A Design Method For Low-Pressure Venturi Nozzles
A Design Method For Low-Pressure Venturi Nozzles
A Design Method For Low-Pressure Venturi Nozzles
School of Mechanical, Industrial, Manufacturing Engineering, Oregon State University, Corvallis, OR 97331, USA;
[email protected] (H.O.); [email protected] (T.M.); [email protected] (X.Z.);
[email protected] (J.L.)
* Correspondence: [email protected]; Tel.: +1-541-706-2093
Abstract: The purpose of this work is to provide empirical design models for low-pressure, subsonic
Venturi nozzles. Experimentally validated simulations were used to determine the effect of nozzle
geometry and operating conditions on the suction ratio (ratio of suction mass flow rate to motive
mass flow rate) of low-pressure, subsonic Venturi nozzles, over a wide range of geometries and
operating conditions, through a parametric study. The results of the parametric study were used to
develop seven empirical models, each with a different range of applicability or calculating a different
indicator of nozzle performance (i.e., suction ratio, momentum ratio, or dynamic pressure ratio), of
the Venturi nozzles using a constrained multi-variable global optimization method. Of the seven
empirical models, the best models were found to be those for low- (less than one) and high-suction
ratios (greater than one), with mean absolute percentage errors of 5% and 18%, respectively. These
empirical models provide a design tool for subsonic, low-pressure Venturi nozzles that is more than
an order of magnitude more accurate than a governing equation approach or conventional flow head
calculations. These newly-developed empirical models can be applied for initial nozzle design when
precise suction ratios are required.
Citation: O’Hern, H.; Murphy, T.;
1. Introduction
Zhang, X.; Liburdy, J.; Abbasi, B. A
Design Method for Low-Pressure Venturi nozzles use a fast-moving motive fluid stream to entrain a nearly quiescent
Venturi Nozzles. Appl. Mech. 2022, 3, suction fluid (Figure 1). In a Venturi nozzle, the motive stream is accelerated by flowing
390–411. https://doi.org/10.3390/ through a converging section, with the highest velocity achieved at the throat of the nozzle.
applmech3020024 The high velocity of the motive fluid creates a region of low static pressure and therefore
a pressure difference between the motive fluid at the throat of the nozzle and the suction
Received: 15 February 2022
Accepted: 31 March 2022
fluid. The pressure difference draws the suction flow into the nozzle, where the suction
Published: 2 April 2022
and motive streams mix before leaving the nozzle outlet. Thermal ejectors can be used to
achieve the same suction and mixing but have a few more internal parts and are typically
Publisher’s Note: MDPI stays neutral in the supersonic regime.
with regard to jurisdictional claims in
Venturi nozzles and ejectors are used in many industries due to their energy efficiency
published maps and institutional affil-
and lack of moving parts [1,2]. The use of such nozzles allows for two streams to mix while
iations.
only using a compressor to move one of the streams, thus reducing the necessary energy
input to operate a system. Venturi mixing nozzles are used in irrigation and fertilization
both to spread water and to mix fertilizers and other chemicals into the water using the
Copyright: © 2022 by the authors.
Venturi effect [3,4]. The concentration of dissolved oxygen in water has also been increased
Licensee MDPI, Basel, Switzerland.
utilizing high-pressure Venturi nozzles [5]. High-pressure or supersonic Venturi nozzles
This article is an open access article are also utilized in refrigeration and chiller applications [6–9]. Variable geometry nozzles
distributed under the terms and have been studied for the application of variable load cooling, where the geometry of the
conditions of the Creative Commons nozzle can be changed as the cooling demand changes [10–12]. Bio-gas injection studies
Attribution (CC BY) license (https:// have also utilized Venturi nozzles to enhance mixing [13]. Venturi nozzles can also be used
creativecommons.org/licenses/by/ for vacuum generation in industrial applications such as vacuum-assisted brakes, powder
4.0/).
Figure
Figure 1.
1. Cross-section
Cross-section of
of representative
representative nozzle
nozzle with
with key
key geometric
geometric parameters
parameters identified.
identified.
Due to the
Despite their widespread
relative simplicity use,ofthe
theperformance
Venturi nozzle andand operation
how wellof these the
known supersonic
Venturi
ejectors
effect is, and high-pressure
it is not straightforward Venturi nozzles have
to calculate been studied
the suction flow rate extensively. In particu-
of these nozzles. The
lar, steamequation
Bernoulli ejector geometry
can be used has to been thoroughly
determine studied
velocity fromfrom a first-principles
a known pressure drop basis,
but as
is
waste steam from industrial processes may be made usable
not applicable to these nozzles because of the mixing of the motive and suction streams.again once entrained in the
nozzle
Gas [17–20].relationships
dynamics Steam ejectors havebe
could been
used studied utilizingthe
to determine CFDlowmethods
pressureasatwell
the as experi-
throat of
mental
the nozzlemethods
based[21–23].
on the Mach number, but the Bernoulli equation or Darcy–Weisbach
Therewhich
equation, has been significant
do not accounteffort to model
for mixing, the behavior
would of high-pressure
still be needed Venturi
to determine nozzles
the suction
and supersonic
flow rate from the thermal ejectors.
calculated pressureKeenan and Neumann
difference. Alternatively,developed
energyahead
one-dimensional
loss calcula-
theory
tions based
could beon gastodynamics
used determine to the
design
outletejectors [24]. and
flow rate, Other first-principal
therefore analyses
the suction flow have
rate,
considered gas dynamics for adiabatic ideal gas air mixing
based on a known pressure drop and major and minor losses across the entire nozzle; and the Bernoulli equation for
incompressible fluid mixing to model the nozzle behavior [3,7].
however, charts and empirical equations for the friction factor are based on constant cross Additionally, second law
analysispipes
section has been usedand
or ducts to define ejector efficiency
it is therefore difficultwith reference determine
to accurately to a reversible ejector, and
for nozzles with it
was found that if the motive and suction fluids are the same fluid,
variable cross sections. Sample calculations for using governing equations and head loss the reversible entrainment
ratio
to efficiency
determine and
the exergetic
suction flowefficiency
rate were are nearly theand
performed same arevalue [25]. Other
presented studies
in Section 3.utilized
CFDIn to determine the effect of geometric features such as the
this paper, we present an empirical model or correlation that can be usedthroat shape, diffuser presence
as a
and angle, and motive inlet shape and diameter, showing that mixing
design guide for low-pressure, subsonic Venturi nozzles for cases of air and air mixing, length, diffuser angle,
as
and effective
well as air andthroatsteamarea are all
mixing. critical parameters
Low-pressure, subsonic to Venturi
nozzle performance
nozzles without [8,9,12,26,27].
