1. Introduction
In complex sea conditions, the environmental loads, such as wind, waves, and currents, inevitably lead to the roll and pitch motion responses of ships and floating platforms [
1,
2,
3], affecting the regular operation of equipment and causing safety risks to personnel. Therefore, it is necessary to choose appropriate anti-roll methods to ensure the motion stability of ships and platforms under load disturbances.
Currently, the widely used anti-roll devices mainly include bilge keels, fin stabilizers, anti-roll tanks, gyrostabilizers, and so on [
4,
5,
6,
7,
8]. Among these, gyrostabilizers employ the gyroscope theorem to obtain amplified precession output torque, effectively suppressing roll motion. Moreover, gyrostabilizers are known for their easy installation, small space requirements, and ability to achieve excellent anti-rolling performance even at zero speed [
9]. Therefore, gyrostabilizers have emerged as a promising and highly regarded solution for reducing roll, captivating the attention of both the international academic and industry communities [
10,
11,
12,
13,
14]. Based on the same principle, the Fluid Momentum Wheel (FMW) has special structural features that provide high safety and stability. Besides their anti-roll function, FMW can also be used for ballasting or adjusting the center of gravity, presenting extensive prospects for application. Arranging the FMW underwater offers advantages in terms of the release and replenishment of fluid medium and heat dissipation [
15]. The concept of FMW and its potential advantages in satellite and ship applications was first proposed by Maynard [
16], laying the foundation for the design of the FMW and providing a clear direction for further research. In 2004, the Texas University Fluid Loop Orientation/Attitude Test (FLOAT) team [
17] developed and tested a fluid momentum controller consisting of multiple water-filled loops driven by mechanical pumps and underwent ground and flight tests to assess the attitude-control capability. Extensive research has been conducted on the application of the FMW in the aerospace industry, specifically for satellite attitude control.
However, there has been limited exploration and systematic investigation of the application of FMW devices in ocean engineering. As a new type of anti-roll measure, the related research contents and key problems of the FMW are extensive. Du et al. [
18] designed a stable offshore platform based on toroidal flow with the principle of angular momentum, demonstrating its ability to resist roll and pitch motion responses. Additionally, Wang et al. [
19] analyzed the flow characteristics of the FMW through a PIV experiment and numerical simulation and confirmed its gyroscopic precession characteristics. These findings indicate that the enclosed volume of the FMW can be easily increased, which means that greater anti-roll performance can be achieved. Compared to traditional rigid gyrostabilizers, FMW utilizes a fluid medium to generate the angular momentum, and thus, the stability of the flow field may significantly affect fluid resistance and the anti-roll effect.
Under the combined effect of centrifugal force and a radial pressure gradient, curved pipes generate a pair of counter-rotating vortices within the plane known as Dean Vortices by Dean [
20], which results in the non-uniform distribution of streamwise fluid velocities across the cross-section. This phenomenon is believed to cause higher fluid resistance compared to straight pipes under similar flow velocities, based on the studies on flow phenomena and mechanical energy loss in various curved pipes [
21,
22,
23,
24]. White [
25] was one of the earliest researchers to conduct experimental studies on pressure loss in helical pipes, and the results were in good agreement with the earlier theoretical work by Dean [
20]. Ito [
26] conducted experimental research on pressure loss in toroidal pipes with flowing water, establishing a correlation between the friction factor and the Dean number. El-Genk and Schriener [
27] comprehensively summarized experimental data on pressure loss and convective heat transfer in toroidal and helical pipes, revealing that secondary flows can enhance convective heat transfer while simultaneously increasing pressure loss. Existing studies indicate that the loss factors in curved pipes rely on not only the Reynolds number but also geometric parameters (such as bend angle, radius of curvature, inlet-outlet area ratio, and so on).
