly defined signals, the particles reflecting the light would have to be smaller than the size of the control volume. While most water droplets are small enough to fulfil this criteria, the 5:1 quartz sand particles used in a size range of 0.5-1mm are certainly not. Test velocity measurements on sand particles poured through a funnel have shown that the LDA is actually able to measure the sand velocity although the resulting signal was rather noisy. It also has to be kept in mind, that the measured velocity for large particles will correspond to their surface velocity, which is influenced by the rotation of the particles. A central limitation and difficulty is the ability to identify an incoming signal as originating from an abrasive particle or from the water phase. Chen [3] argued that the intensity of the signal could be used as a criteria to distinguish a signal as originating from an abrasive particle, Himmelreich [1] however mentioned that the sensitivity of the photo detectors he used was insufficient for such a distinction. With the current LDA system used for the 5:1 measurements presented in this paper only an average intensity (photo multiplier anode current) over time can be measured, thus intensity again cannot be used to identify signals from an abrasive particle. An alternative possibility is to compare the
8/15

velocities of the validated water droplets with the velocities of all signals (validated and non validated) as shown in Fig. 7. (N.B. a non validated signal doesn't imply that it's not from the water phase as even when the abrasive is turned off the droplet validation rate is seldom higher than 30-60%).
1600 1200 "Counts" 800 400 40 60 80 100 Velocity m/s 120 140

All counts Diameter Validated Counts (scaled up)

Fig. 7 Velocity histogram 5:1 AWJ m*abr=0.33 For the test conditions no apparent difference in velocity distribution was recognisable. This could be explained with the long focusing tube used (L*foc=82), which could be the reason for a strong mixing of the different phases. Himmelreich [1] measured the time of flight distribution for different focussing tube lengths and observed two distinct maximas in time of flight for shorter focussing tubes, which gradually merged into one distribution for longer focussing tubes as shown above. Another criteria for the origin of a velocity signal could be its transit time (TT), which is a measure for the duration of the signal and for the time a particle had to cross the control volume. At a given velocity only a very large particle (compared to the control volume size) would have a longer transit time than a smaller particle. However, making use of this as a criteria for identification of the abrasive particles failed, as high transit times were typically observed at high velocities as shown in Fig. 8 for measurements with and without abrasive and even for the velocities of the validated diameters.
0.3

Normalised Counts

0.15

20

40

60

80 100 Velocity m/s

120

140

160

All TT 1%< < 5% of all TT counts (i.e. lower TT) 99%> >95% of all TT counts (i.e. higher TT)

Fig. 8 Velocity distributions for different transit times In all cases where the signal origin cannot be clearly distinguished and other indicators like the ones described above have to be used, a major uncertainty is the bias caused by an unequal number distribution of signals originating from the abrasive and water particles. The volume of a 0.5mm sand particle would be equal to that of more than 15'000 20m water droplets, which without considering all other effects would imply a higher data rate of velocity signals from water than from sand. Quantifying the probability and frequency (rate) of detection for the water and abrasive phase is generally a very difficult task, as there is a plethora of effects which would have to be considered (like for example that a sand particle can only be detected as long as it doesn't completely cover or shade out the control volume or that the measurement volume appears larger for a large sand particle than for a small water droplet etc.). 5.3 LDA velocity measurement results Fig. 9 shows the measured velocity distribution at the exit of the focussing tube with and without abrasive. It can be seen that the distribution resembles the velocity distribution of a pipe flow, an observation also made by Himmelreich [1] for longer focussing tubes. The error bars show the standard deviation of the velocity histograms at the respective points. The third dotted curve is the
9/15

result of scaling down the water velocity distribution by a factor obtained by the law of exchange of momentum (equation 9), with the difference between the dotted curve and the velocity with abrasive being a measure for the amount of dissipation. It has to be kept in mind however, that a bias due to different measurement data rates would result in a deviation from the measured velocity distribution to the actual velocity distribution of the abrasive particles. m water.1 mair.1 v dotted v water.1 m abr.2 m water.2 m air.2 (9) reduced velocity via the law of momentum exchange
1

