The three-dimensional interaction and evolution of a thin rotating body's motion within a surrounding fluid is presented here. The motion of each is shown to affect the other significantly through a dynamic fluid–body interaction. The consideration of three-dimensional spatial effects and the time-dependent rotating motion of the body are new features for this near-wall unsteady problem. The non-linear fluid motion is formulated for an inviscid incompressible fluid and several scenarios are explored in which the body shape, body rotation, and body motion are studied. The problem reduces to solving Poisson's equation within the underbody planform, subject to mixed boundary conditions and to coupling with integral equations. To incorporate rotation and elliptical body shapes, the boundary conditions are rotated about the fixed mesh over each time step, producing a computationally efficient method. Through numerical and analytical investigations, stabilization of the body motion is shown with increased rotational frequency. Additionally, varying the body's ellipticity and center of mass further affects the body's stability and position in the fluid flow.

In a coupled fluid–body interaction, the motion of both a freely moving body and the fluid within which it is situated are considered to interact and mutually affect each other. Below, we investigate an extension to recent three-dimensional work by Smith and Liu,1 which considers a free, finite body located above a solid wall in an incident unidirectional fluid flow over the wall. An unsteady three-dimensional interaction occurs because the fluid flows between the wall and body and over the body in spanwise and streamwise directions. Here, we introduce the effect of body rotation, akin to the flight of a Frisbee.

The real-world motivation for studying such fluid–body interactions arises from industrial, biomedical, and environmental interests, such as aircraft icing,2–6 or environmental transport such as pollen motion7,8 and agricultural spray chemical application.9,10 Typically, such problems are studied through two-dimensional modeling and through a variety of settings and methods, such as direct numerical simulations in near wall flow,11,12 pipe/channel flow scenarios,13,14 dispersed particle/particulate flows,15–20 experimental endeavors,21,22 and in skipping and skimming interactions.23–28 

Within two spatial dimensions, several recent studies have furthered this area of research for high Reynolds numbers (as the scales of background motivations often dictate), through modeling and analysis for inviscid fluid–body interactions in boundary layers and channel flows.11,29–31 Overall, few studies to date have been performed or presented in three spatial dimensions, other than,1 despite the importance and relevance of three-dimensional effects in the practical motivations above. This sets the scope and necessity of the present study, which is focused on an applied mathematical investigation of a dynamic fluid–body interaction in three spatial dimensions.

Section II below describes the formulation of the model, including the scales, derivation of the unsteady three-dimensional system, and inclusion of body rotation. The model is then simplified with a linearized approach3 for an initial assessment of the parameter space and role of body rotation. Section III discusses the extension of the numerical method by Smith and Liu,1 which is applied here, while analytical and numerical cases are presented in Sec. IV with consideration of high-frequency body rotation. Section V provides further discussion and conclusions.

The freely moving three-dimensional body is submerged within a laminar, uniform, inviscid, incompressible fluid flow near a wall (or ground). The thickness of the underbody shape is here taken to be the same order as the thickness of the fluid-filled gap between the wall and the underbody but much less than the representative body lengths in the streamwise and spanwise directions. Our coordinate system moves horizontally with the body, as such the wall is vertically non-stationary in this frame of reference. Figure 1 provides a schematic of the scaled setup and indicates the main new features of the system modeled here, namely prescribed body rotation about the Z-axis and surface roughness, e.g., a small surface mounted mound.

FIG. 1.

(Top) A top-down view of a body with surface roughness, in the form of a small hump situated on the underbody (grey circle), in the (X,Y) plane with incoming stream U aligned with X. The body's leading (solid) and trailing (dotted) edges, X1(Y,T) and X2(Y,T), respectively, vary with both time and Y as the body moves near the wall and rotates. The leading and trailing edges meet when dY/dX=0 on the boundary of the body's planform. The angle of rotation about the Z-axis (yaw) is defined as Θ3=ωT, where ω is a prescribed angular frequency with which the body rotates (independent of the fluid interaction). The body roll, Θ2, denotes the rotation about the body X. (Bottom) A cross sectional view of the body in the (X,Z) plane. The incident fluid flow U aligns with the X-axis. Again, leading (solid) and trailing (dotted) edges, X1(Y,T) and X2(Y,T), respectively, are illustrated in the diagram. The body's pitch, angle of rotation about the body Y-axis, is denoted as Θ1 and the body roll, Θ2, is shown. The rotating and moving underbody shape is denoted F(X,Y,T) and includes an illustrative surface roughness (grey circle), here taking the form of a small mound.

FIG. 1.

(Top) A top-down view of a body with surface roughness, in the form of a small hump situated on the underbody (grey circle), in the (X,Y) plane with incoming stream U aligned with X. The body's leading (solid) and trailing (dotted) edges, X1(Y,T) and X2(Y,T), respectively, vary with both time and Y as the body moves near the wall and rotates. The leading and trailing edges meet when dY/dX=0 on the boundary of the body's planform. The angle of rotation about the Z-axis (yaw) is defined as Θ3=ωT, where ω is a prescribed angular frequency with which the body rotates (independent of the fluid interaction). The body roll, Θ2, denotes the rotation about the body X. (Bottom) A cross sectional view of the body in the (X,Z) plane. The incident fluid flow U aligns with the X-axis. Again, leading (solid) and trailing (dotted) edges, X1(Y,T) and X2(Y,T), respectively, are illustrated in the diagram. The body's pitch, angle of rotation about the body Y-axis, is denoted as Θ1 and the body roll, Θ2, is shown. The rotating and moving underbody shape is denoted F(X,Y,T) and includes an illustrative surface roughness (grey circle), here taking the form of a small mound.