In addition, scholars have also conducted some studies on the flow phenomena and instability of curved pipes [
24,
28,
29]. Sudo et al. [
30] conducted experimental studies on flow separation phenomena in a 90° bend and analyzed the variation in secondary flow in the downstream flow field through the physical quantity of the swirl intensity of secondary flow. The research by Kim et al. [
31] further revealed a strong correlation between the swirl intensity of secondary flow and the curvature radius of a 90° bend while showing a weak correlation with the Reynolds number. Recently, some researchers [
32,
33,
34,
35] employed the PIV measurement technique to study the flow field and turbulence structure in curved pipes. Among them, Ikarashi et al. [
32] investigated the influence of bend curvature (1.0, 1.2, 1.5) on the flow field of a 90° bend, specifically analyzing the phenomena of flow separation and secondary flow. It was observed that flow separation occurred on the inner wall near the bend exit, particularly at smaller radius ratios, increasing the velocity of the secondary flow and turbulent kinetic energy. Noorani et al. [
36,
37] employed the Direct Numerical Simulation (DNS) method to simulate fully developed flow in toroidal pipes with varying curvatures and Reynolds numbers, aiming to analyze the swirl-switching phenomenon of secondary flow. In another study, Wegt et al. [
38] employed the Large Eddy Simulation (LES) method to analyze the variations in the position of the vortex core and the distribution of streamwise velocity downstream of a 90° bend. He et al. [
39] used LES and the Proper Orthogonal Decomposition (POD) method to study the correlation between the secondary flow motion and wall shear stress in 90° bends at different Reynolds numbers (5300, 27,000, and 45,000). Until now, most studies on flow fields and energy losses in curved pipes focus on common configurations, such as 90° bends and U-shaped pipes with inlets and outlets. Due to the enclosed circular structure of the FMW, the relevant research data are limited. The secondary flow and streamwise velocity stratification inside the FMW induce changes in flow instability, resulting in differences in energy losses. A further quantitative evaluation of flow instability is necessary to explore its inherent relationship with energy losses.
The paper can be organized as follows.
Section 2 compares different turbulence models with DNS results and then analyzes the phenomena of in-plane secondary flow and the non-uniform distribution of streamwise velocities in the spatial, as well as temporal, pulsation characteristics in the FMW.
Section 3 establishes evaluation parameters to study the influence of typical parameters of the FMW (for example, pipe diameter, curvature radius, velocity) on the spatiotemporal instability of the flow field. Furthermore, dimensionless mechanical factors of the Reynolds number and curvature ratio are presented to analyze the characteristics of flow instability and quantify its influence on energy losses. Finally, concluding remarks and future perspectives are presented in
Section 4.
2. Numerical Methods and Validation
This study employed the STAR-CCM+ 2022.1.1 software based on the finite volume method to solve the governing equations with the second-order discrete scheme in both time integration and spatial discretization. The chosen time resolution ensured that the Courant number should be less than 1 throughout the entire computational domain, and the
y+ values can satisfy the requirements of the selected turbulence model. The research subject is a unique toroidal pipe that exhibits a closed geometric structure without explicit import/export boundaries. In order to simulate the fully developed flow, periodic boundary conditions were applied at the inlet and outlet with the source term of the pressure gradient in the streamwise direction to balance the fluid resistance, and no-slip conditions were imposed on the wall surfaces, as shown in
Figure 1a. The cylindrical coordinate system takes the geometric center of the toroidal pipe as the origin, and the circumferential direction (
θ) represents the streamwise direction. The mesh generation for the cross-section of a pipe was arranged with an “O-grid” to make the mesh distribution more uniform and the transition smoother. The circular cross-section is discretized by five blocks, with one block located in the center of the pipe and the other four blocks surrounding the pipe to achieve its circular shape, as shown in
Figure 1b. The coordinate system is established as a cylindrical coordinate system with the geometric center of the pipe cross-section as the origin, and the vertical direction (
s) represents the streamwise direction. The selection of mesh parameters for the pipe has been validated in the previous study.