diffusers
Additionally,
were the effect
investigated of adding swirl
experimentally, vanes to the
analytically, and nozzle diffuser to
numerically. Theenhance
resultsthe
of turbulent
these in-
kinetic energy has been studied [28]. Cavitation in high-pressure Venturi nozzles has been
vestigations are combined into empirical models for the suction ratio as a function of the
found to further accelerate the flow and suppress turbulent velocity fluctuations [29,30].
dimensionless groups formed from the geometric parameters, operating conditions, and
For Venturi nozzles with incompressible flow, the effect of the injection angle for the suction
fluid has been studied and a correlation for jet trajectory developed with standard error of
0.27 [31]. The effect of the ratio of the length to diameter of the mixing chamber has been
studied for both supersonic and subsonic cases indicating that as the length to diameter
ratio of the mixing chamber increases, the suction flow rate will first increase and then
decrease [32,33].
Only a few studies have considered subsonic ejectors, and those studies typically only
consider the case with air as both the motive and suction fluid [2,12,34]. For subsonic
air-to-air Venturi mixing nozzles, the effect of the angle of the diverging section of the
nozzle has been considered and found to be optimal between 4◦ and 14◦ [2,33,35]. The
angle at which the suction stream meets the motive stream also impacts the performance of
the nozzle. It was found that a larger angle leads to better penetration of the suction stream
into the motive stream [31]. Additionally, any bend or flow separation in the nozzle will
degrade the performance of the nozzle [31]. Predicting the suction flow rate of an arbitrary
nozzle is still not well quantified.
Appl. Mech. 2022, 3 392
This literature review shows that certain geometric features such as diffuser angle and
throat design have been studied for supersonic or high-pressure Venturi nozzles. However,
similar studies of subsonic, low-pressure Venturi nozzles are lacking. This work fills this
gap by creating a design guide for such nozzles. The purpose of this work is to analyze
subsonic, low-pressure Venturi mixing nozzles in order to characterize their performance
and optimum geometry, and to develop empirical models of Venturi nozzle performance
can be used to determine the suction flow rate and inform the design of subsonic, low-
pressure Venturi nozzles. If the suction flow rate of a particular nozzle is known, there
have been multiple studies demonstrating the effect the addition of a diffuser will have on
that flow rate [2,12,33,35].
There are many possible applications for low-pressure, subsonic Venturi nozzles, such
as wastewater treatment. In this application, such Venturi nozzles can be used to accelerate
air on the motive side and entrain wastewater steam on the suction side. In order to
successfully separate clean water from contaminants in wastewater, the humidity of the
air needs to be carefully controlled, which can be achieved by carefully controlling the
ratio of suction flow rate to motive flow rate. Supersonic or high-pressure Venturi nozzles
would be inappropriate for this application because supersonic nozzles would operate at
temperatures too low for water treatment and high-pressure nozzles would increase the
condensation rate of steam, potentially allowing steam to condense before it is separated
from contaminants. Using a low-pressure, subsonic nozzle is an energy efficient way to
control the humidity of air in some wastewater treatment applications [25,36,37]. Many
other chemical and pharmaceutical processes also use such nozzles and would benefit from
the ability to precisely control gas phase mixtures.
In subsonic Venturi nozzles, the suction flow rate is a function of the low pressure
developed, and therefore, the high velocity at the throat of the nozzle. The velocity and
pressure at the throat are dictated by the geometry of the nozzle and the motive stream flow
rate. The static pressure at the suction inlet also influences the suction flow rate: increased
pressure at the suction inlet leads to a larger pressure difference between the inlet and the
throat and thus increases the suction flow rate. In this study, the effect of four different
geometric parameters (Figure 1) on the suction flow rate are studied: the motive diameter
(30–50 mm); the throat diameter (8–16 mm); the diameter through which the suction stream
enters the nozzle, or the suction diameter (15–27 mm); and the distance between the throat
and outlet of the nozzle, or the mixing length (30–80 mm).
Despite the relative simplicity of the Venturi nozzle and how well known the Venturi
effect is, it is not straightforward to calculate the suction flow rate of these nozzles. The
Bernoulli equation can be used to determine velocity from a known pressure drop but is
not applicable to these nozzles because of the mixing of the motive and suction streams.
Gas dynamics relationships could be used to determine the low pressure at the throat of the
nozzle based on the Mach number, but the Bernoulli equation or Darcy–Weisbach equation,
which do not account for mixing, would still be needed to determine the suction flow rate
from the calculated pressure difference. Alternatively, energy head loss calculations could
be used to determine the outlet flow rate, and therefore the suction flow rate, based on a
known pressure drop and major and minor losses across the entire nozzle; however, charts
and empirical equations for the friction factor are based on constant cross section pipes or
ducts and it is therefore difficult to accurately determine for nozzles with variable cross
sections. Sample calculations for using governing equations and head loss to determine the
suction flow rate were performed and are presented in Section 3.
In this paper, we present an empirical model or correlation that can be used as a design
guide for low-pressure, subsonic Venturi nozzles for cases of air and air mixing, as well as
air and steam mixing. Low-pressure, subsonic Venturi nozzles without diffusers were in-
vestigated experimentally, analytically, and numerically. The results of these investigations
are combined into empirical models for the suction ratio as a function of the dimensionless
groups formed from the geometric parameters, operating conditions, and fluid properties.
The empirical model for suction ratio can be used to inform the design of Venturi nozzles
Appl. Mech. 2022, 3 393
given a desired suction ratio. This work allows one to determine the suction ratio of a
Venturi nozzle based on known geometry and operating conditions.
and the temperature of the air to be increased just before entering the nozzle to minimize
condensation in the nozzle. An industrial 12 kW SteamSpa steam generator (9) was used
to supply steam to the suction inlet. The steam generator produces a single source of
steam, which is then split into two hoses (7 and 8) upon leaving the generator. The first
of these hoses (7) was connected to the suction inlet of the nozzle (6), while the second
(8) was directed away from the experimental setup to serve as a bypass for the steam not
Appl. Mech. 2022, 3, entering
FOR PEER the nozzle. The opening of the second hose (8) was restricted using a clamp so
REVIEW
that constant pressure could be maintained at the suction inlet as the motive flow varied
between experiments.