Dimensionless Axial Velocity v/vs

0.75

0.5

0.25

1.5

0.5 0 Dimensionless Radius in y r/rfoc

0.5

1.5

With abrasive Standard deviation limits Just water (without abrasive) Standard deviation limits Estimated velocity by law of momentum exchange

orifice boundary d0

focussing tube boundary dfoc

Fig. 9 Axial velocity distribution at the exit of the focussing tube The average radial velocity distribution in the cross section of the focussing tube is in the range of the possible misalignment error (i.e. close to zero). More interesting are the fluctuations, which in the centre region of the focussing tube are in the range of 4% of the axial velocities and in the border region reach up to 10% of the axial velocity (Fig. 11). Another effect is a strong reduction in the LDA data rate (number of measured valid velocity signals per time) by more than an order of magnitude. This is caused by the increased disturbances when adding abrasive to the waterjet, an effect that can also be seen in the noisy signal and the low data validation rate.
2.5 .10 Data rate Ch1 #/s 2 .10
4 4 4 4

1.5 .10

1 .10

5000 0 2 1.5 1 0.5 0 0.5 Dimensionless Radius in y r/rfoc 1 1.5

With abrasive Just water (without abrasive)

Fig. 10 LDA measured data rate


Ratio of radial to axial velocity % 20

15

10

1.5

0.5 0 0.5 Dimensionless Radius in y r/rfoc

1.5

10/15

With abrasive Just water (without abrasive)

Fig. 11 Ratio of the radial to the axial velocity in the focussing tube

6 IMAGING
6.1 Setup and principle In order to get an idea of the structure of the 5:1 jet and the applicability of imaging methods, the jet was photographed with an SCO 4QuikE intensified fast shutter (ICCD) camera used with shutter times between 0.1s and 0.5s. Two different light sources in form of a halogen lamp, used for back light illuminated pictures, and a laser light sheet, fed by a continuous Argon-Ion Laser in multi line mode, (Fig. 12) were used.
Halogen Lamp 4 Quik E Laser Light Sheet

Fig. 12 ICCD Camera Setup 6.2 Results and Discussion Fig. 13 shows the picture of the waterjet without abrasive as it exits the focussing tube. It is apparent, that the disintegration of the jet is at an advanced stage, a fact that is contributed to by the rather long focusing tube (length 350mm, nozzle distance ~400mm). For this reason photographs of the jet without the mixing chamber and focusing tube installed were taken (Fig. 17-19), and it was determined that in free air the achievable nozzle distance for a more or less coherent jet was approximately 300mm. Adding abrasives to the flow (m*abr = 0.33) as in Fig. 14 resulted in a much denser jet with many small droplets and a foggy curtain making it impossible to locate the abrasive particles. These pictures show, that the AWJ really is a three phase flow with a high degree of mixing between the three phases. Illuminating the jet with a laser light sheet as in Fig. 15 & 16 showed discrete intense spots on the picture, which probably could be used to determine velocities with the particle tracking method. With the abrasive turned on, there was a strong source of reflection directly at the side of the waterjet where the laser light sheet entered the jet. Again the jet / spray was to dense to spot any abrasives. Images showing the jet. The markers indicate the focussing tube walls and jet orifice diameter do.

Fig. 13 No abrasive, halogen backlight

Fig. 14 W. abrasive, halogen backlight

11/15

Fig. 15 No abr. - light sheet (negative image)

Fig. 16 W. abr. - light sheet (negative image)

Fig. 17 50mm nozzle distance

Fig. 18 250mm nozzle distance

Fig. 19 400mm nozzle distance

7 REACTION FORCE MEASUREMENT


The advantage of reaction force measurements is, that the direction of the flow in and out of the control volume over which the condition of static equilibrium has to be fulfilled is much more known and well definable than for impact force measurements.
F
mwater.in, vwater.in

Control Volume
mabr, vabr.in

mout, vout

min, vin mout, vout Control Volume Freaction

Fig. 21 Impact force


mout, vout

Fig. 20 Reaction force Applying Newton's second axiom to a control volume yields: d v d F v dV v v dA dV dt dt or with the assumptions and simplifications of steady state flow conditions, averaged velocities, frictionless bearings, excluded gravity forces and with from a point of view where the control volume is a non-accelerated inertial system, yields equation 10, which can be found in fluid mechanics literature like [21].