Close modal

Below, the non-linear formulation for the fluid–body interaction in three dimensions is described, wherein the tilde sign denotes dimensional quantities. The body is assumed to be thin, lying close to a fixed solid wall given by z̃=0 as in Fig. 1, and the slopes of the body surfaces are considered to be suitably small. The surrounding fluid is regarded as inviscid and incompressible (the implications of which are discussed in the conclusions). Its oncoming flow relative to the body comprises a unidirectional uniform stream ũ=ũ in the x̃ direction, where (x̃,ỹ,z̃) are Cartesian coordinates with corresponding fluid velocity components being (ũ,ṽ,w̃). Time is denoted by t̃, the pressure is denoted by p̃ and is taken to be zero in the far field without loss of generality. Finally, ρ̃F and ρ̃B are the constant densities of the fluid and solid body, respectively.

Given the above, the Navier–Stokes equations, for an inviscid flow, apply throughout the fluid in which the body is submersed,
(1a)
coupled with the continuity equation given by
(1b)
Taking the streamwise and spanwise scales, x̃ and ỹ, to be comparable with the body length, say l̃, and the wall-normal scale, z̃, comparable with the body height, h̃, the body slopes are thus typically O(h̃/l̃). Note, the thinness of the body necessitates that h̃/l̃ is small. In this setting therefore, the coordinates in the thin gap between the body and wall and fluid flow parameters scale as follows:
(2a)
(2b)
(2c)
(2d)
with X, Y, Z of order unity. Here, (2b) follows from a balancing in the continuity Eq. (1b) and the streamwise and spanwise momentum equations of (1a). The pressure expansion, (2c), is also inferred from (1a); and the positive constant β in (2d) is large since the timescale of a dense body's motion is significantly greater than the surrounding fluid motion, i.e., for large ρ̃B/ρ̃F.
Following this setup and within the scaled setting, a three-dimensional non-linear thin-layer system governs the fluid motion in the thin gap and couples with the body motion via the fluid pressure on the underbody. This interaction depends upon the body's position in the fluid, Z=F(X,Y,T),
(3)
where H(T) is the vertical distance of the body's center of mass from the wall, Xc and Yc are the body's center of mass locations in X and Y, and Θ1(T),Θ2(T) are the scaled pitch and roll angles of rotation about the Y and X axes, respectively. A new aspect of this fluid–body interaction (compared to Ref. 1) is the inclusion of the underbody shape G(X,Y,T) variation with time T representing “Frisbee-like” rotation of the body about the Z axis, concerning the yaw angle of rotation Θ3. The effect of body rotation is of interest in the following analyses.
The fluid motion is governed by the following (scaled) continuity equation and two momentum equations in X, Y (by substituting (2a)–(2d) into (1a) and (1b)). Note that the scaled pressure P=P(X,Y,T) (of order unity) is independent of Z according to the normal momentum balance. Hence, the governing equations are
(4a)
(4b)
(4c)
These are subject to the following boundary conditions at the wall, on the underbody surface F, and time-varying leading edge and trailing edges of the body planform, X=X1(Y,T) and X=X2(Y,T) (due to body rotation about the Z-axis), respectively:
(5a)
(5b)
(5c)
(5d)
(5e)

In summary, (5a) ensures that the flow remains tangential at the solid wall, (5b) provides the kinematic condition at the moving underbody surface F for relatively slow scaled time T, and (5c) matches the oncoming flow in the far field. Thus, the fluid flow remains quasi-steady in the gap, in contrast with the motion of the body which, as described in (11) below, is fully unsteady with scaled time variable T of order unity.

As for the remaining conditions, at the leading edge of the planform, (5d) arises from a thin Euler-like region along the body's leading edge where σ is the angle between the tangent to the curved leading edge and the outer stream direction (X-axis), such that cot σ=X1(Y).32 Finally, constraint (5e) associates with a Kutta-like requirement that the pressure remains continuous across the trailing edge of the planform30,32 and that P is zero for any trailing edge location at the boundary of the planform. This is because the body's thinness induces only small variations in P except in the gap where it is of order unity. Further discussion of these two conditions is provided in Ref. 1.

Seeking solutions for U, V uniform in Z inside the gap, conditions (5a, 5b) establish that the velocity components must satisfy
(6)
while (4b), (4c) with (5c) imply zero vertical vorticity. Therefore, velocities can be written in terms of a scaled vertical potential function Φ(X,Y,T),
(7)
that is governed by the elliptic equation
(8)
with the planform-edge boundary conditions (5d), (5e) becoming
(9a)
(9b)
Finally, for the fluid pressure, (6) gives that
(10)
Overall, the fluid flow is thus controlled by (8)–(9) and couples with the rotating and moving body (of the prescribed underbody shape G(X,Y,T) (8)) by means of the fluid pressure (10).

Specifically, the body moves freely according to forces and moments acting on the underbody, which is a result of the fluid pressure upon the underbody. Hence, under the above scales, the only relevant forces/moments are given by (i) the scaled lift force LZ acting in the normal direction, (ii) the scaled moment MY acting about the Y-axis, and (iii) the scaled moment MX about the X-axis, with all other forces/torques negligible due to magnitude and timescale. Following from Newton's second law, the body force and moments equate to the area integral of the pressure over the underbody as follows:

(11)
(11a)
(11b)
(11c)
with m,j1,j2 the scaled mass and moments of inertia with respect to Θ1,Θ2. The above double integrals extend over the XY-planform of the body. Hence, the body's motion (11) couples to the fluid both via the fluid's motion and pressure (8)–(10) and via the constantly changing body derivative in (8) (due to the contribution of H in (3) and the body's rotation in the fluid).

To proceed, a linearized version of the problem is used as in Ref. 1. In this setting, consider the following scales:
(12a)
(12b)
The setting in (12a) is for notational convenience such that lowercase variables and parameters indicate working in the linearized setting for the remainder. For small ε (which is defined as the ratio of the typical body thickness to the typical gap width), the potential function Φ and pressure P expand as follows:
(13a)
(13b)

For the expansion of pressure in (13b), the leading order term of zero follows from (5c)–(5e) and the ε-order term arises due to the comparison of H,Θ1,Θ2 each with P in (11). For the expansion of ϕ in (13a), the leading order term here follows from the boundary condition (5c), with the ε-order term arising due to the comparison of P and Φ in (4a) and (4b).