For these experiments, the sources of uncertainty were the Coriolis flow meter, the
Instrumentation locations can be seen in Figure 3, represented by dP for the motive
differential pressure transducers, and the thermocouple. The mass flow rate was measured
pressure drop, P for the gage static suction pressure, and T for the thermocouple at the
using an Endress + Hauser Promass I Coriolis flow meter with an instrument uncertainty
motive inlet to measure
of ±0.5% theofairthe
temperature.
reading. TheOne Setrainlet
suction 230pressure
differential pressure
(9) was transducer
measured using a Setra 230
with a full scale ofdifferential
5 psi waspressure
used to transducer
measure the motive pressure drop from the motiveuncertainty
with a full scale of 1 psi and an instrument
inlet to the outlet.ofA±second
0.25% ofSetra 230 differential
full scale or 14 Pa. Thepressure transducer
motive inlet pressure with a full
drop (10) scale
was of
measured using
1 psi was connected to the
a Setra 230steam inlet,pressure
differential with thetransducer
other sidewith open to ambient,
a full to read
scale of 5 psi and antheinstrument
gage static pressureuncertainty of ±0.25%
at that location. Airoftemperature
full scale or 70 Pa. held
was The type k thermocouple
constant at 105 °C to used to measure the
ensure
air temperature had an instrument uncertainty ◦
±2.2 change
it was above the saturation temperature of the steam to avoidofphase C. in the nozzle.
Air mass flowrates were varied between 1.5 and 4.5 g/s. The T1 and T3 nozzles were
2.3. Air Mixing Determination and Validation
tested, the geometric details of these nozzles are shown in Table 1, with a total of 11 data
CFD simulations were used as a tool to determine the suction flow rate from the
points. Each test was repeated
measured threedrop,
pressure times.such that the simulations, once validated, could provide the
For these experiments, the sources
basis for the empirical model of uncertainty
development. were
ANSYSthe Coriolis
Fluent 19.2flow
[38]meter,
was usedthefor all CFD
differential pressure transducers,
simulations. and the thermocouple.
The geometries The massfrom
and boundary conditions flowtherate was meas-
experiment were used to
ured using an Endress + Hauser Promass I Coriolis flow meter
determine the suction mass flow rate for each experimental case. with an instrument uncer-
tainty of ±0.5% of the For each simulation
reading. The suction case,inlet
the motive
pressureflow(9)rate,
was motive pressure
measured dropaacross
using Setra the nozzle,
and static pressure at the suction inlet were known from
230 differential pressure transducer with a full scale of 1 psi and an instrument uncertainty the experiments. For the case
of ±0.25% of full scale or 14 Pa. The motive inlet pressure drop (10) was measured using asimulation
with no suction, the suction inlet was defined as a wall rather than an inlet in
so no flow could cross the boundary. For the case with suction, the suction inlet was open
Setra 230 differential pressure transducer with a full scale of 5 psi and an instrument un-
to ambient conditions and thus the suction inlet gage static pressure was set to zero. The
certainty of ±0.25% of full scale or 70 Pa. The type k thermocouple used to measure the air
motive inlet was defined as a mass flow inlet boundary. The nozzle outlet was to open
temperature had an instrument
ambient conditionsuncertainty of ±2.2
in all cases, °C.
so the outlet boundary condition was defined to be zero
gage static pressure. The energy model and realizable k-ε turbulence model were the only
2.3. Air Mixing Determination
models used.and Validation
The simulations were steady, to mimic the steady state measurements of
the experiments.
CFD simulations were usedAaspressure-based solver with
a tool to determine thea suction
second-order
flow discretization
rate from the scheme was
measured pressure drop, such that the simulations, once validated, could provide the ba- was also
used for each simulation. The SIMPLE pressure-velocity coupling algorithm
used for each simulation. Additionally, the ambient pressure condition was set to match
sis for the empirical model development. ANSYS Fluent 19.2 [38] was used for all CFD
the ambient pressure of the experiments (88 kPa). Each simulation was considered to be
simulations. The geometries and boundary
converged when all residualsconditions
had valuesfrom the experiment
less than 0.0001 for airwere
mixingused
andto0.001 for air
determine the suction mass flow rate for each experimental case.
and steam mixing. The measured motive pressure drop from experiments was compared to
For each simulation case, the motive flow rate, motive pressure drop across the noz-
zle, and static pressure at the suction inlet were known from the experiments. For the case
with no suction, the suction inlet was defined as a wall rather than an inlet in simulation
so no flow could cross the boundary. For the case with suction, the suction inlet was open
Appl. Mech. 2022, 3 396
the simulated motive pressure drop from CFD to validate each simulation. Three nozzles
were tested for each suction condition.
Symmetry was used so only half of the fluid body (Figure 1) was meshed and simu-
lated. In the simulations, the symmetrical half of the nozzle fluid body was meshed, using
a 3D linear mesh, and mesh size was reduced (increasing resolution) until residuals were
less than 0.0001 and the predicted suction mass flow rate varied by less than 0.2% from
one mesh to the next, indicating that the primary result of interest from the simulation
Appl. Mech. 2022, 3, FOR PEER REVIEW independent of the mesh. The results of the mesh independence study are shown in
was
Figure 4. Table 2 shows the grid refinement study and discretization error with a fine-grid
convergence index (CGI 21 f ine ) of 5.5% and 0.7% for two critical parameters [39].
Figure 4. Mesh refinement study. Once the number of elements was increased past 828,360, there
Figure 4. Mesh
was effectively norefinement
change in thestudy. Once
result and meshthe number ofwas
independence elements
reached. was increased past 828,36
was effectively no change in the result and mesh independence was reached.
Table 2. Grid refinement study and discretization error.
Figures 5 and 6 show the experimental results for the closed and open suction inlet,
respectively, for each tested nozzle compared to the simulation result. For all motive flow
rates and suction conditions, the experimental and simulated results for motive mass flow
rate agree to within ±10%. The no suction case has a mean absolute percentage error of 7.5%
and a root mean square error of 8.4%. The suction case has a mean absolute percentage error
of 5.7% and a root mean square error of 6.5%. The error was calculated as the difference
between the experimental and simulated pressure drop. Based on the ±10% agreement
between the simulation and experiments, the simulations were considered to be validated.