The fluid medium investigated in this paper is water, and the turbulent state of the curved pipe is considered to be three-dimensional, incompressible, and viscous without heat exchange. In order to analyze the turbulent flow, each variable in the instantaneous Navier-Stokes equations is separated into its mean value and fluctuating value, which is given by
Φ =
φ +
φ’. Therein,
Φ can denote the velocity and pressure, and the fluctuating mass force is neglected. The governing equations are the unsteady three-dimensional, incompressible Reynolds-averaged Navier-Stokes (RANS) equations, which include the following mass and momentum conservation equations:
Compared to the traditional N-S equations, the Reynolds-averaged equation has an additional Reynolds stress term, which represents the influence of turbulent fluctuations on the time-averaged flow. In order to solve the Reynolds-averaged equations, numerical simulation methods can be divided into direct numerical simulation (DNS) and non-direct numerical simulation. DNS directly solves the instantaneous turbulent governing equations, while non-direct numerical simulation approximates and simplifies turbulent characteristics to solve the Reynolds stress, which can be further categorized into Large Eddy Simulation (LES) and Reynolds-Averaged Navier-Stokes (RANS) methods. Among them, DNS requires no simplifications or approximations for turbulent flow and can provide accurate results with errors mainly from numerical computations. LES assumes that turbulent fluctuations and mixing are primarily caused by large-scale vortices in the flow field, which derive energy from the mean flow and exhibit substantial anisotropy. Therefore, for large-scale vortex motions that can be captured by the grid, the fluctuation terms can be calculated directly by solving the governing equations. For small vortices that cannot be captured, the subgrid-scale model is employed to simulate the impact of small-scale vortex motions on large-scale motions, and the Dynamic Smagorinsky subgrid-scale model of LES was employed in this study. Furthermore, in order to solve the Reynolds-averaged equation, the RANS method makes some assumptions about the Reynolds stress, which involves establishing stress expressions or introducing new turbulence models that relate the Reynolds stress to the time-averaged quantities. The Reynolds Stress Transport (RST) model calculates the components of the Reynolds stress tensor by solving the transport equations. The pressure-strain term in the Reynolds stress transport equation was calculated by employing the Elliptic Blending (EB) model in this study. Moreover, according to the Boussinesq assumption, the Reynolds stress is a function of the turbulent eddy viscosity
μt and can be expressed as Equation (3).
where
δij is the Kronecker delta, and
sij is the average strain tensor. The standard
k-
ε model is one of the most fundamental two-equation models, which introduces the turbulent kinetic energy (
k) and the dissipation rate (
ε) equations, respectively. Additionally, there are other models, such as the standard
k-
ω model and the SST
k-
ω model. The SST
k-
ω model, proposed by Menter [
40], is a hybrid model that combines the advantages of the standard
k-
ε and k-ω models. The SST
k-
ω model is more suitable for studying the flow field in curved pipes due to its combination of the favorable properties of the
k-
ω model for low-Reynolds-number flow near the wall and the
k-
ε model for far-field calculations. The turbulent eddy viscosity in this model is defined by the following:
where
S is the modulus of the average strain tensor,
a* and
β* are model coefficients,
a1 is 0.31, and
d is the distance from the wall.
In order to validate the accuracy of the numerical method, the computed results obtained from different turbulence models were compared with DNS results. The models adopted the same parameters as DNS, including the curvature ratio (
r0/
R) of 0.1 and Reynolds number of 11,700 [
36]. The Fanning friction factor (
f) and the computational costs were taken into consideration for evaluating the analysis validity.
where
τθ denotes the streamwise component of the mean wall shear stress, an overbar denotes the average value along the circumference of the pipe section, and
ub is the bulk velocity.
From the results in
Table 1, it can be observed that the Realizable
k-
ε model demonstrates a larger discrepancy in predicting the friction factor (
f) compared to other turbulence models, which is consistent with the findings in the study of Noorani et al. [
36]. The study indicated that while the
k-
ε model can predict turbulence in straight pipes and mildly curved pipes, it exhibits significant deviations in cases with high curvature. Considering computational efficiency, the SST
k-
ω model is undoubtedly the best choice.
Assuming the rotor of the rigid-body gyrostabilizer is identical to the FMW, the rotational angular velocities of the rotor are the same. As depicted in
Figure 2a, the rotor velocity increases radially from the inside to the outside, with a normalized velocity range of 0.9~1.1.
Figure 2b,c present the transient flow fields of DNS and LES in the cross-section of the FMW. Due to the influence of fluid viscosity, the velocity at the wall is zero, and the range of velocity fluctuations expands to 0~1.5. Additionally, both velocities exhibit strong fluctuation characteristics due to the influence of turbulent eddies, which are generated near the wall and subsequently develop, affecting the flow field at the center. Consequently, this gives rise to the first unstable phenomenon of the FMW compared to the gyrostabilizer, namely the fluid velocity fluctuations.
The average flow field, which can show the spatial distribution characteristic, can be obtained by averaging time on the instantaneous fluid field.
