You are on page 1of 11

Jit Breakup and Mixing Throat Lengt

R. G. CUNNlNGHAi
tie Liquid Jit Gas P u p
Professor of Mechanical Engineering,
The Pennsylvania State University, Gas compression with a liquid jet occurs isothermally and hence with minimum work.
University Park, Pa. Performance characteristics oj the liquid jet gas pump (efficiency arid compression
ratio versus inlet volumetric flow ratio) are predicted accurately by a one-dimensional
R. J. DOPKIN analysis providing (he mixing zone remains in the throat. Jet breakup was investigated
to enable prediction of required throat length and to improve efficiency. Effects of
Engineer throat length, nozzle contour and spacing, nozzle-throat area ratio (0.15 to 0.4-5), jet
E. I. DuPont de Nemours and Co.
Textile Fibers Dept. velocity and suction pressure were investigated. Optimum throat lengths were found;
Spruance Plant corresponding efficiencies exceed 40 percent. Two jet breakup flow regimes were
Richmond, Va.
found: impact and jet disintegration. For the impact regime, jet breakup length
depends on inlet velocity ratio, jet Reynolds number and nozzle-to-throat area ratio.
Optimum throat lengths were found to be an empirical function of nozzle-to-throat
area ratio and ranged from 12 to 82 throat dia. These results, coupled with the one-
dimensional model, permit design of efficient liquid jet gas pumps.

Introduction jet breakup characteristics (which are in turn affected by the


design of the nozzle). The objectives of this investigation were
The liquid jet gas (LJG) pump is a device in which a liquid jet the systematic study of the liquid jet breakup in a mixing tube
entrains and compresses a gas in a cylindrical mixing throat. with a cocurrent gas flow, and the development of throat design
The resulting homogeneous mixture of gas bubbles in liquid is
criteria for LJG pumps.
discharged through a diffuser, to increase the static pressure at
the expense of the kinetic energy of the mixture. The Gas Compression Processes. It is common observation
As shown in Fig. 1, the LJG pump resembles the well known that given sufficient initial velocity, a free liquid jet discharged
liquid-liquid jet pump but with one important distinction: mix- into a gas will ultimately break up, i.e., expand as droplets are
ing throat lengths considerably in excess of the four to eight dia shed from the potential core. The breakup length is the distance
characteristic of liquid-liquid jet pumps are required for ef- from the nozzle tip to the disappearance of the potential core.
ficient LJG pump performance. Properly designed LJG pumps Numerous investigations have confirmed that free jet breakup
are remarkably efficient devices. Isothermal compression ef- starts with the appearance of ripples or waves within a few diam-
ficiencies in excess of forty percent have been reported [1, 11, 15]1 eters of the nozzle tip which then grow in magnitude until drop-
which is quite comparable to the best efficiencies achieved in lets or particles of fluid are shed into the surrounding atmosphere;
liquid-liquid or gas-gas jet pumps. These high efficiencies rep- the shedding process continues until the potential core is con-
resent a significant advance in the design of LJG pumps, which sumed.
heretofore have been characterized as extremely inefficient, i.e., Referring to Fig. 1, the liquid jet and the gas annulus enter
with efficiencies of ten percent or less. the throat at o at velocities Vl0 and Vio, respectively. In the
Although the drawback of low efficiency has been overcome, annular entrance zone the two streams remain essentially un-
earlier work at Penn State [2, 3] has shown the dependence of changed although of course a gas boundary layer begins to de-
LJG pump behavior on mixing throat length. Performance is velop on the wall and interfacial shear forces act between the
theoretically predictable only as long as the mixing is completed surface of the jet and the gas annulus. Within the throat but
in the throat, and so the ability to predict mixing zone behavior prior to the mixing zone the wave disturbances on the jet reach
is very important. Design of the LJG pump to provide maximum critical magnitude and high velocity droplets are shed from the jet
efficiency will require matching the correct throat length to the into the annular space. The potential core diminishes and ul-
timately disappears as the flow proceeds to the right. Down-
stream of the mixing zone, liquid is the continuous medium—in
'Numbers in brackets designate References at end of paper. contrast to annular flow upstream of the mixing zone where the
Contributed by the Fluids Engineering Division of The American Society gas is the continuous medium—and the liquid contains finely
of Mechanical Engineers and presented at the Joint Fluids Engineering & dispersed gas bubbles.
CSME Conference, Montreal, Quebec, Canada, May 13-15, 1974. Manuscript
received at ASME Headquarters February 14, 1974. Paper No. 74-FE-17. A sudden expansion in a pipeline is a diffuser (albeit relatively

Downloaded From:216 /SEPTEMBER 1974


https://fluidsengineering.asmedigitalcollection.asme.org Copyright
on © 1974
06/30/2019 Terms of Use:by ASME Transactions
http://www.asme.org/about-asme/terms-of-use of the ASME
inefficient) in which a pressure rise results from a rapid decelera-
tion of the fluid. The throat of the liquid jet gas pump cor- secondary gas
responds to the sudden expansion fitting, when the gas flow is
restricted to zero. When gas is admitted however an interesting Bubbly muture
and .useful phenomenon occurs. The pressure rise in the throat Oiffuser
is gradually reduced as increasing amounts of gas are admitted.
This decrease in pressure rise is a result of energy transfer from
the liquid to the gas in an essentially isothermal compression
process. A one-dimensional flow model including friction loss
coefficients for the nozzle and the throat wall accurately predicts
the throat compression process as a function of the entering
velocity ratio v = Vso/Vm and the jet pump number n [1],
From point I to d in the diffuser the mixture is decelerated to Fig. 1 Liquid-jet gas pump and nomenclature
achieve a rise in static pressure at the expense of the kinetic
energy of the mixture (largely that of the liquid since the mass
of the gas is relatively small). As the mixture pressure rises,
energy is again transferred from the liquid to the gas but this
time the mechanism is somewhat that of a mechanical piston
compressor. The gas and liquid velocities are identical (unlike
standpoint the Penn State work to date has established that the
the throat process prior to mixing) and mechanical work is per-
performance of the LJG pump can be predicted accurately—if the
formed by the liquid on the gas as the diameter of each entrained
mixing is completed in the throat—and that, high efficiency per-
gas bubble diminishes in moving from t to d. Experiments clearly
formance (> 40 percent) is attainable via the use of proper noz-
confirm that the pressure rise achieved in the diffuser is con-
zles and long mixing throats. "Correct" throat lengths however
siderably reduced by this transfer of energy from the liquid. If
have not yet been defined. To meet this need Dopkin [15] re-
this point is overlooked in treating the flow of two-phase fluids
cently investigated single-jet breakup response to pump geom-
in diffusers, the omission of the isothermal work term will result
etry and operating parameters, a study on which this paper is
in abnormally low diffuser efficiencies or high loss coefficients.
largely based.
Previous Investigations. The application of the LJG pump
for evacuation of stream condensers was discussed by Hoefer
[4] in 1922. A number of similar application references is in-
Theoretical Flow Model
cluded in the recent BHRA literature survey of jet pumps [5) It has been shown that by treating the static pressure changes
but in many cases the information provided has been inadequate across the nozzle, throat and diffuser, one-dimensional continuity
to describe pump operation. Contributions of direct interest in and momentum relations can be used to describe accurately the
this study [2, 3, 6, 7, 8, 9, 10, 11, 12, 13, 14] were reviewed in a liquid jet gas (LJG) pump processes [1]. Using the Nomenclature
companion paper [1]. shown in Fig. 1, the theoretical equations for each component
Further improvements have been reported for the one-di- are repeated here for convenience.
mensional flow model for the liquid-jet gas pump, including an Pressure drop at the nozzle:
energy analysis developed to account for all energy lost or ex-
changed during the pumping processes [1]. From a design Pi — Po Z(l + Kn.) (1)