From the validated simulation, the suction mass flow rate can be found.
Table 3. Summary of nozzle geometries in parametric study. The AR code refers to the area ratio
being varied while all other geometric parameters were held constant. Similarly, the LR, T, and S
codes refer to varying the length ratio, throat diameter, and suction inlet diameter, respectively.
Figure 5. Motive pressure drop (Pa) vs. motive mass flow rate (g/s) for nozzle tests with no suction.
Figure 5. Motive pressure drop (Pa) vs. motive mass flow rate (g/s) for nozzle tests with no suction.
See
Appl. Mech. 2022, 3,See
FOR Table
Table 3 for
PEER3REVIEW
for nozzle
nozzle geometry
geometry details.details. Forgeometry,
For the T1 the T1 geometry, the uncertainty
the uncertainty bars
bars are smaller thanare smaller than
9
theexperimental
the experimental marker
marker used.used.
Figure 6. Motive pressure drop (Pa) vs. motive mass flow rate (g/s) for air mixing tests. See Table 3 for
Figure 6. Motive pressure drop (Pa) vs. motive mass flow rate (g/s) for air mixing tests. See Table 3
nozzle geometry details.
for nozzle For thedetails.
geometry T1 geometry,
For the T1the uncertainty
geometry, bars are smaller
the uncertainty than the
bars are smaller experimental
than the experi-
marker used. mental marker used.
Figure 7. Motive pressure drop (Pa) vs. motive mass flow rate (g/s) for steam tests. See Table 3 for
Figure 7. geometry
nozzle Motive pressure
details.drop (Pa) vs. motive mass flow rate (g/s) for steam tests. See Table 3 for
nozzle geometry details.
In summary, both the air mixing as well as the air and steam mixing cases are val-
idated with a maximum
In summary, error
both the air of 10.6%,
mixing andastherefore
as well thesteam
the air and simulations
mixingwere
casesconsidered
are vali-
to bewith
dated validated and trusted
a maximum error ofmoving forward
10.6%, and withthe
therefore a parametric
simulationsstudy
were and correlation
considered to
bedevelopment.
validated and trusted moving forward with a parametric study and correlation devel-
opment.
3. Governing Equations and Flow Head Calculations
As previously
3. Governing discussed,
Equations and Flow some applications,
Head such as water treatment, require precise
Calculations
knowledge of the suction flow rate, or suction
As previously discussed, some applications, such ratio, in order
as water to successfully
treatment, operate.
require precise
Unfortunately, the suction ratio of the nozzles can be difficult to accurately
knowledge of the suction flow rate, or suction ratio, in order to successfully operate. Un- determine
using flowthe
fortunately, design calculations.
suction Asnozzles
ratio of the an example,
can behere governing
difficult equation
to accurately calculations
determine and
using
flow design calculations. As an example, here governing equation calculations and energyto
energy head loss calculations are used to calculate the suction flow rate and compared
the loss
head suction flow rateare
calculations from thetovalidated
used calculatesimulations
the suction [40].
flow rate and compared to the suc-
The governing equation calculation approach was evaluated based on the continuity
tion flow rate from the validated simulations [40].
(Equation (1)), conservation of energy (Equation (2)), and conservation of momentum
The governing equation calculation approach was evaluated based on the continuity
(Equation (3)) equations. For this analysis, a control volume that crosses the throat, suction
(Equation (1)), conservation of energy (Equation (2)), and conservation of momentum
inlet, and outlet of each nozzle was considered. For air and air mixing, Equation (6) was
(Equation (3)) equations. For this analysis, a control volume that crosses the throat, suction
used to determine the enthalpy of the outlet stream. For air and steam mixing, where
inlet, and outlet of each nozzle was considered. For air and air mixing, Equation (6) was
the psychrometrics of the humid air must be considered, Equations (6)–(9) were used to
used to determine the enthalpy of the outlet stream. For air and steam mixing, where the
determine the relative humidity and thus the enthalpy of the outlet humid stream. For both
psychrometrics of the humid air must be considered, Equations (6)–(9) were used to de-
cases, Equations (4) and (5) were used to determine the densities of the air at the throat
termine the relative humidity and thus the enthalpy of the outlet humid stream. For both
and outlet stream. Alternatively, a flow head loss method based on the head form of the
energy equation could be used to calculate the suction mass flow rate based on the major
and minor losses in each nozzle; however, it was found that the head loss method is less
accurate than the governing equation approach, as shown in Figure 8.
Figure 8 shows the suction mass flow rate as determined by the governing equations
and validated simulations versus the motive mass flow rate for each experimental data
point. On average, the governing equations predict the suction mass flow rate with a 270%
error. This method is insufficiently accurate to determine the suction mass flow rate of the
low-pressure Venturi nozzles considered in this study. A different method is necessary to
precisely calculate the suction mass flow rate of these nozzles, and thus inform the design
of the nozzles. The empirical models presented in Section 5 of this paper allow for precise
calculation of the suction ratio and therefore suction mass flow rate.
ℎ = 𝑓(𝜌 ,𝑃 ) (6)
ℎ = 𝑓(𝑇 ,𝑃 ) (7)
𝑚
𝜔= 𝑚 (8)
Appl. Mech. 2022, 3 400
ℎ =ℎ 𝜔ℎ (9)
Figure 8. Head loss predicted suction mass flow rate, governing equations predicted suction mass
flow rate, and validated simulation suction mass flow rate vs. motive mass flow rate. The suction
mass flow rate predicted by the governing equations method is 270% higher than the simulation
result. The suction mass flow rate predicted by the head loss method is approximately 380% higher
than the mass flow rate from the validated simulation. The simulation has an average error of 8%
relative to the experimental data.
. . .
mmotive + msuction = moutlet (1)
. 1 2 . 1 2 . 1 2
mmotive hthroat + Vthroat + msuction hsuction + Vsuction = moutlet houtlet + Voutlet (2)
2 2 2
. . .
Pthroat gage Amotive = moutlet Voutlet − mmotive Vthroat − msuction Vsuction (3)
ρthroat = f ( Pthroat , Tambient ) (4)
. .
mmotive ρthroat + msuction ρsuction
ρoutlet = . . (5)
mmotive + msuction
houtletair = f (ρoutlet , Poutlet ) (6)
houtletsteam = f ( Toutlet , Poutlet ) (7)
. .