Substitution into (8), (9a), (9b), and (10) provides the following system to leading order for the perturbation potential ϕ(x,y,t):

(14)
(14a)
(14b)
(14c)
(14d)

Finally, within this linearized setting, the underbody shape perturbation of (3) and body motion (11) become

(15)
(15a)
(15b)
(15c)
(15d)
integrating over the xy-planform of the body.

The governing equations and boundary conditions of the interaction are now thus (14a–15); initial conditions on h,θ1,θ2 and their temporal derivatives h,θ1,θ2 are assumed known at time zero. For prescribed underbody g(x,y,t) and planform shape, undergoing a constant rotation about the z-axis, θ3(t)=ωt, the above constitutes a closed problem for ϕ(x,y,t), with the locations of the inflow and outflow edges (leading and trailing edges) taken as known in advance for the current linearized setting.

Two observations are particularly worth mentioning: (i) The body's roll angle θ2 in our linearized model has no influence on the fluid flow and pressure due to the dominance of the horizontal scale as seen in (14a), (15a), and (ii) the body's vertical displacement h(t) in (15a) does not affect the pitch angle θ1(t) for the same reason. Of interest now are the effects of body rotation about the z-axis, and how such motion for a non-radially symmetric underwater shape can influence the coupled fluid–body interactions. Added complexity arises in the numerical scheme since the leading and trailing edges of a non-circular body planform also vary in time.

Having formulated the equations of motion for the fluid–body interaction, (14) and (15), we now present a method for computationally solving them. In this instance, a finite difference approach will be discussed that uses a fixed-mesh framework to reduce the computational complexity.

We present an extension of both the problem and the methodology formulated in Smith and Liu1 to include rotation about the z-axis of a body with an elliptical projected planform. The fluid flow in the gap between the body and the wall is of interest, especially for a non-smooth, non-rotationally symmetric underbody surface function (15a). If the body were radially symmetric about the z-axis, the rotation of the body would not affect the fluid–body motion significantly since the x-derivative of f would remain constant in time as the body rotates.

The problem to solve is the Poisson Eq. (14a) with mixed boundary conditions (14b), (14c) prescribed on the evolving leading and trailing edges, x1(y,t) and x2(y,t), of a rotating body. At each time step, the finite difference discretization of this system is encoded as the coefficients of a sparse linear equation matrix, see Smith and Liu1 for the discretization scheme. However, the rotating body introduces programming complexity.

In a frame of reference with a fixed flow direction, the sparse finite difference matrix formulated for this problem needs to be updated at every time step due to the body rotating. This process can become computationally inefficient, depending on the implementation, since it requires a rewriting of a sparse matrix at each time step. As the size of the finite difference matrix increases (e.g., smaller spatial discretization), the computational costs become significant.

Consider a frame of reference fixed on the body such that the flow direction rotates relative to the body. In this instance, a finite difference matrix that is fixed over time (discussed below) can be derived, thereby circumventing the evolving flow domain challenge. To proceed in this fixed body setting, the (i) coordinate system of the flow and (ii) relative spatial derivatives need careful consideration, as does (iii) the application of boundary conditions at the body leading and trailing edges (x1(t) and x2(t)).

Translating the model into elliptical coordinates, we take the flow domain to be centered at the origin and conform to the (unrotated) body's elliptical planform,
(16)
such that the computational mesh can be produced in elliptical coordinates with x̂=rcos(θ) and ŷ=qrsin(θ), for r(0,e] and θ(0,2π]. In the above, q is the constant ratio of lengths between the ellipse's two axes that describes the body's shape. While q and e are free parameters in general, throughout the following analyses we take e to be 1/q to maintain an overall planform area of π (without loss of generality).
At any time t, the x,y flow directions relative to a body rotating at a rate ω will rotate by an angle ψ=ωt such that
(17)
for which (16) continues to hold. The spatial derivatives with respect to x and y, and therefore with respect to r and θ, need careful consideration in order to solve the equations of motion (14). In the current setting, the required derivatives with respect to the flow direction are given by
(18a)
(18b)
(18c)
(18d)

Therefore, the derivatives with respect to x in the equations of motion (14a), (14c), (14d) take the body rotation into account.

Under the above considerations, the Poisson Eq. (14a) becomes
(19)
the left-hand side of which is independent of ψ (remaining the same as that of a non-rotating body), while the right-hand side incorporates the rotated x-derivative that varies with time, as per (18a) and (18c). A consequence of this is that the finite differencing formulated by Smith and Liu1 can be directly implemented here. In practice, the body rotation requires that the vector of constant terms b of the linear system is updated at each time step (due to the time evolving body derivatives on the right-hand side), which is more computationally efficient than updating the finite difference matrix.

Furthermore, given the fixed-body frame of reference, the boundary conditions now rotate about the body boundary where previously they had remained fixed in time for the non-rotating case. Two conditions are applied to the body at the trailing and leading edges, x1(t) and x2(t), respectively. These two edges meet at the two points on the boundary where dy/dx=0, i.e., at the body tangents that align with the flow direction. For the computational application of the boundary conditions, we only require the polar coordinate of these interface points, denoted β1,β2[0,2π). At the trailing edge, a Dirichlet boundary condition (14b) is applied when θ(β1,β2). In contrast, at the leading edge, a Neumann boundary condition (14c) is applied over the rest of the boundary curve, where θ[β2,β1]. Note here the dependence again on the body rotation due to the definition of θ/x by (18c) in the current setting. Once again, practically, the above adaptation amounts to updating the vector of constants b in our linear system. As a note of computational caution, when the body rotates, β1 and β2 may cross the point of θ=0 requiring careful implementation to ensure the conditions are applied correctly.

