1. Introduction
High-pressure water jets have been successfully applied in various fields such as rock breaking [
1], coal mining [
2], vessel cleaning [
3], and machining and cutting [
4], among others. In these engineering applications, the primary advantage lies in the high-impact pressure generated by high-pressure water jets on target objects without inducing adverse factors such as high temperature, sparks, dust, and fragment splashing. Gaining an understanding of the characteristics of high-pressure water jets in air can facilitate their more effective utilization.
Leu et al. [
5] conducted an analysis of the anatomy of high-pressure water jets in air, dividing the entire jet into three distinct regions: the initial (potential core), main (water droplet), and final (diffused droplet) regions. Jiang et al. [
6] captured the entire structure of the high-pressure water jet flow instantaneously using a high-speed camera. They analyzed the overall morphology and velocity characteristics, subsequently reclassifying the flow structure from the conventional three regions into five regions: the potential core region, stripped droplet region, primary shear region, secondary shear region, and diffused droplet region. The principal characteristics of high-pressure water jets in air are that, under the strong action of shear stress at the interface, the flow structure gradually transforms from a continuous jet to a discrete flow dominated by droplets. During this process, due to momentum exchange between water and air, the water is decelerated while the air is accelerated. Rajaratnam et al. [
7] investigated the flow field of very-high-velocity water jets in air using LDA (Laser Doppler Anemometry). They found that the velocity at the centerline of the jet remains approximately constant at
U0 (inlet velocity) within an axial distance from the nozzle of about 100
D (where
D is the nozzle diameter), and then decreases continuously to about 0.25
U0 at an axial distance of 2500
D. Rajaratnam et al. [
8] also measured the water distribution along the axis of high-velocity jets. Their measurements revealed that the water concentration on the axis of the jet decreased rapidly with the axial distance, falling to about 5% at 100
D and 2% at 200
D. It is evident that, due to the significantly lower density of air compared to water, a high degree of air entrainment in the jet will dramatically reduce the jet’s kinetic energy, despite the air having the same velocity as the water phase. Furthermore, the impact pressure on the plate, which results from the transformation of the jet’s kinetic energy, will inevitably decrease significantly. From an application perspective, the pressure distribution on the plate is the most critical aspect. Therefore, greater attention should be directed towards scenarios where the nozzle is not too far from the targets.
Experiments are also a commonly employed method for investigating the characteristics of short-distance impinging high-pressure water jets. Leach et al. [
9] investigated the overall characteristics of high-pressure water jets using short-exposure optical photography and X-ray photography. They also measured the distribution of impact pressure on the flat plate and analyzed it theoretically. In their experiments, the distance from the nozzle exit to the flat plate was 76
D. Additionally, the effect of nozzle shape was also examined in detail. Guha et al. [
10] conducted experimental studies on the pressure distribution on a flat plate (with the distance from the nozzle exit to the flat plate being 43
D) for the purpose of designing an efficient cleaning system. They measured the pressure distributions and found that the pressure along the radial direction can be expressed by a Gaussian distribution. Moreover, the stagnation pressure on the flat plate decreased linearly with axial distance.
Numerical simulation is another widely used method by some researchers to investigate the characteristics of high-pressure water jets. Liu et al. [
11] employed the multiphase volume of fluid (VOF) method in conjunction with the standard
k-
ε turbulence model to simulate the dynamic characteristics of an abrasive jet from a very fine nozzle. Guha et al. [
10,
12] conducted simulations for high-pressure water jet flow using the Eulerian multiphase model in combination with the standard
k-
ε turbulence model and standard wall functions. By introducing a novel model for mass and momentum transfer, they achieved favorable results compared to the experimental findings of Leach et al. [
9]. Liu et al. [
13] comparatively investigated the impingement capability of high-pressure submerged and non-submerged water jets using the renormalization group (RNG)
k-
ε turbulence model. They found that the velocity in the submerged water jet was significantly lower than that in the non-submerged water jet at the same nozzle pressure.
The dynamics of high-pressure water jets in air exhibit significant complexity. Numerical simulation involves multiple challenging issues, including the establishment of multiphase flow models, the selection of turbulence models, and the handling of turbulence in multiphase flows, among others. Crucially, the high flow velocities (typically 100–500 m/s) and elevated Reynolds numbers of these jets render turbulence modeling paramount for the accurate prediction of impact pressures and spray dispersion patterns.