-Nomenclature-
JP= jet Reynolds number rjmt mixing throat efficiency
A = cross-sectional area, ft2
D = diameter, ft
o- _ liquid surface tension lbf/ft
3
Kn!, Ken, MlSc p = density, lbm/ft
friction loss coefficients Vio'A.pi V = velocity ratio at the throat
K,h, Kdi — IjJ— jet Weber number inlet, TWFio = <t><,/c
LJG = liquid jet gas pump 2
L, = throat length, ft a = diffuser area ratio, At/A a 111 = liquid viscosity, lbf sec/ft
b = area ratio, A„M i 4>o, <t>l, 4>d =s: volumetric flow ratios, Qi„/
LT = throat length/dia ratio L</
Di
1 - b Qi, Qj./Qi and Qu/Qi
c — area ratio, Ai0/A„ — —;—
Lm = jet breakup length, ft o Subscripts
p = pressure, lbf/ft2 e = energy rate, ft lbf/sec 1 = liquid primary flow
P = total or stagnation pressure, gc = gravitational constant ft 2 = gas secondary flow
lbf/ft* lbm/lbf sec* 3 two-component mixture of
3
Q = volumetric flow rate, ft /sec in = mass flow rate, lbm/sec gas in liquid as the con-
R = gas constant, ft lbf/lbm °R 2ZWc tinuous medium
S = nozzle-to-throat spacing in
n = jet pump number ——
Po i, n, s, o,
throat diameters, Si/D, rt0, rat, r,,s = pressure ratios Pt/P0, Pa/ I, d = location, see sketch
St = nozzle-to-throat spacing, ft P„ Pa/P, f = force
T = temperature, °R xo = specific work ft lbf/lbm = fric = friction
V = velocity, ft/sec W/m 9 = gas partial pressure
W = work rate ft Ibf/sec y = density ratio, pio/pu where he = kinetic energy
z= Pio = Pto/RTia ideal gas m = mass or mixing length
jet. velocity head,
2<? c relation opt = optimum
lbf/ft2 V>p = ie^ P u m P efficiency V = vapor

Journal of Fluids Engineering SEPTEMBER


Downloaded From: https://fluidsengineering.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use
1974/ 217
therr'(,f~eter
Q., ft3/ min of gal; frum P, to i'd, so that
Woul == Q.,RT, In Pd/P,. (4)
rotameter -+
(Q,o)_ The energy input rate from the liquid stream is ein == Ql(P i - Pd.)
surge Pump efficiency is defined as
tank

W oul P,<p, In Pd/P,


l/jp == (.5)
Pi - Pd
The corresponding throat efficiency is

JET PUMP TEST SECTION Po<Po In PdPo


(6)
Z + Po - P,
By combining energy conservation relations with the con-
orifice meter
(01 )
tinuity-momentum relations employed in derh ing equations
(1), (2) and (3), it has been shown that the throat pressure and
diffuser pressure rises are both diminished by friction losses

rc=
cooling water
thermometer
and the isothermal compression work on the gas by the liquid.
manometer The throat pressure rise is also diminished by the mixing loss
-~~rfloi",'
which occurs whenever two streams of dissimilar velocity are
mixed [1J. This phenomenon is normally the largest source of
energy dissipation and it controls the overall performance of
Fig.2 Test stand and instrumentation the LJG pump. Mixing loss is a function of the inlet velocity
2Zb2c
ratio v == V'o/V lo and the jet pump number n == - - ; and the
Po
Pressure rise in the throat: design of the LJG pump involves selecting the ar£a ratio b to
permit operation at the lowest possible jet pump number n, in
order to minimize mixing losses. In addition to v and n, the dif-
fuser compression ratio rd' and the overall pump compression
ratio rd, are also functions of the nozzle-throat area ratio band
(2) the diffuser area ratio a, i.e., rdl (v, n, b, a) and rd,(v, n, b, a).

Pressure rise 1:1/ the diffuser: Experimental Program


The experimental apparatus was described in references [11
Pd - P, == Z(l + 'Y<Po)[b 2(1 + <p1)2 - a2b2(1 + <PdP and [15] and will be summarized here. A sketch of the te~t
ICli b2(1 + <PI)] Po<Po In Pd/P , (3) stand is shown in Fig. 2. The test pump is shown in the Fig. 3
photograph. The building-block throat sections, several of
The effect of pump geometry is reflected in the nozzle-to- which appear in Fig. 3, permit assembly and testing of a series
throat area ratio b == An/A,. The mass flow ratio is l'<Po == Ill'; of LJG pumps with throat lengths ranging from 10 to 52 dia.
Inl. Z is the jet velocity head. The volumetric flow ratio <P == Q./
The throat sections (0.500 in. LD.) were machined from 3 in. X
QI varies along the pump flow path as the isothermal function of 3 in. cross-section plexiglas blocks with flat sides to permit ob-
the static pressure, so that the throat and diffuser equations are serving and photographing the mixing process. Various nozzle
nonlinear. When <P, == Q2I/QI == <PoPO/P , is eliminated, the re- tips were soldered to lengths of one-in. LD. copper tubing u~ed
sulting throat equation is a quadratic in P" or in the compression as nozzle approach sections; one appears at the left in Fig. 3.
ratio r,o == PdP o. Similarly, the diffuser equation is a cubic in All tests were made with tap water maintained at constant tem-
1"dl == Pd/P ,. The K terms are friction loss coefficients based on perature by adding fresh water, and with atmospheric air serving
the expression
as the gas component. The primary (liquid) flow rate was
V2 measured with an ASME flange-tap orifice meter and the
D.Plric == Kp - secondary (gas) flow with high- and low-range rotometers.
2(Jc
Three pressures (Pi, P, and P d ) must also be measured to deter-
where p and V are the local density and velocity of the flow mine pump efficiency and compression ratio; pressure gauges
mixture. Gas is entrained from pressure P, and isothermally and manometers are shown in Fig. 2. To assess the performances
compressed to the pump discharge pressure Pd. Heat rejection of the throat and diffuser components, the throat discharge
to cool the gas and from frictional dissipation cause a negligible pressure P, was also measured. Pressure profiles describing the
liquid temperature increase. Separate phases enter the throat
but a homogeneous mixture is produced in the mixing zone;
there is no slip between gas and liquid phases after mixing.
Changes in solubility of the gas in the liquid are neglected. The
static pressure P == (P. + P q ) is the sum of the liquid vapor
pressure and the gas partial pressure, and the minimum static
pressure attainable will be P == Pc at zero gas flow.
A computer program is used to calculate throat and pump
pressure ratios, mixing and friction losses, and efficiencies point
by point for increments of the inlet flow ratio <Po. Input values
are pump area ratios a and b, flow variables Z, P" T, and R;
friction levels K n" Ken, K 'h and J(di; and the interval on the
independent variable <Po.
The useful pump output is the isothermal compression of Fig. 3 Liquid-Jet gas pump with variable throat 'enQthl