ω = msuctionsteam /mmotive (8)
houtlet = houtletair + ωhoutletsteam (9)
4. Parametric Study
In order to determine the suction mass flow rate for different nozzles, a parametric
study was completed in using the validated CFD simulations. Fifteen different geometries
with varying motive inlet diameters, throat diameters, suction inlet diameters, and mixing
lengths were simulated with varying boundary conditions. All geometries are given in
Table 3. Geometries were chosen to provide a range of values for each of the selected
geometrical parameters: motive diameter, throat diameter, suction diameter, and mixing
length. Each parameter was varied to provide at least four different values. These values
Appl. Mech. 2022, 3 401
were chosen such that the average of each geometric parameter provides a fixed suction
ratio to keep the ratio of steam to air below the carrying capacity of water in air for the
majority of steam mixing cases. It was confirmed that flow in each geometry remains
subsonic and incompressible for all relevant conditions prior to including the geometry in
the parametric study.
All flow conditions are shown in the simulation test matrix in Table 4. The motive
inlet was defined to be a mass flow inlet with a flow rate of either 5.2 or 20.8 g/s. These
flow rates were chosen because 20.8 g/s is the desired flow rate for one application of
these nozzles and 5.2 g/s was selected to provide a lower range of suction ratios [33]. The
motive inlet fluid was ideal gas air for all cases. The suction inlet boundary condition was a
pressure inlet with a static gage pressure of either 10 Pa, 100 Pa, or 500 Pa. The suction fluid
was either ideal gas air or steam. For all cases, the nozzle outlet boundary condition was
defined to be 0 Pa gage. Every combination of geometry and boundary conditions summed
to 109 different cases considered in the parametric study. For each case, the suction mass
flow rate and dimensionless suction ratio, or ratio of suction mass flow rate to motive mass
flow rate, were calculated.
Table 4. Test matrix for parametric study. A total of 109 cases were studied. The AR code refers to the
area ratio being varied while all other geometric parameters were held constant. Similarly, the LR, T,
and S codes refer to varying the length ratio, throat diameter, and suction inlet diameter, respectively.
Name/Code Motive Flow Rate (g/s) Suction Static Pressure (Pa) Suction Fluid
20.8 10,100,500 Air, Steam
D2
5.2 10,500 Air, Steam
AR1 20.8 10,100,500 Air, Steam
AR2 20.8 10,100,500 Air, Steam
AR3 20.8 10,100,500 Air, Steam
AR4 20.8 10,100,500 Air, Steam
AR5 20.8 10,100,500 Air, Steam
AR6 20.8 10,100,500 Air, Steam
20.8 10,100,500 Air, Steam
LR1
5.2 10,500 Air, Steam
LR2 20.8 10,100,500 Air, Steam
20.8 10,100,500 Air, Steam
LR3
5.2 10,500 Air, Steam
20.8 10,100,500 Air, Steam
T1
5.2 10,500 Air, Steam
T2 20.8 10,100,500 Air, Steam
20.8 10,100,500 Air, Steam
T3
5.2 10,100,500 Air, Steam
20.8 10,100,500 Air, Steam
S1
5.2 100 Air, Steam
20.8 10,100,500 Air, Steam
S2
5.2 100 Air, Steam
The Buckingham Pi Theorem was used to determine the dimensionless groups that
define this system as:
. " a b . c d !e #
ms Am L mm νs Pstatic Dt4 ρm π 2
. =C . 2 (11)
mm At Ds µ m Dt νm 8 mm
The third independent dimensionless group on the right side of the above expression
.
mm
( µ m Dt )
is the Reynolds number at the throat of the nozzle. The last independent dimen-
Pstatic Dt4 ρm π 2
sionless group ( . 2 ) is the ratio of the gage static pressure at the suction inlet to
8 mm
the dynamic pressure at the throat. Using these definitions, Equation (11) can be written as:
. " a b d !e #
ms Am L c νs Pstatic, suction
. =C ( Remotive, throat ) (12)
mm At Ds νm Pdynamic, throat
The coefficient and exponents of the correlation were determined using a multi-
variable global optimization code in Python. The global constrained minimization algo-
rithm determined the best fit for the coefficient and exponents of the correlation based on
the 109 parametric study cases, using the Levenberg–Marquardt scheme [41,42].
5. Results
Seven different empirical models were developed and evaluated to determine which
parameters are most important to prediction of nozzle performance, and to find which
empirical model is best able to predict the nozzle performance. Details of each empirical
model are given below. Every empirical model considered, as well as their errors and
ranges of applicability are summarized in Table 5 in order to provide a design reference for
Venturi nozzles.
Figure 9. Suction ratio predicted by global suction ratio correlation vs. suction ratio from validate
Figure 9. Suction ratio predicted by global suction ratio correlation vs. suction ratio from validated
simulation.
simulation.
The increase in mean absolute percentage error for the high suction ratio cases ind
If only the low suction ratio cases are considered in the optimization, the result is
cates that the correlation does not well predict the behavior of the mixing nozzles for thos
Equation (20), given in Tableof5.the
cases. Each The low
high suction
suction ratio
ratio correlation
cases predicts number
has a low Reynolds the suction
and aratio
high-pressur
with a mean absolute percentage error of 18%, and a root mean square error
ratio. This indicates that the high suction ratio cases may be driven moreof 22%byasthe applie
shown in Figure static
10. pressure at the suction inlet than the Venturi effect from the motive mass flow rat
If only the high suction
and throat ratio cases
diameter. are considered
Additionally, the lowinsuction
the optimization,
ratio cases allthe
haveresult is
a relatively hig
Equation (21). The high suction ratio correlation predicts the suction ratio with a mean
Reynolds number and a relatively low pressure ratio. If the low and high suction rati
absolute percentage error of 5%, and a root mean square error of 6%, as shown in Figure 11.
Separating the global correlation including both high and low suction ratios into one
correlation for low suction ratio and one correlation for high suction ratio allows for more
accurately informed decisions about the design of a Venturi nozzle geometry, assuming
that the desired suction ratio can be identified as either high or low. Figure 12 shows the
suction ratio predicted by the low and high suction ratios on one plot, with the correlation
used for each suction ratio range.