v CV.in m CV.in (10) - balance of momentum Applying equation (10) to the system in Fig. 21 to determine the jet reaction force causes great difficulties, as the flow out of the control volume is not well defined and changes with the cut geometry, so there are too many unknowns for the equation to be solved. Equation (10) applied to the system in Fig. 20 in axial direction however allows direct determination of a momentum averaged jet velocity:

F CV

v CV.out mCV.out

12/15

v w.jet.momentum

F axial m water m abr mair (11) - momentum averaged velocity

At the time of writing this paper the force measurement system of the 5:1 model still had some difficulties, mainly due to a drift which was probably due to a connection pipe. Nevertheless, the effect of turning the abrasive on and off can be seen in Fig. 22&23, where the reaction force decreases when the abrasive is turned on (indicated with an arrow). The force shown is the force gravity superimposed to the reaction force, which acts in the other direction.
171.6

171 Force / N
6 6.5 7 7.5 time / min 8 8.5

Force / N

171.4 171.2 171 170.8

170.7 170.4

170.1 8.3

9.2 10.1 time / min

11

Fig. 22 Force w/wo abr, m*abr=0.33

Force

Force

Fig. 23Force w/wo abr, m*abr=0.16

8 OTHER METHODS
Non of the applied methods allowed the unambiguous measurement of the abrasive particle velocities, and of all the known methods only the ones measuring current induced by particles moving though coils appear to work reliably. An optical method to measure the velocity of the abrasive particles would be to use fluorescent particles with a sort of LIF (laser induced fluorescence) method. After excitation by a high power laser light the particles emit light at a higher wavelength which can be seperated from the laser light by means of an optical filter, allowing the application of a particle tracking algorithm. Using fluorescent particles was already suggested by Himmelreich in [1] but he couldn't pursuit utilising the method as for the high price of such particles. The authors intend to commence work on utilising either UV excitable sand, which occurs naturally, or fluorescent coloured sand particles for AWJ measurements in the near future. A method which could lead to very interesting results is the use of flash x-ray photography, which as far as the authors know wasn't applied for abrasive waterjets yet, but could give some very interesting results. An overview of the technique can be found under [22].

9 CONCLUSIONS
From this paper, which has given an overview of different measurement methods applied to a 5:1 scale model of an abrasive waterjet, the following conclusions can be drawn: Similarity between 5:1 and 1:1 scale AWJ models is never complete. For Reynolds-similar flow the Weber- and Froude-similarity, which both are critical for the AWJ process, cannot be fulfilled. The isentropic jet velocity, which can be determined by solving the energy balance over the nozzle, shows excellent agreement with an equation presented by Hashish [7]. Using the LDA / PDA measurement methods, allows measurement of the 5:1 AWJ velocity at high spacial resolution, however the difference between signals from the abrasive- and water particles cannot yet be distinguished. High speed imaging methods allow the characterisation of the AWJ to some degree, however the AWJ is too dense to make out any abrasive in the image. It can be seen, that the AWJ flow of the test conditions is a three phase flow with a strong degree of mixing between the three phases. Reaction force measurements allow the estimation of an average AWJ velocity for comparison with other measurements. It was also shown, that doing the same with impact force measurements is very difficult and dependant of the cut geometry. Promising methods that could be applied in the future are the LIF / PTV method utilising fluorescent abrasive particles and perhaps flash x-ray photography.

ACKNOWLEDGEMENTS
The authors would like to thank the CTI Swiss Innovation Promotion Agency, WaterJet Holding AG Aarwangen, Switzerland and MVT AG Nidau, Switzerland for support and funding.
13/15