218 / S E PTE M B E R 1 9 7 4 Transactions of the ASME


Downloaded From: https://fluidsengineering.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use
transparent
throat from the entrance of the LT = 22 diameter throat. At the <j>0'
departure point, the downstream end of the mixing zone was at
the exit of the throat (L,„ = 22). Further reduction of Pa causes
the mixing zone to move into the diffuser, and agreement be-
tween theory and experiment ceases beyond the 4>„' point. This
Sliding 1 in 1.0.
is to be expected since the mixing loss (closely related to the sud-
copper tube.
Pj pressure tap 9.5 in.
den expansion loss in a conduit) is dependent on the jet-to-mixing
upstream from tip. zone area ratio, and allowing mixing to extend into larger cross-
sectional areas in the diffuser incurs increased mixing losses. Per-
formance under these off design conditions may also be degraded
by observed wall separation in the diffuser. The net result is
reflected in the rapid decrease in overall pump efficiency rj,-,
as shown to the right of <f>„' in Fig. 5.
In contrast to the liquid-liquid (LL) jet pump, LJG jet pump
compression ratios (Fig. 5) are relatively flat plotted versus <£<,,
as long as the mixing zone is maintained in the mixing throat
SEl .252 .196)t
(operation to the left of 4>0'), (The LJG and LL pump charac-
.025 151
SE2 .352 .274) .035 300 teristics were compared at similar jet pump numbers n in [1].)
SE3 .429 .336) .035 449
El .274 .100 300 The top curves in Fig. 5 show the measured friction loss coef-
:3« 27 c 0".r "n"
U (0 -- )2-) .275 ,!?•> VT ficients Kik and Ka% versus the flow ratio <f>o- Significantly, the
7H" .104! 7) .0£2 300 friction coefficients remain essentially constant versus <p„, and
" Contracted jet diameters Used on area contraction coefficients of this constancy is cited as good evidence that the theoretical model
.605, .?06 and .612 rescectively, ohtained bv mass flow neas'Jrenents.
* Has 5° divergent section, lenqth .030 in, followinn X. satisfactorily describes the friction losses including the two-phase
**Seven hole no2zle with six .104 inch diameter holes on a circle of .312
diameter rlus center hole. Inlets rounded.
flow effects. Measured K values are sensitive indices to pump

Fig. 4 Pump Inlet configuration (top), and no»la details

1.0
mixing process were measured with a series of static pressure
taps in the throat sections.
The influence of nozzle design was investigated and five of the
nozzle contours are described in Fig. 4. The key design param-
eter is the nozzle-to-throat area ratio 6 = A„/At] in the case of
jets formed by the square-edge orifices, the cross-sectional area
A„ is taken to be that of the vena contracta. The contraction
coefficients, measured by mass flow tests, and orifice diameters
are listed in Fig. 4 for nozzles sized to provide three area ratios:
b = 0.153, 0.300 and 0.449.
Nozzle-to-throat spacing was <S = 3 throat dia. Performance
appeared to be relatively insensitive to spacing (from S = 1 to 7
throat dia) when corrected for the S effect on total jet length
from nozzle tip to break up of the jet [15].

Theory Experiment Comparison


Performance tests for a given 6-ratio pump were made with
fixed jet velocity Vi0 and suction port pressure P„ and hence
with a fixed jet pump number n - —— , where Z is the jet
velocity head and

An
Shown in Fig. 5 are results for a b = 0.300 pump in terms of
efficiencies and compression ratios for the pump, and for the
mixing throat alone; the separation of the curves is of course a (
result of the diffuser compression process. Note that the experi- /
Theory; D
mental points agree well with the theoretical predictions up to
the "departure point" flow ratio which is labelled $„' on the Tjj> Z = 96.3 psi
D
curve. The flow ratio <po (the independent variable) is deter- 10 P = 13.0 ppsia
o
mined by adjustment of the pump discharge valve and hence Kt„ - .39 b = 0.300
pressure Pd- At the maximum Pd setting, the mixing zone is Kdl. - .26
located near the throat inlet; further increase of P d will cause
flooding or backflow into the air suction port. As Pd is reduced
in finite steps, air is progressively induced in greater quantities
and simultaneously the mixing zone moves downstream in the i Flow Ratio
throat. The initial data points (at <£„ = 0.6 in Fig. 5) correspond o
to Lm = 8, that is to mixing completed about eight diameters Fig. 8 Theory-experiment comparison including behavior of mea-
sured friction coefficients

Journal of Fluids Engineering SEPTEMBER 1 9 7 4 / 219


Downloaded From: https://fluidsengineering.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use
and 0.14 versus 0 for the SE2 nozzle). The E 3 is similar to E l
but with a short divergent section; the jet spread too rapidly
and obstructed air flow at the throat entry. The conical nozzle
CI results are complicated by apparent boundary layer separa-
tion near the exit, which caused a very high nozzle loss (Kn! =
0.25) and a partially contracted jet. As a result the jet-to-throat
area ratio was slightly less than 0.30, a parametric change which
reduces the slope of the ij,> (0„) curve for LJG pumps.
Nozzle comparisons at one throat length {LT = 28 in Fig. 6)
may be misleading because an optimum throat length exists for
each basic nozzle design. For example the correct length for the
CI conical nozzle is approximately 48, compared with only
19-20 dia for the SE2 nozzle, and added length would improve
the CI nozzle pump performance.
The square-edged orifice nozzle was selected as standard for
this study of confined jet break up. This design has a number of
attributes which may be summarized by noting t h a t : (a) air
induction was maximized, (6) required mixing throat length was
minimized (about one-half the length for a converging nozzle),
(c) operation was stable for all locations of the mixing zone (con-
verging nozzles such as CI produced surging a t low flow ratios,
i.e., at high pump discharge pressures), and (<2) the square-edged
orifice is the most reproducible, and immune to scaling (size)
effects. (The multihole nozzle is the least desirable in this re-
gard, particularly when the pump parts are small in size.)
Apparently the superior air induction effectiveness results
from the absence of boundary layer buildup upstream, and the
flat (or even concave) velocity profile of the jet immediately
downstream of the orifice. Results from studies of breakdown of
submerged jets tend to confirm these findings: jets from square-
edged orifices exhibit higher breakdown frequencies and con-
sequently shorter breakdown distances than jets from converging
nozzles with developed velocity profiles [16, 17, 18].