A sensitivity analysis was conducted by increasing then decreasing the value of each
dimensionless group by 10% compared to the original value and calculating the maximum
relative error, mean absolute percentage error, and root mean square error for each case. The
sensitivity analysis revealed that the area ratio AAmt had the largest impact on the error of
each
correlation of all the dimensionless groups, followed by the kinematic viscosity ratio
.
νs
νm and then the Reynolds number µmm mDt . Comparing the effect of each dimensionless
group between the low and high suction ratio cases, it was found that the geometry has a
larger impact on the suction ratio for the low suction ratio cases than the high suction ratio
cases. The high suction ratio cases are more dependent on operating conditions than the
geometry of the nozzle. These results support the hypothesis that the suction flow for high
suction ratio cases is largely driven by the applied static pressure at the suction inlet, while
the low suction ratio cases are more dependent on geometry because they are truly Venturi
driven flow and the area ratio is critical to the performance.
cases are considered to be driven by different phenomena, inertia and pressure, respec-
tively, then it may be better to model each regime separately.
If only the low suction ratio cases are considered in the optimization, the result is
Appl. Mech. 2022, 3 Equation (20), given in Table 5. The low suction ratio correlation predicts the suction ratio 404
with a mean absolute percentage error of 18%, and a root mean square error of 22% as
shown in Figure 10.
Figure 10.
Figure
Appl. Mech. 2022, 3, FOR PEER REVIEW
10. Suction
Suction ratio
ratiopredicted
predictedby
bylow
lowsuction
suctionratio correlation
ratio vs.vs.
correlation suction ratioratio
suction fromfrom
validated
validated
16
simulation.
simulation.
If only the high suction ratio cases are considered in the optimization, the result is
Equation (21). The high suction ratio correlation predicts the suction ratio with a mean
absolute percentage error of 5%, and a root mean square error of 6%, as shown in Figure
11. Separating the global correlation including both high and low suction ratios into one
correlation for low suction ratio and one correlation for high suction ratio allows for more
accurately informed decisions about the design of a Venturi nozzle geometry, assuming
that the desired suction ratio can be identified as either high or low. Figure 12 shows the
suction ratio predicted by the low and high suction ratios on one plot, with the correlation
used for each suction ratio range.
A sensitivity analysis was conducted by increasing then decreasing the value of each
dimensionless group by 10% compared to the original value and calculating the maximum
relative error, mean absolute percentage error, and root mean square error for each case.
The sensitivity analysis revealed that the area ratio had the largest impact on the
error of each correlation of all the dimensionless groups, followed by the kinematic vis-
cosity ratio and then the Reynolds number . Comparing the effect of each di-
mensionless group between the low and high suction ratio cases, it was found that the
geometry has a larger impact on the suction ratio for the low suction ratio cases than the
high suction ratio cases. The high suction ratio cases are more dependent on operating
conditions than the geometry of the nozzle. These results support the hypothesis that the
suction flow for high suction ratio cases is largely driven by the applied static pressure at
the suction inlet, while the low suction ratio cases are more dependent on geometry be-
Figure
Figure 11. Suction
11.
cause they Suction ratio
ratio
are truly predicted
predicted
Venturi by
bythe
driven high
the suction
high
flow theratio
suction
and correlation
ratio
area vs. vs.
correlation
ratio theto
is critical suction
the ratioratio
suction
the fromfrom
performance. the the
validated simulation.
validated simulation.
Figure 12.
Figure 12.Suction
Suctionratio predicted
ratio by low
predicted byand
lowhigh
andsuction
high ratio correlations
suction vs. the suction
ratio correlations vs.ratio
the suction ratio
from the validated simulation.
from the validated simulation.
5.2. Momentum Ratio and Dynamic Pressure Ratio Models .
For the momentum ratio, the suction momentum term (ρ m ) was considered to be a
s s
Given the apparent dependence of the global . suction ratio correlation on pressure,
function of theempirical
two alternative motive models
mass flow
were rate (mm ),momentum
evaluated: the motiveratio inlet
andarea (Am ),pressure
dynamic the throat diameter
(D t ), the
ratio. suction
For these inleteither
models, diameter (Ds ), the ratio
the momentum mixing length
or the (L),pressure
dynamic the motive fluid
ratio is pre- density (ρm ),
the
dictedmotive
by thefluid
globalviscosity (µinstead
correlation, m ), andof the gage static
the suction ratio. pressure
For each ofatthese
the cases,
suctionthe inlet (Pstatic ),
form of when
which, the correlation can be determined,yields
non-dimensionalized, again, the
using the Buckingham Pi Theorem.
following:
. .
ρs ms = f mm , Am , Dt , L, Ds , µm , ρm , νs , Pstatic (13)
. " a b . c d !e #
ρs ms Am L mm νs Pstatic Dt4 ρm π 2
. =C . 2 (14)
ρm mm At Ds µ m Dt νm 8 mm
When the momentum ratio correlation is optimized, it yields Equation (22), also given
in Table 5. The resulting correlation yields a mean absolute percentage error of 28%, and a
root mean square error of 36% when compared to the validated simulations.
. 2
For the dynamic pressure ratio, the suction dynamic pressure (8 mm /(Dt4 ρs π 2 )) was
.
considered to be a function of the motive mass flow rate (mm ), the motive inlet area (Am ),
the throat diameter (Dt ), the mixing length (L), the motive fluid density (ρm ), the motive
fluid viscosity (µm ), and the gage static pressure at the suction inlet (Pstatic ), yielding
Equation (15) below, which when non-dimensionalized gives Equation (16). The coefficient
and exponents determined using the global optimization are shown in Equation (23). The
global dynamic pressure ratio has a mean absolute percentage error of 48%, and a root
mean square error of 56%. In Equation (16), the mixing length is nondimensionalized
using the throat diameter, rather than the suction inlet diameter as in the suction ratio
and momentum ratio models because the suction diameter is on the independent side of
the equation, but the remaining terms are identical to those of the previously discussed
correlations. .
8m s .
4 2
= f mm , Am , Dt , L, µm , ρm , νs , Pstatic (15)
DS ρ s π
. " . c d !e #
ms Dt4 ρm Am a L b mm νs Pstatic Dt4 ρm π 2
. =C . 2 (16)
mm Ds4 ρs At Dt µ m Dt νm 8m m
Appl. Mech. 2022, 3 406
Table 5. Summary of proposed empirical models, ranges of applicability, mean absolute percentage error (MAPE), and root mean square error (RSME).