The current literature review indicates that the
k-
ε model and its variants have dominated previous numerical investigations of similar jet systems. While emerging evidence has suggested that elliptic–relaxation-based turbulence models (e.g., the
φ-
α model) demonstrate superior predictive capability for single-phase impinging jets at velocities below 50 m/s [
14,
15,
16], the application of such advanced turbulence closure schemes to multiphase, high-velocity (≥100 m/s) water jets remains unexplored. This study employed a self-developed innovative elliptic-blending turbulence model (hereafter, the SST
k-
ω-
φ-
α model), extending the
φ-
α model through integration with shear stress transport (SST) formulations. By incorporating
ω-based eddy viscosity modulation, this model not only preserves elliptic blending advantages for near-wall treatments but also enhances prediction accuracy in free shear layers. This dual enhancement mechanism explains its proven superiority over the SST
k-
ω model across six benchmark flow configurations—including 2D channel flow, axisymmetric pipe flow, 2D step flow, axisymmetric sudden expansion flow, 2D impinging jet flow, and axisymmetric impinging jet flow [
17]. In addition, critical parameters influencing numerical simulations—particularly bubble (droplet) diameter and surface tension—have not been systematically investigated. The present study specifically examines these interfacial phenomena.
Two distinct configurations of high-pressure water jets in air are simulated in this work. One is an impinging jet (normal impact configuration with a stand-off distance of 76D). The other is a free jet configuration without boundary interference. Benchmark comparisons were conducted using two industry-standard models: the realizable k-ε model and the SST k-ω model. The model formulations and implementation are briefly described in the next section. Subsequently, the results are presented and discussed in detail. Some conclusions are presented in the final section.
2. Model Formulation and Implementation
The high-pressure water jet in air is a typical gas–liquid multiphase flow. There are three conventional models that can be used for the numerical simulation of gas–liquid multiphase flows: the volume of fluid (VOF) model, the mixture model, and the Eulerian multiphase model.
The VOF model is suitable for problems where the gas and liquid phases do not penetrate each other and have clear interfaces. Conversely, the mixture model is appropriate for flows in which the gas and liquid mix well. Evidently, the problem of a high-pressure water jet in air does not fall into these two categories. The Eulerian multiphase model can solve complex multiphase flows and is a suitable choice for the problems we need to address in this work. When the Eulerian multiphase model is applied to isothermal, incompressible, viscous, and turbulent two-phase flows, its governing equations consist of the continuity equations, the momentum equations, and the turbulence transport equations.
The Eulerian multiphase model used in the ANSYS Fluent (17.0) CFD package is constructed based on the following assumptions:
All phases share a single pressure.
The continuity equations and the momentum equations are solved for each phase.
The turbulence model can be applied to the mixture or for each phase.
For the flow of high-pressure water jets in air, there is no mass exchange among the phases. Large velocities result in a high Reynolds number (Re) and Weber number (We), rendering the effects of surface tension and wall lubrication insignificant and thus negligible. Moreover, due to the relatively high Froude’s number, the gravitational force can also be excluded.
Based on the aforementioned assumptions and simplifications, the governing equations can be written as follows.
2.1. Continuity Equations
For phase
q, the continuity equation takes on the form
where
, are the volume fraction, density, and velocity of phase
q, respectively.
2.2. Momentum Equations
The momentum equation for phase
q can be written as
where
P is the pressure,
is the turbulent dispersion force, and
is an interaction force between phase
p and
q.
represents the stress–strain tensor of phase
q and can be written as
in which
and
are the shear and bulk viscosity of phase
q.
The interaction force between phase
p and
q is simply modeled as
where
represents the interphase exchange coefficient of momentum. For gas–liquid flow, it can be modeled as the general form:
where
and
are the density and diameter of the bubbles or droplets of the secondary phase,
p.
is the interfacial area.
f is the drag function.
is the particulate relaxation time, which is defined as
Many definitions for the drag function are adopted by different researchers. In this work, the Schiller and Naumann model is used. In this model, the drag function is
where
and
are the drag coefficient and relative Reynolds number, respectively. They are defined as
and
Several methods have been provided for calculating the interfacial area. In this work, the symmetric model is selected. The interfacial area is calculated as
2.3. Turbulence Dispersion Force
The turbulent dispersion force can be treated as an interfacial momentum force in the phase momentum equations (
in Equation (2)). Alternatively, its effect can be modeled by introducing a turbulent diffusion term into the phase continuity equations. The continuity equation for phase
q can be written as
where the turbulent dispersion term
must satisfy the constraint of
The diffusion coefficient is estimated from the turbulent viscosity as
In this work, the adjustable parameter = 1.0.
2.4. Turbulence Models
In ANSYS Fluent (17.0), three methodologies are available for the treatment of turbulence within the Eulerian multiphase model: the mixture turbulence model methodology, the dispersed turbulence model methodology, and the per-phase turbulence model methodology.
In the mixture turbulence model methodology, different substances are treated as a single mixture, and turbulence quantities—such as the turbulence kinetic energy, turbulence energy dissipation rate, and specific dissipation rate—are computed based on this mixture.
The density, molecular viscosity, and velocity vectors of the mixture are determined using the following equations:
For turbulence models, after substituting all quantities of single-phase flow with those of the mixture, the calculation methods and corresponding constants for each term in the turbulent transport equations for multiphase flow are completely consistent with those of single-phase flow. For instance, the turbulent transport equations for the SST
k-
ω-
φ-
α turbulence model, which is a self-developed model and has been demonstrated to have a strong capability to simulate jet flow and heat transfer, are presented below.