Fixed Throat-Length Tests. A series of tests was made to ob-


serve mixing length (Lm) response to suction port pressure P „
jet velocity and area ratio 6. The throat length LT was held
constant. Some interesting data were obtained but the results
were of little help toward understanding liquid jet breakup
phenomena. T h e use of Z, or P, or 6 as the flow parameter
A Flow Ratio simultaneously changes the jet pump number (n = 2Zb'ic/P0).
Consequently the results are masked by the changes in mixing
Fig. 6 Effect of nozzle design (Fig. 4) oh LJG pump performance
losses, which vary directly as n. The fixed-length tests did sug-
gest relatively small responses of L„ to Z and P, and great
sensitivity to area ratio 6. The balance of this paper is con-
performance: a slight impairment of pump performance is re- cerned with variable throat-length tests, aimed at developing a
flected by large changes in the corresponding K coefficients. T h e means of predicting correct LJG pump throat lengths.
K value divergence from the constant values (for 0 O > <£</)
Variable Throat-Length Results. The importance of throat
coincides with the movement of the mixing zone into the dif-
length is shown in Fig. 7. Dimensionless throat length L T is
fuser. This K behavior is of no practical design consequence be-
cause the L J G pump should be operated only at flow ratios to
the left of 4>o', where the K's are constant. These results demon-
strate that the K values, which are quite comparable to those
for liquid-liquid pumps, can be assumed constant in designing
pumps for operation over a wide range of flow ratios. (The K
values measured in this investigation are summarized in the
discussion of Fig. 8.)

Nozzle Design Effects. The early work at Penn State was


based on converging nozzles such as CI in Fig. 4, selected for
single-jet simplicity and ease of jet area definition. Multiple jet
nozzles are known to promote more rapid mixing in single phase
pumps. Witte [11] confirmed this for his LJG pump and in ef-
fect optimized design (with fixed 6 and throat) by trying a suc-
cession of multihole nozzles; 19 holes gave the best results.
Shown in Fig. 6 are performance curves for five nozzles with
b = 0.30, LT = 28 and other parameters constant. T h e square-
edged orifice was superior both in terms of the departure flow
ratio 0 / and efficiency r]JP. The poorer results with the E l and
7H nozzles are largely due to higher nozzle losses {K„, — 0.10 Fig. 7 Performance curves, response to throat length LT

220 / S E P T E M B E R 1974 Transactions of the ASME


Downloaded From: https://fluidsengineering.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use
Nozzle SE2, b » .300
LT = 28 Z =• 152, R., - 13.0 10 20 30
1.0 n > 4.91, S • 3
LT throat length, diameters
r~ K th .43
Fig. S Throat-length effeet sn theory experiment departyrs for
efficiency and flow rati® >h

01—
1.0 1.2 1.4 1.6 in Figs. 7 and 8. Apparently Kdi tends to be smaller for LT
2.0
values near the optimum length and this is attributable to the
* 0 Flow Ratio
beneficial effects of intense mixing at the diffuser entry which
Fig. S Throat and diffyser las* coefficients versus flew ratio mea- prevents or delays phase separation. Variations of the K curves
sured with throe throat lengths for 4>0 > 4>o, i.e., beyond the departure points are not of practical
concern since this is an off-design condition. The Kth values rise
for 4>o > 4>o', because the pressure recovery from momentum
the test parameter with b = 0.300 and n = 4.91. The LT = 10 transfer is partially lost as mixing occurs in the diffuser. Con-
pump gave poor results, i.e., a maximum efficiency of 30 percent versely, Kdi values drop for 4>, > 4>J because a limited amount
and a departure flow ratio <j>0' of about 1.1. Increasing the throat of mixing in the diffuser entry is beneficial. Negative Ka values
length to LT = 16, 22, and 28 dia provided much better <j>„' which occur for <$><>>> <$><>' simply attest to the departure from
levels. The longer throat lengths caused greater Motional losses the essentially one-dimensional homogeneous flow which is ap-
and thus depressed the efficiency at a given flow ratio (compare parent for 0,, < <t>„'.
at 4>0 = 1.0). The curves suggest that an optimum exists since
the $0' gain for LT - 28 over the LIT = 22 pump was marginal Summary of Friction Coefficients. Friction coefficient results
and at the price of a drop in pump efficiency. The theoretical may be summarized as follows: Ka< — 0.26 to 0.35 for all con-
curve shown in Fig. 7 is for KM, = 0 and this curve of course lies ditions; Kth = 0.32 to 0.48 for optimum throat lengths; Ka =
above all four sets of experimental data. (By replacing K,?, = 0 0.15 and = 0.60 for short and for very long throat lengths, re-
with the appropriate K,n values [see Fig. 8] experiment-theory spectively. Because the gas throat-entry pressure loss is neg-
comparison similar to Fig. 5 would result.) ligible, Km = 0 is recommended. Nozzle coefficients may be
taken from direct tests or from the liquid-liquid jet pump liter-
The corresponding K curves for LT = 16, 22, and 28 appear ature. For square-edged orifice nozzles with accurate contrac-
in Fig. 8. The constancy of K values is again evident; the aver- tion coefficients (jet area is based on the vena contracta) K„, = 0
age Ku, and Kdi values in the 4>„ < </>/ regions are shown for is applicable. Nozzle coefficients corresponding to Fig. 6 are
each curve. Note that K,h increases from 0.15 to 0.43 in propor- K„, = 0, 0.10, 0.25, 0.14 and 0.16, respectively, for the SE1,
tion to length LT. Additional data including effective friction El, CI, 7H and E3 nozzles.
factor values/(A = Ktli/LT were reported in [15]; the approxi-
mately linear increase in Kth with LT will facilitate prediction Shown in Fig. 9 as a function of throat length are maximum
of performance with excess LT. Subsequent tests showed that jet pump efficiencies and corresponding inlet flow ratios <f>„''.
LT = 19 - 20 was optimum for 6 = 0.300, i.e., slightly shorter These data were cross-plotted from performance curves for pumps
than for the LT = 22 center curves in Fig. 8. The flatness of the with mixing throat lengths from ten to 38 dia in length. First,
K(4>„) curves for <j>o < 4>o will fail under extreme LT conditions: confirming the suggestion in Fig. 7, there is an optimum length
(a) very short lengths prevent complete mixing in the throat (19-20 dia) for this 6 = 0.300 pump operated at a jet pump
and the K's vary continuously; (or LT = 10 the range was 0.3 number of n = 4.91. Secondly the plot of flow ratio <f><>' as a
to 1.4 for K, and —0.2 to 1.5 for Kdi [151; (&) very long throats function of mixing throat lengths reveals two flow regimes.
h
at low area ratios may promote phase separation and the re- Two Flow Regimes. In Fig. 9 there is a relatively rapid re-
sulting impairment of pressure recovery will be reflected by sponse of air induction effectiveness (hence 4>o) to increased
abnormally high Kdi values. throat lengths up to the optimum value of LT = 19-20 dia. The
The diffuser coefficients are relatively immune to LT varia- slope then changes abruptly, i.e., there is a significantly reduced
tions. A value of Kd< = 0.30 is representative for the series response of ($><,' gain for throat lengths beyond LT ~ 19-20.