REFERENCES
[1] [2] [3] Himmelreich U.: Fluiddynamische Modelluntersuchungen an Wasserabrasivstrahlen. VDIVerlag, Dsseldorf, 1993. (In German). Himmelreich U., Riess W: Hydrodynamic Investigations on Abrasive-Waterjet Cutting Tools. Proceedings of the 10th International Conference on Jet Cutting Technology, 1991, pp. 3-22. Chen W.L., Geskin E.S.: Measurement of the Velocity of Abrasive Waterjet by the use of Laser Transit Anemometer. Proceedings of the 10th International Conference on Jet Cutting Technology, 1991, pp. 23-36. Neusen K.F., Gores T.J., Amano R.S.: Axial Variation of Particle and Drop Velocities Downstream from an Abrasive Water Jet Mixing Tube. Proceedings from the 12th International Conference on Jet Cutting Technology, 1994, pp. 93-103. Sawamura T., Fukunishi Y., Kobayashi R.: Study of the abrasive waterjet structure by measuring water and abrasive velocities separately. Proceedings from the 14th International Conference on Jet Cutting Technology, 1998, pp. 185-193. Claude X., Merlen A., Thery B., Gatti O.: Abrasive waterjet velocity measurements. Proceedings from the 14th International Conference on Jet Cutting Technology, 1998, pp. 235251. Hashish M., Abrasive-waterjet (AWJ) studies, Proceedings from the 16th International Conference on Water Jetting, 2002, pp. 13-48. Osman A.H., Busine D., Thery B., Houssaye G.: Visual Information of the Mixing Process Inside the AWJ Cutting Head. Proceedings from the 9th American Waterjet Conference, 1997, pp. 189-209. Baar R., Riess W.: Two Phase Flow Velocimetry Measurements by Conductive-Correlative Method. Journal of Flow Measurement and Instrumentation, Vol. 8. No. 1, pp 1-6, 1997.

[4]

[5]

[6]

[7] [8]

[9]

[10] Swanson R.K., Kilman M., Cerwin S., Tarver W.: Study of Particle Velocities in Water Driven Abrasive Jet Cutting. Proceedings from the 4th American Waterjet Conference, 1987, pp. 163-171. [11] Miller A.L., Archibald J.H.: Measurement of Particle Velocities in an Abrasive Jet Cutting System. Proceedings from the 6th American Water Jet Conference, 1991, pp. 291-304. [12] Dorle A., Tyler L.J, Summers D.A.: Measurement of Particle Velocities in High Speed Waterjet Technology. Proceedings from the 2003 WJTA American Waterjet Conference, Paper 2-G. [13] Momber A. W.: Energy transfer during the mixing of air and solid particles into a high-speed waterjet: an impact force study. Journal of Experimental Thermal and Fluid Science 25, 2001, pp. 31-41. [14] Isobe T., Yoshida H., Nishi K.: Distribution of Abrasive Particles in Abrasive Water Jet and Acceleration Mechanism. Proceedings of the 9th International Symposium on Jet Cutting Technology, 1988, pp. 217-238. [15] Liu H.-T., Miles P.J., Cooksey N., Hibbard C.: Measurements of Water-Droplet and Abrasive Speeds in a Ultrahigh-Pressure Abrasive-Waterjets. Proceedings of the 10th American Waterjet Conference, 1999, Paper 14. [16] Summers D.A., Fossey R.D., Newkirk J.W., Galecki G.: Results of Comparative Nozzle Testing Using Abrasive Waterjet Cutting. Proceedings of the 2001 WJTA American Waterjet Conference, 2001, Paper 15. [17] Neusen K.F., Alberts D.G., Gores T.J., Labus T.J.: Distribution of Mass in a Three-Phase Abrasive Waterjet Using Scanning X-Ray Densitometry. Proceedings of the 10th International Conference on Jet Cutting Technology, 1991, pp. 83-98. [18] Lefebvre A.H.: Atomization and Sprays. Taylor and Francis, 1989, pp. 37-45. [19] Michael J. Moran, Howard N. Shapiro: Fundamentals of Engineering Thermodynamics. John Wiley & Sons, Inc., New York, 1996.
14/15

[20] Albrecht H.-E., Damaschke N., Borys M., Tropea C.: Laser Doppler and Phase Doppler Measurement Techniques. Springer, Berlin, 2002. [21] Fox R.W., McDonald A.T.: Introduction to Fluid Mechanics. John Wiley, 1994. [22] Observation of high-speed processes by means of x-ray photography. http://www.emi.fraunhofer.de/english/Departments/ExperimentalBallistics/DeptPages/Projects/ Observation.html

15/15

You might also like