More details about the SST
k-
ω-
φ-
α turbulence model can be found in a study by Yang et al. [
17].
As the realizable k-ε turbulence model and the SST k-ω turbulence model have already been described in detail in the ANSYS Fluent (17.0) help documentation, they will not be elaborated upon here.
The dispersed turbulence model methodology is applicable when there is clearly one primary continuous phase and the rest are dispersed, dilute secondary phases. In the flow of high-pressure water jets in air, this condition is clearly not met.
In the per-phase turbulence model methodology, a set of turbulence transport equations are solved for each phase. When turbulence transfer among phases plays a dominant role, it may be an appropriate choice. However, for the flow of high-pressure water jets in air, our results showed that the per-phase turbulence model methodology is not feasible (see
Section 3.1.2). Additionally, due to the complex construction and lengthy description of this method, it will not be described in detail here. Those interested can refer to the ANSYS Fluent (17.0) help documentation.
2.5. Model Implementation
The ANSYS Fluent (17.0) CFD package was adopted as the computing platform. In the present work, three turbulence models, namely the SST
k-
ω-
φ-
α turbulence model, the SST
k-
ω turbulence model, and the realizable
k-
ε turbulence model, were used. The realizable
k-
ε model and the SST
k-
ω model are embedded models within ANSYS Fluent (17.0) and can be directly utilized. The user-defined scalar (UDS) functionality of ANSYS Fluent (17.0) was employed to implement the SST
k-
ω-
φ-
α model [
17].
The governing equation system, which comprises the multiphase form of continuity equations, momentum equations, and turbulence transportation equations, was solved using a pressure-based coupled algorithm. The convection terms in the momentum equations and turbulence transportation equations were discretized using the second-order upwind scheme. The gradients and derivatives were evaluated using the least-squares cell-based method.
Different wall treatment methods were employed for various turbulence models. Specifically, for the SST
k-
ω-
φ-
α model,
,
,
,
, and
were applied. The automatic near-wall treatment method was employed for the SST
k-
ω model [
18]. For the realizable
k-
ε model, the enhanced wall treatment method was utilized. At the inlet boundary of the nozzle, all quantities were assumed to be uniformly distributed. The mean velocity (
U) was set according to requirements. The turbulent quantities were calculated using
where
= 0.09,
I, and
VR are the turbulence intensity and the ratio of turbulent viscosity to molecular viscosity of the fluid at the nozzle inlet, respectively. In all simulations,
I = 0.05,
VR = 10,
φ = 0.3, and
α = 1.0 were adopted. At the outlets, the static pressure was set to zero.
4. Conclusions
Based on the Eulerian multiphase flow model, high-pressure water impinging jets in air and free high-pressure water jets in air were numerically simulated. The methods of turbulence treatment, the effects of different turbulence models, the influence of the bubble diameter of the secondary phase, and the effects of surface tension were thoroughly investigated. The results led to the following conclusions:
- (1)
The mixture turbulence model methodology not only has simple governing equations and is easy to use, but also can yield good results, making it suitable for simulating high-pressure water jets.
- (2)
For high-pressure water impinging jets in air, comparatively, the SST k-ω-φ-α model yielded the best results. Both the realizable k-ε model and the SST k-ω model slightly overpredicted the pressure near the stagnation point.
- (3)
The bubble diameter of the secondary phase has an influence on the impact pressure. The impact pressure decreases as the bubble diameter decreases. However, the bubble diameter has less effect on the impact pressure after it decreases below a threshold.
- (4)
Surface tension has a negligible effect on the impact pressure of high-pressure water impinging jets with a small stand-off distance.
- (5)
For free high-pressure water jets with long ranges, the difference between the SST k-ω-φ-α model and the SST k-ω model is not large. The results of the realizable k-ε model are quite different from those of the other two models, and they are not consistent with the trend of the experiments. Relatively, the results of the SST k-ω-φ-α model are better.
The main focus of this work is to investigate several key factors which affect the numerical simulation of high-pressure water jets in air. The obtained results can help to establish a dual-pronged framework that synergistically integrates computational mechanics with predictive modeling methodologies. From a theoretical perspective, the proposed model enables precise prediction of multiphase flow dynamics and material interaction mechanisms inherent to high-pressure water jet cleaning and cutting processes through combinging with rigorous solid mechanics computations. From a technological perspective, the rapid advancement of artificial intelligence (AI) and machine learning (ML) algorithms holds significant potential for revolutionizing operational parameter optimization in water jet applications. Specifically, this research can help to generate comprehensive datasets—encompassing fluid velocity fields, pressure distribution patterns, and volume of fraction characteristics—so it serves as a critical foundation for training robust ML architectures. This dual contribution can not only advance our fundamental understanding of fluid–structure interactions but can also support autonomous, data-driven decision-making systems in next-generation industrial water jet technologies.