Journal of Fluids Engineering SEPTEMBER


Downloaded From: https://fluidsengineering.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use
1974/ 221
50 ~ T - — i 1 r-
fVoizle S£?
b • .153
Z • 152 psl
40 - P ; - ) 3 . 0 psia
= 3.03
S =• 3

M
30

Inadequate Throat Length

J
20

10

Oi—. 1 _J 1 I I i i
0 1.0 2.0 3.0 *.0

* 0 Flow Ratio

Optimum Throat Length Fig. 11 Mixing throat performance: effect of length

low pump efficiencies. The right-hand curves on the other hand


represent the desired condition: lower compression ratios but
at relatively high volumetric flow ratios thus producing high
jet pump efficiencies.
Observations concerning Figs. 11-12 are as follows:
1 Forcing short breakup lengths (by high P„ pressures) causes
Excessive Throat Length extremely rapid rates of pressure rise (curves A, B, C).
Fig. 10 Mixing flow regimes: Impact at top, and jet disintegration in
2 With adequate throat lengths the mixing process is relative-
middle and bottom sketches ly gradual and high flow ratios and pump efficiencies are ob-
tained—see curves G, H and I and the corresponding efficiency
(Air entrainment in free liquid sprays [19] exhibits a linear points on Fig. 11.
relation between air volume and distance along the spray axis 3 Although the rates of pressure rise for the proper throat
up to a point where the "slope" changes suddenly.) lengths (curves G, H and I) are relatively gradual, note that the
These results coupled with visual observations of the behavior pressure increase—hence mixing—spans approximately 28 dia
of the mixing zone in the transparent throat suggest that there for this LJG pump which is very large indeed compared with the
are two rather distinct mixing regimes in the LJG pump mixing 4-8 dia mixing length which suffices for liquid-liquid or gas-gas
process. These regimes are illustrated by the sketches (not to jet pumps. Additionally, the mixing zones for curves G, H and
scale) in Fig. 10. The top sketch illustrates impact mixing. This I are preceded by approximately ten dia of "inert" throat
phenomenon occurs in short throats, and also in adequate-length lengths, in which essentially no pressure increase occurs because
throats when the back pressure Pd is sufficiently high to force the jet remains virtually intact with little shedding of droplets
the mixing zone to positions near the inlet of the throat. In the and hence no compression of the secondary fluid.
top sketch the potential core of the jet is diminished relatively 4 The coincidence of the LT = 22 and LT = 28 throat
little by the shedding of high velocity droplets and hence gas lengths (curves D and E in Fig. 12) shows that the mixing process
entrainment effectiveness (and <t>0') is very poor. The middle is essentially independent of the downstream diffuser section of
sketch in Fig. 10 corresponds to the optimum mixing length the LJG pump.
(19-20 dia in Fig. 9). The added length permits normal jet Previous investigators [2, 3, 6, 8, 11] have cited extremely
disintegration, i.e., exhaustion of the jet core, at the entry to rapid rates of throat pressure rise as characteristic of LJG
the mixing zone. The resulting shedding of high velocity droplets pumps. Witte [11] referred to this phenomenon as "mixing
for a large axial distance provides good air entrainment and hence shock." Fig. 12 reveals however that gradual rates of pressure
considerably higher breakway flow ratios $„'• The bottom sketch rise along the throat will obtain if adequate throat length is
shows the effect of further addition of mixing throat length
beyond the optimum. Although the length of the mixing zone is
increased, the exhaustion of the high velocity potential core
permits only limited improvement in air induction effectiveness,
thus explaining the (Fig. 9) change in slope of the departure
flow ratio <t>0' versus throat length beyond the optimum point.

Static Pressure Profiles. The behavior of static pressure pro-


files in the mixing throat helps to explain further the two-
regime hypothesis. Referring first to Fig. 11, efficiency data are
plotted for three throat lengths: LT = 22, 28 and 38 dia (6 =
0.153 and n = 3.03). The beneficial effect of added length is
obvious. Note also that any additional friction losses from
longer throats are negligible at this lower b value (compare Fig.
7). (A relative friction response to area ratios which has been
predicted by theory [1].) The letters A, B, C, etc., along the
Fig. 11 curves identify the corresponding static pressure pro-
files plotted versus throat length in Fig. 12. The left-hand pres-
sure profiles show rapid increases to relatively high compression
Fig. 12 Pressure distribution In mixing throat, for thr«» I T throat
ratios and correspond to very low inlet flow ratios <£„ and hence lengths

222
Downloaded From: /SEPTEMBER 1974
https://fluidsengineering.asmedigitalcollection.asme.org Transactions of the ASME
on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use
~r— r™" ~~~ 1™ - — — ~ T ~ ~ — < — ~ T -

Nozzle SE1
b = .153
Z = 152 p s i
/ °^r^'^
-P = 13.0 psia
-
r, = 3.03 p
S = 3

a -/ 0
Data Type Method -
P $ ' Departure Point A
Sk
a J- A $ n " Departure Point 8
&/o O Pressure Distribution C

" /
Fig, 14 Mixing l@ngth-f!ow ratio correlations for three jet vaiochioa
(mathod C pressure distribution data)

is attributable to the fact that the 0 O ' method is influenced by the


diffuser: for short throat lengths, the diffuser performance (and
hence pump efficiency) benefits from the delayed mixing in the
entrance of the diffuser. This is also revealed by the behavior
of measured Kat values which may be quite low or even negative
when 4>0 > 4>o—see the Kdi curves in Fig. 8 for L T = 10 and
22. This phenomenon causes the Method A procedure to call for
slightly shorter required mixing throats for a specified <f>0 than
Methods B and C which are independent of the diffuser and
predict more conservative (longer) throat lengths. Method C
was adopted for the remainder of the study of jet breakup
phenomena.