Figure 3 compares the average flow fields in the DNS and LES cross-section with the RANS method results, including the RST model, the SST
k-
ω model, and the Realizable
k-
ε model. The normalized velocities in the plane (
) and in the streamwise direction (
uθ/
ub) represent the upper and lower sides of the figure, respectively. From the velocity distribution in the streamwise direction for DNS and other models, it can be observed that under the action of inertial centrifugal force, regions with higher velocities tend to be located towards the outer side, while lower velocities are found on the inner side, presenting the phenomenon of velocity stratification similar to that of the rigid-body gyrostabilizer. In the presence of an imbalance between inertial centrifugal force and the radial pressure gradient, the secondary flow phenomenon is observed in the plane of the FMW. The velocity profile on the upper plane clearly demonstrates that LES and SST
k-
ω models provide a closer approximation to the DNS results than other models. The in-plane velocities are primarily concentrated near the upper wall, while the counter-rotating vortices with an outward shift are observed near the inner side. In contrast, the backflow effect of the secondary flow near the inner wall is weak in the RST model, resulting in significant deviations from the DNS results. The Realizable
k-
ε model overestimates the influence of inertial centrifugal force, resulting in a greater outward shift and an expanded range of influence from the secondary flow. Therefore, regarding the velocity distribution in the flow field of the FMW, LES and the SST
k-
ω model demonstrate better similarity to the DNS results.
Turbulent kinetic energy is commonly used to quantify the intensity of velocity fluctuations and describe the temporal characteristics of the flow field.
Figure 4 presents the distribution of normalized turbulent kinetic energy (
k/
uτ2) for different turbulence models. The inertial centrifugal force causes high-velocity fluid particles to shift towards the outer side, leading to more pronounced turbulent fluctuations in that region. On the inner side, the swirl-switching of the secondary flow results in a concentrated turbulent kinetic energy shown as two vortices. Compared with the findings obtained from DNS, the LES model exhibits the closest agreement, followed by the SST
k-
ω model. The previous study found that the distribution of turbulent production terms in both the LES and SST
k-
ω models closely resembles the results obtained from DNS. The SST
k-
ω model, as a Reynolds-averaged method, encounters challenges in capturing the time-dependent swirl-switching phenomenon of the inner secondary flow, leading to lower values of turbulent kinetic energy on the inner side. Conversely, LES avoids this limitation and thus achieves the best predictive performance. However, the RST and Realizable
k-
ε models significantly overestimate the turbulence intensity in the inner region, resulting in higher values of predicted turbulent kinetic energy.
Although LES has higher accuracy compared to other models, its computational cost increases rapidly, with a larger Reynolds number. Therefore, the RANS method to solve the Reynolds-averaged equation is an efficient and reasonable approach in engineering applications. Therein, the RST and Realizable k-ε models exhibit significant errors in simulating secondary flows in the plane. The SST k-ω model combines the advantages of the k-ω model in calculating low-Reynolds-number flows near the wall and the k-ε model in far-field calculations, which can be better used to study the flow field information in curved pipes. Considering the computational precision and efficiency of various turbulence models, the SST k-ω model was chosen for the subsequent study of the flow instability in the FMW.
In order to verify the accuracy of the SST
k-
ω model, the numerical results of the driving pump with different rotational speeds are compared with the PIV experimental data [
19]. The experiment revealed that the flow field exhibited a non-uniform distribution of the streamwise velocity and spatial instability of the secondary flow in the cross-section. Here, the test case with 250 rpm is selected for the comparison validation, and the fluid velocity in the FMW is shown in
Figure 5. The mean velocities along the main flow direction obtained using the flowmeter, PIV, and numerical simulation are 0.78, 0.79, and 0.85 m/s, respectively. The maximum error is less than 9%, which verifies the reliability of the numerical method.
The velocity distributions in the radial direction of the FMW can be obtained using PIV and numerical simulation methods, as shown in
Figure 6. Therein, the
x-axis is the non-dimensional ratio between the distance to the center and the radius, and the
y-axis is the velocity value divided by the average velocity to obtain the dimensionless value of the velocity. As can be seen from the figure, the velocity distribution measured via PIV is almost consistent with the numerical solution. The velocity gradually increases from the inner region to the outer region, indicating the presence of clear velocity stratification. However, it should be noted that the relatively big tracer particles in the boundary layer can result in data loss using the PIV method. The experiment revealed that the flow field exhibited a non-uniform distribution of the streamwise velocity and spatial instability of the secondary flow in the cross-section.