The computational domain and finite difference equations can now be formulated as in Smith and Liu,1 taking care to introduce the new discretization coefficients for the directional body derivatives. The domain is discretized into an M×N elliptical grid with Δr=1/M and Δθ=2π/N, where ri=iΔr, θj=jΔθ and i[1,M], j[1,N], thus excluding the origin. At each time step, the equations of motion are solved as a linear system using an (M×N)×(M×N) finite difference matrix to model the body trajectory over time. To emphasize, the novel extension herein lies in the modeling of a freely pitching and rotating elliptical body without the need for a moving mesh or grid interpolation.

In the proceeding analysis, we (i) consider the effect of body rotation and ellipticity on the fluid–body interaction, (ii) study the influence that moving the body's center of mass has on stabilizing or destabilizing the interactions, and (iii) analytically consider the behavior of the body and fluid in a high-frequency regime.

Throughout the proceeding analysis, the same example of underbody shape function is considered. The body planform may be either a unit disk or elliptical, depending on the prescribed ratio of the body's minor and major axes, q (following from (16)). The underbody is largely flat apart from a surface roughness in the form of a parabolic-shaped hump of circular base, radius c>0, sited off-center at some point x=a/q and y=bq. Hence, the body shape is given by
(20)

The values of the underbody shape constants are taken to be a,b,c=0.8,0.1,0.1 producing a hump of height c2=0.01. Note the scaling of a and b in (20) with q to ensure the hump's base remains within the planform boundary.

Furthermore, a grid independence study was carried out ahead of obtaining the presented results to ensure an accurate mesh resolution was used. The following results are performed on a 400×400 grid, uniformly discretizing the polar domain of r(0,e] and θ(0,2π].

Beginning now with the case of a circular planform, q=1, we investigate the influence of rotation about the z-axis on the body's motion, in particular with regard to its height, h(t), in the flow and pitch, θ1(t). Figure 2 shows the evolution of these two quantities over time for different values of angular velocity ω=0,π/2, π,2π,4π,10π. In both top (h) and bottom (θ) panels, the respective values can be seen to tend to a limit with increasing ω. The insets in each show similar values for 2π,4π,10π. The results show increased stability in the body motion with delayed blow up/fly away in the motion as shown by a reduced height and pitch at time t=10. The existence of a limit here indicates that the gyroscopic stability provided by rotation in the z-axis is of a finite influence with increased ω. Notably, there is a small correction for higher frequencies with the solution for ω=4π slightly “overshooting” that of ω=10π. A dotted black line in Fig. 2 shows the formal limit, the existence and nature of which are discussed in Sec. IV B, which also shows a slight correction back from the highest frequency solutions.

FIG. 2.

The motion of a body of circular planform rotating with angular velocities ω=0,π/2,π,2π,4π,10π, according to its height from the wall (top plot) and the body's pitch (bottom plot). The dashed line indicates the high-frequency limiting case for an averaged body function gav(r) as discussed in Sec. IV B.

FIG. 2.

The motion of a body of circular planform rotating with angular velocities ω=0,π/2,π,2π,4π,10π, according to its height from the wall (top plot) and the body's pitch (bottom plot). The dashed line indicates the high-frequency limiting case for an averaged body function gav(r) as discussed in Sec. IV B.

Close modal

Investigating further, Fig. 3 displays the effect of body rotation on both the lift force acting on the body (green curve, left axis) and the moment (purple curve, right axis) at time t=10. With increased ω, the forces acting over the body reach a limit around ω=2π, demonstrating why little further variation in the body motion is seen for larger angular velocities since the body forces remain relatively similar.

FIG. 3.

The lift (green, left axis) and moment (purple, right axis) acting on a rotating body with circular planform at t=10 due to the fluid–body interaction. The x-axis corresponds to values of the angular velocity with ω=0,π/2,π,2π,4π,10π.

FIG. 3.

The lift (green, left axis) and moment (purple, right axis) acting on a rotating body with circular planform at t=10 due to the fluid–body interaction. The x-axis corresponds to values of the angular velocity with ω=0,π/2,π,2π,4π,10π.

Close modal

A final point of interest in the circular planform case is the fluid pressure (p), and horizontal and vertical velocity components (u and v respectively) over the underbody. Figure 4 presents the results for these three quantities in turn at different radii on the body r=0.33,0.66,0.99. The top plot shows the pressure, while the bottom left and right show u and v, respectively. Interestingly, there is a scaling factor of 1/7 between the ω=0 (black curves, left axes) case and the ω=10π (purple curves, right axis) case causing the profiles for each case and radii to match exactly.

FIG. 4.

The pressure (p), horizontal velocity (u), and vertical velocity (v) on the underbody of a rotating body with circular planform at t=10. Each variable is shown at three radial distances from the body center, r=0.33,0.66,0.99, solid, dashed, and dotted curves, respectively, for polar co-ordinates θ=[0,2π). Black curves (left axis values) correspond to the non-rotating body and the purple curves (right axis values) correspond to ω=10π. The close alignment when overlaid reveals an exact scaling by a factor of 1/7 with rotation.

FIG. 4.

The pressure (p), horizontal velocity (u), and vertical velocity (v) on the underbody of a rotating body with circular planform at t=10. Each variable is shown at three radial distances from the body center, r=0.33,0.66,0.99, solid, dashed, and dotted curves, respectively, for polar co-ordinates θ=[0,2π). Black curves (left axis values) correspond to the non-rotating body and the purple curves (right axis values) correspond to ω=10π. The close alignment when overlaid reveals an exact scaling by a factor of 1/7 with rotation.