Dimensionless Correlations of Jet Breakup


0 10 20 30 10
Lengths
Length, Mixing Throat Diameters
Dimensionless mixing lengths are plotted versus flow ratio in
Fig, 13 Comparison of methods A, B,.and C for Measuring departure
points; and 4>0 flow regime change Fig. 14 for three jet velocities. Note that (a) the slopes of the
Lm/Dt(4>o) curves are dependent on jet velocity, (6) the Lm/
D,(4>„) relations are linear at the lower flow ratios, but the points
provided. Very rapid rates of pressure rise—of the order of a breakaway from the straight lines, curving upward at the higher
few diameters—are the (undesirable) result of impact mixing flow ratios, (c) the breakaway curvature becomes progressively
which occur when liquid jets are discharged into essentially sharper as the jet velocity increases, i.e., as Z is increased from
flooded mixing throats. 40 to 96 to 152 psi, and (d) the breakaway points for the two
Assessing throat length effects by obtaining a complete per- highest Z curves are approximately the same, 19-20 throat dia.
formance curve (efficiency as a function of inlet flow ratio) for (This may also be true for the Z = 40.4 curve, but the curvature
each assembly and repeating this process for a series of pumps of is so gradual that it is difficult to identify a distinct breakaway
varying throat lengths is laborious and time consuming: seven point with any degree of accuracy.)
to 15 test points must be measured to plot the curve required to The effects of suction port pressure P, are shown in terms of
establish peak efficiency and <j>0' for one throat length at one Lm/D,(4>0). in Fig. 15. Note t h a t (a) data for all three pressures
operating condition. Accordingly, the use of static pressure fall on the same straight line at low flow ratio values, (6) the
profiles to define jet breakup or mixing lengths was investigated departure point for all three curves is 19-20 dia, (c) to the right
as a replacement for the measurement of <£</ departure points. of this separation point the data form three curves. Consider
Defining the distance between inlet and attainment of maximum the curves at 4>o = 1-4: Lm/Dt varies inversely as P„ i.e., mixing
throat pressure as the mixing length, a series of tests was made to length was shortest for the highest air pressure. While these
compare the performance curve (method A) versus pressure pro- data are limited in scope, the trend is consistent with a free-jet
file (Method C). Results are plotted at the top of Fig. 13. Also study [20] which predicted an infinite breakup length for a jet
included in this plot are data for Method B, based on the de- in a vacuum, if surface tension is neglected.
parture of throat efficiency points from the normal or theoretical The third variable in the jet pump number n is the area ratio
curve (Fig. 5 illustrates such data). T h e bottom curves in Fig. b, and Fig. 16 shows mixing length data for 6 = 0.153 compared
13 show that Methods A and B yield essentially similar informa- with b = 0.300. Note (a) one Lln/Dt(4>o) curve represents both
tion concerning efficiency as an index to optimum mixing throat sets of data before breakaway, and (6) the breakaway point is
length (32-33 dia for this 6 = 0.153 pump). Regarding maximum highly dependent on area ratio 6: 19-20 dia for b = 0.300 and
flow ratio versus length the top curve indicates that Methods A, 32-33 dia for b = 0.153. These lengths are also the optimum
B and C yield essentially the same results. Again—as in Fig. throat lengths from a pump efficiency >/,;> standpoint, as shown
9—two distinct regimes are evident in the top curve, i.e., be- in Figs. 9 and 13.
yond the optimum length of 32-33 dia the slope of the cj>0(LT)
To summarize, the Lm/D,(4>0) plots in Figs. 14, 15 and 16
curve decreases noticeably. Further increase in LT beyond
show that linear relations exist up to a breakaway point which
32-33 dia causes a loss in ?/;,, due to friction, and with only
coincides with the optimum length throat. The jet velocity and
marginal improvements in <jf>„'. Examination of the coded data
suction port pressure P, exert little effect on jet breakup length
points shows that the 4>0' departure points (Method A) fall some-
(hence optimum LT), compared with the significant influence of
what above the Method B and C points. The slight divergence
area ratio b.

Journal of Fluids Engineering SEPTEMBER 1 9 7 4 / 223


Downloaded From: https://fluidsengineering.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use
S^nbo]_

• Flow Patio

Fig. 15 Mixing length-flow ratio correlations, with variable P, (nwthod


C data)

1.2 1.4x10"-
The remaining task is to identify a dimensionless correlation
that will collapse all of the data to a common curve. Free jet Fig. 17 Mixing lengths versus flow ratio/Reynolds number ratio
breakup studies [21, 22] have been expressed in terms of length
in jet diameters plotted versus the reciprocal of the jet Reynolds
number. Fig. 17 shows that this correlation is apparently ap-
plicable for the L J G pump. Mixing lengths in Fig. 17 are non- of pumps with LT :S LTovt whenever the mixing zone is posi-
dimensionalized with jet diameter Du and the flow ratio 4>„ is tioned (off-design) at Lm < LT0!,t as a result of setting Pd a t
divided by the jet Reynolds number ??. As a result data forming
discharge pressures in excess of the design P,i level.) Breakup
the several curves in Figs. 14, 15 and 16 are collapsed to one
length in the impact regime is affected by jet velocity Via,
relation, L,n/Du = 7.86 X 10° <£»/£?, up to the three breakaway
velocity ratio v, area ratio 6, and by the jet Reynolds number.
points. Included are data for b = 0.153 and 0.300, and in addi-
The linear relation permits calculation of Lm from flow ratio (j>0
tion 0.449, and for three values each of jet velocity heads Z
and conversely.
and suction port pressure P,. The breakaway points are evident-
2 The breakaway point Ln — 15Di0c represents the optimum
ly dependent on area ratio 6, while velocity and suction port pres-
LT values for the range of V\„, P, and b values in this study, us-
sure effects have been "eliminated" through use of these coor-
ing water and air as the liquid and gas components. Apparently
dinates.
jet velocity and suction port pressure exert only second order ef-
The final correlation is shown in Fig. 18. By dividing the di- fects on the optimum value of LT for the confined mixing of a
mensionless length L/Di0 and the parameter 4>0/1? data by the liquid jet and gas annulus. The formula optimum lengths in
Aio 1 —b throat and jet diameter terms are:
area ratio c = —- = —-— , all of the data up to breakaway
An b
fall on a common curve Lm/Dioc = 7.86 X 10° vfR. The coor- LJD, LJDi0
dinates of the common breakaway point, in Fig. 18 are v/1?. =
1.91 X 10-6 a n d LJDloc = 15. 0.153 32.4 83.0
The significance of Fig. 18 may be summarized as follows: 0.300 19.2 35.0
0.449 12.3 18.4
1 The empirical expression Lm/Dioc = 7.86 X 10« v/1Z. de-
scribes the jet breakup length behavior in the impact flow regime The data used in arriving at Lm/Duc = 15 as the optimum point
up to the optimum LT length. This flow regime obtains for an were based primarily on the 6 = 0.153 and 0.300 pumps. D a t a
LJG pump with LT < LTopt. (It can also occur in the operation for 6 = 0.449 were few in comparison, b u t Figs. 17 and 18 show
that the breakaway point (hence optimum length) for b = 0.449
fits the behavior pattern established with the smaller — b pumps.
3 From a design standpoint, 1 and 2 describe the useful
l * 152.0 pst
Ps • 13.0 psia
5 • 3