Table 5. Cont.
6. Conclusions
A study on the effect of geometry and operating conditions on the suction ratio of
low-pressure, subsonic Venturi mixing nozzles was conducted. An ANSYS CFD model
of the Venturi nozzle mixing was experimentally validated, and then used to calculated
nozzle performance over a wide range of geometries and operating conditions. Governing
equation calculations and flow head calculations were also used to determine the suction
ratio of the experimentally tested nozzles and was found to be very inaccurate in these
cases. To determine the suction ratio more accurately, seven potential empirical models
were developed to examine the effect of different thermophysical parameters on the suction
ratio and identify the parameters most critical to accurate prediction. The foundation for
each of the empirical models is the results of a parametric study of nozzle geometry and
operating conditions.
Appl. Mech. 2022, 3 409
The empirical models for suction ratio are more accurate than the empirical models
for either momentum ratio or dynamic pressure ratio. For the five suction ratio models
developed, the average mean absolute percentage error is 17%. Separating the flow into
high-suction ratio and low-suction ratio regimes had the largest impact on the error of the
models indicating that the regime change is the most critical aspect of nozzle operation.
Based on these results, any of the presented suction ratio empirical models (global, air
only, air and steam mixing only, high-suction ratio, or low-suction ratio) can be used to
determine the suction ratio of a low-pressure, subsonic Venturi nozzle within 22%, or a
more specific model may be used if the application of the nozzle (low-suction ratio, air only
mixing, etc.) with reduced error.
This work can be used to inform the design of low-pressure, subsonic Venturi nozzles
for many applications. The suction ratio empirical models are, on average, 34-foldmore
accurate than the flow head loss approach. The suction ratio empirical models can be used
to determine the suction ratio or nozzle design when precise mixing is required for a given
application.
While the correlations proposed in this study provide a good initial design, it will
be advantageous to have a secondary tool for a more accurate nozzle design. To that end,
these correlations can be used as the basis for physics-guided machine learning algorithms
to serve as a more accurate secondary tool for detailed nozzle design and analysis. The
authors are in the process of developing such a design tool. The results will be evaluated
and published in a follow-up article.
Author Contributions: Conceptualization, H.O., J.L. and B.A.; methodology, H.O., T.M., and X.Z.;
software, H.O. and X.Z.; validation, H.O.; formal analysis, H.O.; writing—original draft prepara-
tion, H.O. and T.M.; writing—review and editing, H.O., B.A. and J.L.; supervision, B.A.; funding
acquisition, B.A. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by US Department of Energy, Advanced Research Projects
Agency—Energy (ARPA-E) award number DE-AR-0001000.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: The authors thank ARPA-E for their support of this research.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Ariafar, K. Performance evaluation of a model thermocompressor using computational fluid dynamics. Int. J. Mech. 2012, 6,
35–42.
2. Meakhail, T.A.; Zien, Y.; Elsallak, M.; Abdelhady, S. Experimental study of the effect of some geometric variables and number of
nozzles on the performance of a subsonic air—Air ejector. Proc. Inst. Mech. Eng. 2008, 222, 809–818. [CrossRef]
3. Manzano, J.; Palau, C.V.; Moreira de Azevedo, B.; do Bomfim, G.V.; Vasconcelos, D.V. Characterization and selection method of
Venturi injectors for pressurized irrigation. Rev. Ciência Agronômica 2018, 49, 201–210. [CrossRef]
4. Bilalis, D.; Karkanis, A.; Savvas, D.; Kontopoulou, C.-K.; Efthimiadou, A. Effects of fertilization and salinity on weed flora in
common bean (‘Phaseolus vulgaris’ L.) grown following organic or conventional cultural practices. Aust. J. Crop Sci. 2014, 8,
178–182.
5. Okzan, F.; Ozturk, M.; Baylar, A. Experimental investigations of air and liquid injection by venturi tubes. Water Environ. J. 2006,
20, 114–122.
6. Little, A.B.; Garimella, S. A critical review linking ejector flow phenomena with component- and system-level performance. Int. J.
Refrig. 2016, 70, 243–268. [CrossRef]
7. Little, A.B.; Bartosiewicz, Y.; Garimella, S. Visualization and validation of ejector flow field with computational and first-principal
analysis. J. Fluids Eng. 2015, 137, 1107. [CrossRef]
8. Lin, C.; Cai, W.; Li, Y.; Hu, Y.; Giridharan, K. Numerical investigation of geometry parameters for pressure recovery of an
adjustable ejector in multi-evaporator refrigeration system. Appl. Therm. Eng. 2013, 61, 649–656. [CrossRef]
9. Varga, S.; Oliveira, A.C.; Ma, X.; Omer, S.A.; Zhang, W.; Riffat, S.B. Experimental and numerical analysis of a variable area ratio
steam ejector. Int. J. Refrig. 2011, 34, 1668–1675. [CrossRef]
Appl. Mech. 2022, 3 410
10. Sun, D.-W. Variable geometry ejectors and their applications in ejector refrigeration systems. Energy 1996, 10, 919–929. [CrossRef]
11. Huang, B.J.; Hu, S.S.; Le, S.H. Development of an ejector cooling system with thermal pumping effect. Int. J. Refrig. 2006, 29,
476–484. [CrossRef]
12. Li, X.; Wang, T.; Day, B. Numerical analysis of the performance of a thermal ejector is a steam evaporator. Appl. Therm. Eng. 2010,
30, 2708–2717. [CrossRef]
13. Bora, B.J.; Debnath, B.K.; Gupta, N.; Saha, U.K.; Sahoo, N. Investigation on the flow behaviour of a venturi type gas mixer
designed for dual fuel diesel engines. Int. J. Emerg. Technol. Adv. Eng. 2013, 3, 202–209.
14. Papadakis, E.; Raptopoulos, F.; Koskinopoulou, M.; Maniadakis, M. On the use of vacuum technology for applied robotic systems.
In Proceedings of the 2020 6th International Conference on Mechatronics and Robotics Engineering (ICMRE), Barcelona, Spain,
12–15 February 2020; pp. 73–77.
15. Rodrigues, A.; Camargo, E.; Ciolfi, M. Venturi Tube Application to Improve the Vacuum Assistance for Brake Systems; SAE International:
Warrendale, PA, USA, 2013.