Close modal
The results above indicate that for large angular velocities ω, the trajectory of the body tends to some fixed solution, independent of ω. That is, at high rotational speeds, the body behaves as if it has an averaged shape, thus reducing the impact of local irregularities on fluid flow (such as the presence of the hump). To see this, given (20), the governing Eqs. (14)–(14d) reduce to
(21)
For ω=0, the body is not rotating and the hump remains of fixed location relative to the body center. Hence, the right-hand side of (21) becomes
indicating that the hump location stays fixed, centered at (x,y)=(a,b).
For high frequencies, with ω1, first consider the forcing terms acos(ωt)+bsin(ωt) in (21), which occurs due to rotation of the body. The complementary function for this system must have the form ϕ(x,y,t)=ϕ*(x,y)(acos(ωt)+bsin(ωt)), thus giving a corresponding p=p*(x,y)(acos(ωt)+bsin(ωt)) from (14d). Hence, continuing with the complementary solution (15c) takes the form
(22)
The above approximation assumes that the θ1 term in (21) contributes little to p* for large ω, which is justified later. For now, the solution to (22) is
(23)
Due to the presence of the ω2 term in the denominator of (23), the terms acos(ωt)+bsin(ωt) contribute negligibly to θ1(t) when ω is large. Hence, the right-hand side of (21) is in effect replaced by
for high angular velocity and frequency, indicating that the solution approaches that for a new non-rotating hump centered at x=0 for the complementary function.
Overall, this result points to the existence of the limit approached in the above presented cases. Indeed, the limit itself should be determined. For a rapidly rotating body, the fluid–body no longer responds as an off-center hump but rather that of the average body, gav, generally a non-rotating body with a ring. To formulate the average shape, the underbody form is integrated over θ for all r values such that
(24)

Hence, the average body only varies with r, taking the same cross-sectional form for all θ. Solving the equations of motion (14) with this new body function and ω=0, we obtain the limiting curves for h and θ in Fig. 2 that closely match the high angular velocity computational cases. The three-dimensional profile of this shape and consequent fluid pressure over its surface are presented in Fig. 5 wherein the results again match very closely across the whole body shape between the averaged body and rotating body cases.

FIG. 5.

A comparison of surface pressure over the underbody of an averaged body gav (top) and an underbody with a surface mounted hump g rotating at a high-frequency ω=10π (bottom), the solution for each is displayed at time t=10. The averaged body takes the form of ringed hump about the underbody with a height equal to the rotationally averaged height of the hump. The color gradient depicts the pressure values across the body, matching very closely at every x, y point.

FIG. 5.

A comparison of surface pressure over the underbody of an averaged body gav (top) and an underbody with a surface mounted hump g rotating at a high-frequency ω=10π (bottom), the solution for each is displayed at time t=10. The averaged body takes the form of ringed hump about the underbody with a height equal to the rotationally averaged height of the hump. The color gradient depicts the pressure values across the body, matching very closely at every x, y point.

Close modal

For a final comparison, the fluid pressure (p) and horizontal and vertical velocity components (u and v respectively) over the underbody at different radii on the body r=0.33,0.66,0.99 are presented in Fig. 6, again showing good agreement between the limiting averaged body case and the rotating body.

FIG. 6.

Comparison of a body of circular planform rotating with angular velocity ω=10π (red) to a non-rotating averaged body gav (black). The values of underbody pressure (p), horizontal velocity (u), and vertical velocity (v) are presented at t=10 for each body shape at three radial distances from the body center, r=0.33,0.66,0.99, solid, dashed, and dotted curves, respectively.

FIG. 6.

Comparison of a body of circular planform rotating with angular velocity ω=10π (red) to a non-rotating averaged body gav (black). The values of underbody pressure (p), horizontal velocity (u), and vertical velocity (v) are presented at t=10 for each body shape at three radial distances from the body center, r=0.33,0.66,0.99, solid, dashed, and dotted curves, respectively.

Close modal

In (15a) and (15c), the location of the body's center of mass plays a role in determining the evolution of its pitch within the flow and its roll (however, the latter is negligible in the overall body motion). Again for a circular planform, we compare the influence of xc on a non-rotating body to one rotating with ω=10π (left column and right column). Figure 7 shows the evolution of the body height h and pitch θ1 over time. For the non-rotating body, xc>0 can have a strong destabilizing effect blowing up the solutions for h(t) and θ1(t) with the magnitudes of both h and θ1 increasing faster as xc approaches 1. Alternatively, a stabilizing effect is seen for xc<0, with h and θ1 remaining small.

FIG. 7.

The motion of a body of circular planform rotating with angular velocities ω=0 (left column) and ω=10π (right column), as given by its height from the wall (top plots) and its pitch θ1, (bottom plots). The body's centre of mass, xc, is varied taking values of 1,2/3,1/3,0,1/3,2/3,1.

FIG. 7.

The motion of a body of circular planform rotating with angular velocities ω=0 (left column) and ω=10π (right column), as given by its height from the wall (top plots) and its pitch θ1, (bottom plots). The body's centre of mass, xc, is varied taking values of 1,2/3,1/3,0,1/3,2/3,1.

Close modal

The opposite trend is seen for a body that rotates, although the magnitudes of h and θ1 are several orders smaller. Indeed, for xc=1 a curious interaction occurs in which the body trajectory reverses, decreasing h and increasing θ1 over the t=[0,10] interval. Overall, rotation serves to provide stability for each value of xc.

Similar trends are seen in the pressure under the body in Fig. 8 where p at a radius r=0.99 is shown for both the non-rotating and high-frequency rotation cases (top and bottom, respectively) at t=10. Indeed, the magnitude and size ordering of the pressure curves follow that of Fig. 9, with greatly diminished pressures for xc<0 and when rotating. The link between the trends in p, h, and θ1 comes from the body motion Eq. (15), Eq. (15c) wherein it is the pressure over the underbody that governs the overall motion and coupled interaction.

FIG. 8.

The pressure (p) on the underbody of a rotating body with circular planform at t=10 at radial distance r=0.99 from the body center. The body rotates with angular velocities ω=0 (top) and ω=10π (bottom). Non-central center of mass, xc=[1,2/3,1/3,0,1/3,2/3,1].

FIG. 8.