Nozzle b _b "j L_
O SE! .153 .153 13.0 152.0
A SE2 .300 .153 12,5 244.0
.300 13.0 40.4
.300 13.0 96.3
.300 13.0 152.0
.300 9.0 95.3
.300 4.58 96.3
.449 13.0 71.5

! ,/0 1/c . 7.86x10' i

•4 .8 1.2 1.6 2.0 2.4 2.8x10


( e 0 / " ] / c = •-./"

Fig. 16 Mixing lensth-flow ratio correlation for b =0.153 and 0,309 Fig. 18 Mixing l«ng ths ¥ersus inl©t velocity ratlo/Reynolitt number
(msthod C data) ratio

224 / S E P T E M B E R 1974 Transactions of the ASNtE


Downloaded From: https://fluidsengineering.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use
jet. For the special case of a free jet the effect of jet velocity on
the jet length required to achieve breakup is complex [21, 22] as
shown by the sketch a t top of Fig. 19. At low jet velocities, points
A to C in the sketch, the breakup length increases with jet
velocity. Beyond a critical point the reverse is true: the breakup
length decreases with further increase in jet velocity. There is
no agreement in the literature concerning the course of events
beyond point D.
Leinhard and Day [23] have shown t h a t the breakup length
of high-Reynolds number jets can be correlated as a function of
the Weber number and Reynolds number as shown in the lower
Jet Velocity V J 0
sketch in Fig. 19. Leinhard's Reynolds number conditions ap-
proximate those in this study, and his curve is shown in the mid-
dle of Fig, 19. Our optimization results from correlations similar
to Figs. 9, 11 are shown as data points on these same coordinates
for b = 0.153 and 0.300. Unfortunately, testing of the 6 = 0.449
pump did not produce sufficient data to permit plotting points
200 on these coordinates. The broken horizontal lines represent the
empirical equation Lm/Di0 = 15c for all three b values. Fig. 19
suggests t h a t confinement of the liquid jet in a mixing tube
150 rapidly accelerates jet breakup, i.e., as 6 —> 1.0 very short throats
o D Optimum length data will suffice. Conversely, as b —> 0 the jet breakup length will ap-
Eq. Lm/D1Q - 15c
proach t h a t of the free jet. 'The expression Lm/Di0c — 15 is in
agreement, i.e., it predicts an infinite L,„ as b —» 0, and a zero
100 b = .153
Lm as 5 —* 1.0. The horizontal line for the b = 0.300 pumps sug-
gests t h a t the influence of b overpowers surface tension, viscosity
and jet velocity effects as reflected in the Weber and Reynolds
b = .300
numbers, since the jet breakup length is constant, at least for
2 50
the range of parameters covered in this study.
b = .449

3.0 5.0 7.0xl0" 9 Conclusions and Symmary


2 1 Additional experimental data have confirmed the one-
W/R
dimensional model for the liquid jet gas pump developed in
Fis. 19 Upper: free jet breakup curva; lower: comparison el eottiiiiat!
j e t , md free jet breakup length correlations reference [1], and the constancy of the throat and friction coef-
ficients versus volumetric flow ratio verifies the frictional loss
model.
2 Studies of jet breakup length using a series of LJG pumps
with varying throat lengths have confirmed the existence of
portion of Fig. 18, viz, the optimum point Lm/Dioc = 15 and the optimum throat lengths. Adequate throat length improves air
linear correlation to the left, Lm/Dloc(v/%?); this is because induction, and isothermal efficiencies in excess of 40 percent
throat length should not exceed 2/7'0pt for maximum efficiency. have been achieved.
The behavior of the limited data beyond the breakaway point in 3 The use of throat lengths in excess of the optimum causes
Fig. 18 is potentially interesting, however. Examination of the increased frictional losses in the throat, which depresses maximum
coded data points reveals a pattern. Consider the intersections efficiency while providing only marginal improvement in inlet
at one value of v/1?., 2.2 X 10~"6 for example: (a) Lm increases volumetric flow ratio or inlet velocity ratio. Efficiency loss in
with decrease in P , for constant Z and 6, (6) Lm increases with negligible however if b is small.
increase in Z, for constant P, and b, (c) Lm increases with in- 4 A relatively rapid technique of measuring jet breakup
crease in b (compare the solid circle (b = 0.153) and square lengths by measuring static pressure profiles has been developed.
(b — 0.300) points.) These three parameters comprise the jet The profile method provides mixing length data that agree with
22b J c the more laborious method of measuring peak efficiencies and
pump number n = —— . (P„ = P,). Apparently Lm/Di„ in- flow ratios from performance curves for a series of fixed length
L J G pumps.
creases as some function of n for LJG mixing when L,„ > LTopt.
Additional work in this area is indicated. For practical design 5 Two mixing flow regimes have been identified. Impact
purposes however, the relation Lm/Dloc — 15 is recommended mixing accompanied by very sudden pressure rise—but with
for use in selecting LJG throat lengths for optimum performance. poor air induction effectiveness—occurs when the liquid jet
If a pump is built with LT > 15Dloc and then operated at enters essentially a "flooded" mixing throat. On the other hand,
very gradual rates of pressure rise—and improved air induction
L,„ < 15D l0 c (mixing moved upstream in throat), 1},P and <j>„
effectiveness—occur when adequate throat lengths are provided.
will be impaired and <f>0 will be related to L„, via u/1?, see Fig. 18.
Mixing lengths for area ratio b = 0.15 to 0.45 in this study
On the other hand, a pump with LT > 15Di„c would permit
range from 32 to 12 throat dia respectively, considerably in
Lm > 15DicC Results with such extended Lm values would be
excess of the 4 to 8 dia normally used for liquid-liquid jet pumps.
(a) a minor gain in <j>0 and (b) a loss in efficiency caused by long-
throat friction; losses will be minor in magnitude for small b 6 The relation LJDloc = 7.86 X IWv/R represents the
but significant for large & pumps. response of mixing length Lm to inlet velocity ratio and Reynolds
number for mixing in the throat up to the optimum length
Comparison with Free Jet Breakup Results. As described in the L op t, for liquid jets from square-edged orifice nozzles. According-
introduction, the breakup of a liquid jet starts with the develop- ly, L,„ can be calculated for a given flow ratio 4>o or v, and con-
ment of ripples at the gas-liquid interface which increase in versely.
amplitude until droplets are shed from the surface of the liquid 7 The optimum throat length LTopt is primarily a function