4. Conclusions
The inertial centrifugal force and radial pressure gradient induce the in-plane secondary flow and non-uniform velocity distribution, significantly affecting the fluid resistance. Additionally, the temporal instability caused by turbulent fluctuation also affects the energy conversion of the flow field. Based on previous research on the energy loss mechanism in the FMW, this study used numerical simulation methods to analyze the effects of typical parameters on the flow field instability and mechanical energy loss. Some meaningful conclusions can be made as follows:
(1) This study conducted calculations using different turbulence models (LES, RST, SST k-ω, and Realizable k-ε models) and compared the results with DNS and experimental results. The differences in the streamwise and in-plane velocities and turbulent kinetic energy were further analyzed. Among the models, LES showed the best agreement in terms of spatial and temporal characteristics with high computational costs. However, the RST and Realizable k-ε models exhibited larger errors in the distributions of secondary flow and turbulent kinetic energy. The SST k-ω model provided high accuracy in both the Fanning friction factor and flow field characteristics while maintaining computational efficiency, and reproduced the velocity stratification phenomenon observed in experiments well. Therefore, the SST k-ω model was selected to analyze the influence of flow field instability on mechanical energy loss in the FMW.
(2) The flow characteristics of the FMW differ significantly from the rigid-body gyrostabilizer with a constant angular velocity due to the influence of fluid viscosity. Specifically, the flow field exhibits significant turbulent fluctuations with time and velocity stratification and in-plane secondary flow in the spatial characteristics. In the analysis of the spatiotemporal instability of the FMW flow field, the spatial instability parameters (angular velocity non-uniformity and swirl intensity of secondary flow) and temporal instability parameter (turbulent kinetic energy) were defined to assess the impact of typical design parameters on the flow field instability. The computational results demonstrated that the instability increases as the typical parameters (pipe diameter, radius of curvature, and velocity) increase. The in-plane secondary flow significantly affects the streamwise velocity distribution, transforming the profile of the high-velocity region on the outer side into a crescent shape and C-shape. Consequently, the angular velocity non-uniformity and the magnitude and distribution of turbulent kinetic energy change with it, ultimately influencing fluid resistance. Subsequently, the typical parameters were represented as dimensionless parameters (Reynolds number and curvature ratio). Under the same conditions of these dimensionless parameters, the flow field instability maintains good agreement, indicating that using dimensionless parameters to evaluate the flow field instability is reasonable enough.
(3) Finally, this study examined the impact of flow field instability on mechanical energy loss. The distribution of wall shear forces is closely related to the velocity distribution, showing a notable non-uniform distribution with higher values on the outer side compared to the inner side. The increase in typical parameters enhances the instability of the flow field, which leads to higher wall shear forces and increased resistance effects. Moreover, the flow field instability parameters are normalized, and the Fanning friction factor is employed to evaluate the energy loss characteristics of the flow field. Non-dimensional parameters are established to quantitatively evaluate the relationship between flow field instability and energy loss. The findings revealed that an increase in the Reynolds number corresponds to a decrease in the flow field instability and friction factor. Conversely, a rise in the curvature ratio amplifies the in-plane secondary motion of the flow field, resulting in an increased friction factor. When comparing their relative influences, it is evident that the Reynolds number has a significantly larger impact on the Fanning friction factor than the curvature ratio. Specifically, the curvature ratio has a stronger influence on the in-plane motion of the flow field, indirectly affecting the local Reynolds number and leading to changes in the friction factor.
This study analyzed the temporal fluctuation represented by turbulent kinetic energy to provide support for quantitatively evaluating the mechanical energy loss in the FMW. However, the RANS method employed in this study has limitations in capturing the details of temporal fluctuation characteristics in the flow field, such as the swirl-switching phenomenon of secondary flow in the main flow region. In the boundary layer near the wall, the stretching, deformation, and splitting of turbulent vortices affect the energy transfer and transformation. In future studies, LES or DNS methods will be used to investigate the near-wall turbulent structures and the relationship between the switching frequency of the secondary flow and the fluctuations in the wall shear force.