The pressure (p) on the underbody of a rotating body with circular planform at t=10 at radial distance r=0.99 from the body center. The body rotates with angular velocities ω=0 (top) and ω=10π (bottom). Non-central center of mass, xc=[1,2/3,1/3,0,1/3,2/3,1].

Close modal
FIG. 9.

Plot showing the limiting cases for a body of circular planform rotating at high frequency, ω=10π. There is an interaction between the location of the hump and the center of mass, with similar limiting cases for (i) a=0.8,b=0.1,xc=1 and a=0.8,b=0.1,xc=1, and (ii) a=0.8,b=0.1,xc=1 and a=0.8,b=0.1,xc=1. Left: body height. Right: body pitch.

FIG. 9.

Plot showing the limiting cases for a body of circular planform rotating at high frequency, ω=10π. There is an interaction between the location of the hump and the center of mass, with similar limiting cases for (i) a=0.8,b=0.1,xc=1 and a=0.8,b=0.1,xc=1, and (ii) a=0.8,b=0.1,xc=1 and a=0.8,b=0.1,xc=1. Left: body height. Right: body pitch.

Close modal

One point of further interest is how the hump location (a, b, c) and xc interact. The above results relate to a body with a=0.8,b=0.1,c=0.1: see (20). Producing the same analysis for a=0.8,b=0.1, as in Fig. 9 it can be seen that the aforementioned trends reverse. That is, when the hump and center of mass are located on the same side of the body (i.e., a and xc are of the same sign), they serve to destabilize the body in comparison to when they are on opposite sides.

The final case we consider is that of bodies with varying elliptical planforms and the effect of rotation on their motion. In Fig. 10, the variations in body height and pitch over time are shown for different elliptical bodies. Once again, the left column depicts non-rotating bodies and the right column shows bodies rotating with velocity ω=10π. As a reminder, the parameter q is the ratio of the body's x-axis to the body's y-axis, with circular bodies, q=1, depicted for comparison (yellow curves). In each variation of q, the position of the hump location is scaled accordingly with a=0.8/q and b=0.8q to ensure that it remains within the planform.

FIG. 10.

The motion of a body of elliptical planform rotating with angular velocities ω=0 (left column) and ω=10 (right column), as given by its height from the wall (top plots) and its pitch θ1 (bottom plots). The ellipticity of the body, e, is varied taking values of 0.4,0.7,1,1.3,and1.6.

FIG. 10.

The motion of a body of elliptical planform rotating with angular velocities ω=0 (left column) and ω=10 (right column), as given by its height from the wall (top plots) and its pitch θ1 (bottom plots). The ellipticity of the body, e, is varied taking values of 0.4,0.7,1,1.3,and1.6.

Close modal

For the non-rotating case, q>1, the body's y-axis is larger than the x-axis, hence the body is wide in the spanwise direction and thin in the streamwise direction. Treating the circular body q=1 as a reference case, for q>1 the body is seen to interact more strongly with the flow, moving away from the wall (h) and pitching at a faster rate. The opposite is seen for q<1 where the body is spanwise thin and streamwise thick. The body trajectory, according to h, is now much shallower throughout and evolves more slowly.

Under rotation, two properties of interest are seen. First, q=1, the circular case is now the fastest growing solution, however the cases for other q values remain of similar order. Second, the cases for q=0.4 and 1.6 produce similar body trajectories, as do q=0.7 and 1.3 since the thicknesses of the body in spanwise and streamwise directions continually vary becoming periodically thick and thin. Thus, these pairings of q values present similarly sized and averaged body shaped bodies when rotating. The similarity in these results is further emphasized by Fig. 11 in which the body height (left axis, green) and pitch (right axis, purple) are shown for different q at t=10. For the non-rotating case, the circular body sits between the results for q<1 and q>1, with the magnitude of the lift force and moment both increasing with q. Under rotation, the circular case is seen to be the peak value with rough symmetry either side for high and low q. We thus anticipate that if q=q̂, where q̂ is some real scalar, under rotation the body trajectory results would match the case where q=1/q̂.

FIG. 11.

The lift (green, left axis) and moment (purple, right axis) acting on a rotating body with circular planform at t=10 due to the fluid–body interaction. The x-axis corresponds to values of ellipticity with q=0.4,0.7,1,1.3,1.6. The legend indicates the relevant line style, solid and dashed, of the considered angular rotations, ω=0,10π, respectively.

FIG. 11.

The lift (green, left axis) and moment (purple, right axis) acting on a rotating body with circular planform at t=10 due to the fluid–body interaction. The x-axis corresponds to values of ellipticity with q=0.4,0.7,1,1.3,1.6. The legend indicates the relevant line style, solid and dashed, of the considered angular rotations, ω=0,10π, respectively.

Close modal

In this paper, we have studied the three-dimensional fluid–body interaction of a freely moving, submerged body in a near-wall flow. Building upon the work of Smith and Liu,1 we extended their computational framework to include a rotating body in order to study the effect of a body's rotation about its vertical axis, perpendicular to the wall, on its motion over time. To this end, we used a fixed-frame finite difference method to efficiently solve the equations of motion and produce novel studies into how a body's rotation, its ellipticity and its center of mass location affect its motion. We demonstrated previously unseen (within this context) stabilization or destabilization of the body over time. The inclusion of a non-smooth underbody added further novelty to the results and study. In this setting, and through a series of different computational analyses, three key findings have emerged.

First, the high-frequency rotation of a circular body stabilizes its motion through the fluid (i.e., delays and slows its fly away and pitching) toward a finite limit, that of a non-rotating rotationally averaged underbody. As such, the high-frequency solution is independent of the angular velocity above a certain threshold. From Sec. IV B and Figs. 2, 6, and 5 it is apparent that this gyroscopic effect comes into play for large ω essentially scaling the body variables found for ω=0 by a factor 1/7 (for our prescribed g). For a rotationally averaged body shape gav (that does not rotate, i.e., ω=0), this gyroscopic stabilization is built into the underbody shape. The reasoning behind this phenomenon can be observed in (14a) due to the dependence on f/x on the right-hand side. Since the contribution of f/x depends on the body configuration and alignment in the flow, when the timescale of body rotation is fast compared to the evolution of its height and pitch (i.e., large ω), the fluid essentially only “sees” the rotationally averaged body and its averaged derivative. Finally, we note that the plateau of the stabilization effect at high rotational frequencies may well be due simply to the signs of the small perturbations about the limit.