Downloaded From:Journal of Fluids Engineering


https://fluidsengineering.asmedigitalcollection.asme.org SEPTEMBER
on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use 1 9 7 4 / 225
of nozzle-to-throat area ratio b. Based on a large number of 6 Von Pawel-Rammigen, G., dissertation, Braunschweig,
1936.
performance tests including area ratios b = 0.15, 0.30 and 0.45
7 Folsom, R. G., "Jet Pumps with Liquid Drive," Chemical
and for a range of jet velocities and suction port pressures, the Engineering Progress, Vol. 44, No. 10, Oct. 1948, pp. 765-770.
optimum throat length for single jets issuing from square-edged 8 Takashima, Y., "Studies on Liquid J e t Gas P u m p s , "
orifice nozzles can be expressed as Journal of the Scientific Research Institute (Tokyo), Vol. 46, 1952,
pp. 230-246.
M'opt = 15Di„c 9 Bonnington, S. T., "Water J e t Ejectors," B H R A Pub-
lication RR540, Nov. 1956.
thus permitting calculation of LTovt as a simple function of jet 10 Bonnington, S. T., "Water Driven Air Ejectors," B H R A
1 - 6 Publication SP664, Oct. 1960.
diameter Di0 and the pump area ratio c = —-— . 11 Witte, J. H „ "Mixing Shocks and Their Influence on t h e
o Design of Liquid-Gas Ejectors," dissertation, Delft, 1962.
12 Witte, J . H., "Efficiency and Design of Liquid-Gas
8 Among the several nozzle contours tested, square-edged Ejectors," British Chemical Engineering, Vol. 10, No. 9, Sept.
orifices were best in terms of producing liquid jets with superior 1966, pp. 602-607.
air induction effectiveness, ease of fabrication and pump op- 13 Witte, J. H., "Mixing Shocks in Two Phase Flow, Journal
erating stability a t all flow ratios. of Fluid Mechanics, Vol. 36, Part 4, 1969, pp. 639-655.
14 Miller, J. R., " T h e Liquid-Gas Ejector," dissertation,
9 T h e results of this study of jet breakup and optimum N. M . State University, May 1969.
throat lengths coupled with the one-dimensional model should 15 Dopkin, R. J., "The Liquid J e t Gas P u m p : A Study of
facilitate the rational design of liquid-jet gas pumps for practical Jet Breakup and Required ' ' Throat
'"' ' "Length.
"' "' MS thesis, T h e
applications, including the correct throat lengths for maximum Pennsylvania State University, Aug. 1973.
16 Rockwell, D . O., and Niccolls, W. O, "Natural Break-
efficiency. down of Planar J e t s , " ASME Paper 72-FE-5, 1972.
17 Beavers, G. S., and Wilson, T . A., ' Vortex Growth in
References Jets," Journal of Fluid Mechanics, Vol. 44, 1970, pp. 97-112.
18 Sato, H., "The Stability and Transition of a Two-Dimen-
1 Cunningham, R. G., "Gas Compression With t h e Liquid sional J e t , " Journal of Fluid Mechanics, Vol. 7, 1959, pp. 53-80.
J e t P u m p , " JOURNAL O F F L U I D S E N G I N E E R I N G , T R A N S . A S M E , 19 Benatt, F . G. S., and Eisenklam, P., "Gaseous Entrain-
Series I, Vol. 96, No. 3, Sept. 1974, pp. 203-215. ment into Axisymmetric Liquid Sprays," Journal of the Institute
2 Higgins, H . W., "Water J e t Air P u m p Theory and Per- of Fuel, Aug. 1969, pp. 309-315.
formance," M S thesis, T h e Pennsylvania State University, 20 Phinney, R. E., and Humphries, W., "Stability of a
Sept. 1964. Viscous Jet—Newtonian Liquids," United States Naval Ord-
3 Betzler, R. L., " T h e Liquid-Gas J e t Pump Analysis and nance Laboratory, NOLTR-70-5, Jan. 1970.
Experimental Results," MS thesis, T h e Pennsylvania State 21 Grant, R. P., and Middlemen, S., "Newtonian J e t
University, June 1966. Stability," AIChE Journal, Vol. 12, No. 4, 1966, pp. 669-678.
4 Hoefer, K., "Experiments on Vacuum Pumps for Con- 22 Miesse, C. C , "Correlation of Experimental D a t a on t h e
densers," V.D.I. Forshung. Geb. Ihg. Weseng, No. 253, 1922. Disintegration of Liquid J e t s , " Industrial and Engineering Chem-
5 Bonnington, S. T., and King, A. L., " J e t Pumps and istry, Vol. 47, No. 9, Sept. 1955, pp. 1690-1701.
Ejectors, A State of the Art Review and Bibliography," B H R A 23 Lienhard, J . H., and Day, J . B., "The Breakup of Super-
Fluid Engineering, Cranfield, Bedford, England, Nov. 1972. heated Liquid J e t s , " ASME Paper No. 69-WA/FE-19, 1969.

N A M E S O F 1973 R E V I E W E R S (Continued from p. 194)

C. Rodgers J. G. Skifstad Jean-Claude Tatinclaux L. Weaver


W. T. Rouleau A. M . O. Smith D. B . Taulbee H. E. Weber
G. Rudinger T. G. Sofrin H. Tennekes W. W. Weltmer
P. W. Runstadler, Jr. Shao Lee Soo A. Thiruvengadam F. M . White
C. J. Sagi G. Sovran P. A. Thompson R. D . Whitney
D . Salinas E. M. Sparrow W. E . Thompson D. C. Wiggert
N. L. Sanger L. A. Spielman A. R. D . Thorley P. O. Witze
T. Sarpkaya G. S. Springer W. G. Tiederman D. J. Wood
H. A. Scarton F. W. Staub S. C. Traugott B. J. C. Wu
D . J. Schneck M. A. Stoner L. M . Trefethen
M. A. Schultz E . Benjamin Wylie
V. L. Streeter K. L. Treiber
D. Scott B. Sturtevant A. Trikha Ji Wu Yang
G. K. Serovy D. A. Summers C. W. VanAtta T a h Teh Yang
M. M . Sevik S. P . Sutera G. B . Wallis D . F . Young
L. E . Sissom W. M . Swanson H. A. Walls S. W. Zelazny
G. D. Skellenger I. Swiecicky R. L. Wang W. Zielke

Downloaded From:226 / S E P T E M B E R 1974


https://fluidsengineering.asmedigitalcollection.asme.org Transactions
on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use of the ASME

You might also like