16. Xu, J.; Liu, X.; Pang, M. Numerical and experimental studies on transport properties of powder ejector based on double venturi
effect. Vacuum 2016, 134, 92–98. [CrossRef]
17. Kroll, E.A. The design of jet pumps. Chem. Eng. Prog. 1947, 1, 21–24.
18. Eames, I.W. A new prescription for the design of supersonic jet pumps: The constant rate of momentum change method. Appl.
Therm. Eng. 2002, 22, 121–131. [CrossRef]
19. Hoggarth, M.L. The design and performance of high-pressure injectors and gas jet boosters. Proc. Inst. Mech. Eng. 1970, 185,
755–766. [CrossRef]
20. Munday, J.T.; Bagster, D.F. A new theory applied to steam jet refrigeration. Ind. Eng. Chem. Process Des. Dev. 1997, 16, 442–449.
[CrossRef]
21. Riffat, S.B.; Everitt, P. Experimental and CFD modeling of an ejector system for vehicle air conditioning. J. Inst. Energy 1999, 72,
41–47.
22. Bartosiewicz, Y.; Aidoun, Z.; Desevaux, P.; Mercadier, Y. Numerical and experimental investigations on supersonic ejectors. Int. J.
Heat Fluid Flow 2005, 1, 56–70. [CrossRef]
23. Fu, W.; Liu, Z.; Li, Y.; Wu, H.; Tang, Y. Numerical study for the influences of primary steam nozzle distance and mixing chamber
throat diameter on steam ejector performance. Int. J. Therm. Sci. 2018, 132, 509–516. [CrossRef]
24. Keenan, J.H.; Neumann, E.P. A simple air ejector. ASME J. Appl. Mech. 1942, 62, 75–81. [CrossRef]
25. McGovern, R.K.; Narayan, G.P.; Lienhard V, J.H. Analysis of reversible ejectors and definition of an ejector efficiency. Int. J. Therm.
Sci. 2012, 54, 153–166. [CrossRef]
26. Yang, X.; Long, X.; Yao, X. Numerical investigation on the mixing process in a steam ejector with different nozzle structures. Int. J.
Therm. Sci. 2012, 56, 95–106. [CrossRef]
27. Ouzzane, M.; Aidoun, Z. Model development and numerical procedure for detailed ejector analysis and design. Appl. Therm.
Eng. 2003, 23, 2337–2351. [CrossRef]
28. Shin, D.H.; Gim, Y.; Sohn, D.K.; Ko, H.S. Development of venturi-tube with spiral-shaped fin for water treatment. J. Fluids Eng.
2019, 141, 051303. [CrossRef]
29. Rudolf, P.; Hudec, M.; Stefan, D. Numerical and experimental investigation of the cavitating flow within venturi tube. J. Fluids
Eng. 2019, 141, 1101.
30. Frankel, S. Large eddy simulation of turbulent-cavitation interactions in a venturi nozzle. J. Fluids Eng. 2010, 132, 121301.
31. Sundararaj, S.; Selladurai, V. Numerical and experimental study on jet trajectories and mixing behavior of venturi-jet. J. Fluids
Eng. 2010, 132, 101104. [CrossRef]
32. Chen, W.; Chong, D.; Yan, J.; Liu, J. The numerical analysis of the effect of geometrical factors on natural gas ejector performance.
Appl. Therm. Eng. 2013, 59, 21–29. [CrossRef]
33. Hu, J.; Cao, X.; He, H.; Meng, Z.; Ding, M. Numerical optimization on the geometrical factors of subsonic air-air ejector. In
Proceedings of the 9th International Symposium on Symbiotic Nuclear Power Systems for 21st Century, Harbin, China, 9–11 July
2018.
34. Maqsood, A. A study of subsonic air-air ejectors with short bent mixing tubes. Diss. Abstr. Int. 2008, 69, 1911.
35. Vaclav, D. Air to air ejector with various divergent mixing chambers. Appl. Mech. Mater. 2014, 493, 50–55.
36. O’Hern, H.; Nikooei, N.; Zhang, X.; Hagen, C.; AuYeung, N.; Tew, D.; Abbasi, B. Reducing the water intensity of hydraulic
fracturing: A review of treatment technologies. Desalin. Water Treat. 2021, 221, 121–138. [CrossRef]
37. Abbasi, B.; Zhang, X.; O’Hern, H.; Nikooei, E. Method and System for Purifying Contaminated Water. U.S. Patent Application
No. 16/985,043, 4 August 2020.
38. Ansys Academic Research Fluent, Release 19.2. Ansys, Canonsburg, PA, USA. Available online: https://www.ansys.com/
academic (accessed on 18 November 2021).
39. Celik, I.B.; Ghia, U.; Roache, P.J.; Freitas, C.J. Procedure for estimation and reporting of uncertainty due to discretization in CFD
applications. J. Fluids Eng. 2008, 130, 078001.
40. Cengel, Y.A.; Cimbala, J.M. Fluid Mechanics, Fundamentals and Applications; McGraw-Hill: New York City, NY, USA, 2014.
41. Transtrum, M.K.; Sethna, J.P. Improvements to the Levenberg-Marquardt algorithm for nonlinear least-squares minimization.
arXiv 2012, arXiv:1201.5885.
Appl. Mech. 2022, 3 411
42. Elhashimi, M.; Zhang, X.; Abbasi, B. Empirical prediction of saline water atomization pressure loss and spray phase change using
local flow pressure analysis. Desalination 2021, 514, 115156. [CrossRef]
43. McMaster-Carr. Available online: https://www.mcmaster.com/venturi/water-aspirator-pumps/ (accessed on 18 November
2021).
44. McMaster-Carr. Available online: https://www.mcmaster.com/venturi/fixed-flow-air-powered-vacuum-pumps-7/ (accessed
on 18 November 2021).
45. Steel Venturi Style Pneumatic Air Blower. Grainger, Lakeforest, IL, USA. Available online: https://www.grainger.com/product/
ALLEGRO-Steel-Venturi-Style-Pneumatic-3TCK2 (accessed on 18 November 2021).
46. Aluminum Venturi Nozzle For Blow Gun M 12x1.25 (MT). Tameson, Eindhoven, The Netherlands. Available online:
https://tameson.com/pneumatics/air-tools/air-nozzles/venturi/aluminum-venturi-nozzle-for-blow-gun-m-12x1p25-mt.
html?id_currency=4 (accessed on 18 November 2021).