Second, the body motion is greatly affected by the center of mass location, leading to a blow up in the body height and pitch of several orders for non-rotating bodies. Under rotation, the center of mass and underbody hump show further interaction such that when located on the same side the body motion is greatly diminished, even showing a slight reversal in direction (decreasing h and increasing θ), compared to the case when both hump and center of mass are on opposite sides causing a mild destabilization (in comparison to a centrally located center of mass). Indeed, throughout all of the analyses in this paper the body is shown to fly away, i.e., move away from the wall, in almost all cases, except for one scenario. When the body is rotating and the center of mass is ahead of the hump, xc>a>0, the body moves toward to wall. This is highlighted in Fig. 9 wherein the mirrored scenario (xc<a<0) is shown to match and show the same destabilization. In these instances, the center of mass and hump interact to “over-stabilize” the body, reversing its typical trajectory. When the body moves closer to wall, we anticipate that viscous effects due to the non-slip wall condition will emerge, requiring further analysis as to whether impact ensues, as considered in several two-dimensional analyses.12,25,29,33–35

Third, varying the ellipticity of the body leads to further changes in the body trajectory. For bodies of greater spanwise than streamwise thickness, the magnitude of both h and θ1 increases faster than in the opposite case. Hence, for a non-rotating body, the streamwise and spanwise presentation (shape) of the body to the fluid affect the time until the body flies way. When q is small, its streamwise dimension is greater than its spanwise dimension and serves to stabilize the body in the flow; while, for large q the opposite occurs. When the body rotates, the differences in the body's relative spanwise and streamwise extent average out and all body shapes remain similarly stable in the flow. This is further seen since, under rotation, results for q=q̂, where q̂ is some real scalar, showing agreement with the case where q=1/q̂ (up to the location of the hump), e.g., on average the fluid “sees” the same body shape for small and large q.

Thus, overall, it is seen that (1) body rotation has a generally favorable effect causing a body to stabilize and thus maintain proximity to a wall for longer through gyroscopic effects, (2) for elliptical bodies, the eccentricity of the body has very limited effect on the body trajectory when rotating, especially in comparison to the non-rotating case in which more significant fly away and blow up motions are seen, and (3) the location of the center of mass quantitatively and qualitatively affects the body trajectory, with body rotation reversing the trends seen in the non-rotating case for different values of xc. The novelty of this study in including these rotational body dynamics and its present three-dimensional setting enable the evaluation of the above phenomena.

There are several limitations to note with this modeling framework. First, the assumption of inviscid flow combined with a thin attached viscous three-dimensional boundary layer on the solid surface might limit the practical applications of this work. Since this study is an initial attempt at modeling such three-dimensional interactions, this assumption is reasonable to enable theoretical progress.

Second, we have assumed the flow remains attached throughout the interaction. The relevance of this assumption depends on the nature of the boundary layer. If the viscous boundary layer is predominantly turbulent, then the attached flow assumption is largely valid at high Reynolds numbers. If the boundary layer is predominantly laminar, then attached flow usually still applies for a finite scaled time before detachment and large-scale separation take place. These two points require further investigation and form potentially fruitful directions for future work. Indeed, a separated inviscid flow model could be built upon the present attached version and in principle to describe the separated flow scenario once separation is fairly widespread.

Third, further regarding the applicability of this method to real-world scenarios, a comparison to experimental data or direct numerical simulations of this interaction would be valuable for validating the results; however, such data and simulations are currently unavailable to us.

Fourth, our computational method is fast and efficient for the current setting; however, it is likely to lack flexibility and applicability for other settings. Indeed, we have presented results for the linearized problem and as such our proposed method is likely to be unsuitable for more complex scenarios, e.g., the inclusion of viscous effects and where a Poisson equation no longer holds in the wall gap. For such scenarios, we anticipate that different numerical methods and frameworks will be required.

In light of these limitations we anticipate several fruitful directions for future work, which include (1) three-dimensional formulation and body rotation effects in different flow settings of fluid–body interactions, such as in boundary layers and channel flows, and different scalings of the body relative to the wall, (2) further numerical study of this setting using finite element methods and direct numerical simulation as well as comparable experimental results to validate the applicability of the developed theory, (3) a non-linear version of the interactive system would be of value wherein it is expected that the roll motion of the body becomes important, (4) extension of fluid–body skimming interactions, particularly,25,28,36 and (5) an exploration of this model's relevance to real-world applications. This includes its applicability to aerospace and automotive engineering, particularly the movement of ice particles, droplets, and debris within a vehicle's boundary layer; underwater robotics and the motion of bodies close to the seabed or boundary layer under ice floes; and ecological and environmental phenomena such as pollen and agrochemical transport in environmental flows. This last case may also feature multiphysical effects such as electrostatics37–39 that will serve as a novel extension of fluid–body interaction problems.

The authors have no conflicts to disclose.

Ryan A. Palmer: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Kevin Liu: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Frank T. Smith: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
F. T.
Smith
and
K.
Liu
, “
Three-dimensional evolution of body and fluid motion near a wall
,”
Theor. Comput. Fluid Dyn.
36
(
6
),
969
992
(
2022
).
2.
R. W.
Gent
,
N. P.
Dart
, and
J. T.
Cansdale
, “
Aircraft icing
,”
Philos. Trans. R. Soc. A
358
(
1776
),
2873
(
2000
).
3.
J.
Mason
,
W.
Strapp
, and
P.
Chow
, “
The ice particle threat to engines in flight
,” in
44th AIAA Aerospace Sciences Meeting and Exhibit
(
ARC
,
2006
), p.
206
.
4.
E.
Norde
,
Eulerian Method for Ice Crystal Icing in Turbofan Engines
(
University of Twente
,
2017
).
5.
R.
Palmer
,
I.
Roberts
,
R.
Moser
,
C.
Hatch
, and
F. T.
Smith
, “
Non-spherical particle trajectory modelling for ice crystal conditions
,” No. 2019‐01-1961 (
SAE International
,
2019
).
6.
R.
Purvis
and
F. T.
Smith
, “
Improving aircraft safety in icing conditions
,” in
UK Success Stories in Industrial Mathematics
(
Springer
,
Cham
,
2016
), pp.
145
151
.
7.
M.
Chamecki
,
C.
Meneveau
, and
M. B.
Parlange
, “
Large eddy simulation of pollen transport in the atmospheric boundary layer
,”
J. Aerosol Sci.
40
(
3
),
241
255
(
2009
).
8.
M.
Sofiev
,
J.
Belmonte
,
R.
Gehrig
,
R.
Izquierdo
,
M.
Smith
,
Å.
Dahl
, and
P.
Siljamo
, “
Airborne pollen transport
,” in
Allergenic Pollen: A Review Production, Release, Distribution Health Impacts
(
Springer Netherlands
,
2013
), pp.
127
159
.
9.
A. S.
Felsot
,
J. B.
Unsworth
,
J. B.
Linders
,
G.
Roberts
,
D.
Rautman
,
C.
Harris
, and
E.
Carazo
, “
Agrochemical spray drift; assessment and mitigation—a review
,”
J. Environ. Sci. Health Part B
46
(
1
),
1
23
(
2010
).
10.
Y.
Gil
and
C.
Sinfort
, “
Emission of pesticides to the air during sprayer application: A bibliographic review
,”
Atmos. Environ.
39
(
28
),
5183
5193
(
2005
).
11.
E. M.
Jolley
,
R. A.
Palmer
, and
F. T.
Smith
, “
Particle movement in a boundary layer
,”
J. Eng. Math.
128
(
1
),
6
(
2021
).
12.
R. A.
Palmer
and
F. T.
Smith
, “
A body in nonlinear near-wall shear flow: Numerical results for a flat plate
,”
J. Fluid Mech.
915
,
A35
(
2021
).
13.
V.
Loisel
,
M.
Abbas
,
O.
Masbernat
, and
E.
Climent
, “
The effect of neutrally buoyant finite-size particles on channel flows in the laminar-turbulent transition regime
,”
Phys. Fluids
25
(
12
),
123304
123304
(
2013
).
14.
L.
Portela
,
P.
Cota
, and
R.
Oliemans
, “
Numerical study of the near-wall behaviour of particles in turbulent pipe flows
,”
Powder Technol.
125
,
149
157
(
2002
).
15.
M.
Dehghan
and
T. H.
Basirat
, “
Turbulence effects on the granular model of particle motion in a boundary layer flow
,”
Can. J. Chem. Eng.
92
,
189
(
2014
).
16.
P.
Diplas
,
C.
Dancey
,
A.
Celik
,
M.
Valyrakis
,
K.
Greer
, and
T.
Akar
, “
The role of impulse on the initiation of particle movement under turbulent flow conditions
,”
Science
322
,
717
720
(
2008
).
17.
A.
Kudrolli
,
G.
Lumay
,
D.
Volfson
, and
L. S.
Tsimring
, “
Swarming and swirling in self-propelled polar granular rods
,”
Phys. Rev. Lett.
100
,
058001
(
2008
).
18.
M. C.
Miller
,
IN.
McCave
, and
P. D.
Komar
, “
Threshold of sediment motion under unidirectional currents
,”
Sedimentology
24
(
4
),
507
527
(
1977
).
19.
C.
Schmidt
and
T.
Young
, “
The Impact of freely suspended particles on laminar boundary layers
,” in
47th AIAA Aerospace Sciences Meeting including The New Horizons Forum and Aerospace Exposition
2009
.
20.
J.
Wang
and
E. K.
Levy
, “
Particle behavior in the turbulent boundary layer of a dilute gas-particle flow past a flat plate
,”
Exp. Therm. Fluid Sci.
30
(
5
),
473
483
(
2006
).
21.
M. S.
Aköz
and
M. S.
Kırkgöz
, “
Numerical and experimental analyses of the flow around a horizontal wall-mounted circular cylinder
,”
Trans. Can. Soc. Mech. Eng.
33
,
189
215
(
2009
).
22.
J.-M.
Foucaut
and
M.
Stanislas
, “
Experimental study of saltating particle trajectories
,”
Exp. Fluids
22
(
4
),
321
326
(
1997
).
23.
I. J.
Hewitt
,
N. J.
Balmforth
, and
J. N.
McElwaine
, “
Continual skipping on water
,”
J. Fluid Mech.
669
,
328
(
2011
).
24.
P. D.
Hicks
and
F. T.
Smith
, “
Skimming impacts and rebounds on shallow liquid layers
,”
Proc. R. Soc. A
467
(
2127
),
653
674
(
2011
).
25.
E. M.
Jolley
and
F. T.
Smith
, “
Interactions between a heavy particle, air, and a layer of liquid
,”
Phys. Fluids
35
(
4
),
043311
(
2023
).
26.
R. A.
Palmer
and
F. T.
Smith
, “
Skimming impacts and rebounds of smoothly shaped bodies on shallow liquid layers
,”
J. Eng. Math.
124
(
1
),
41
73
(
2020
).
27.
R. A.
Palmer
and
F. T.
Smith
, “
Skimming impact of a thin heavy body on a shallow liquid layer
,”
J. Fluid Mech.
940
,
A6
(
2022
).
28.
R. A.
Palmer
and
F. T.
Smith
, “
The role of body shape and mass in skimming on water
,”
Proc. R. Soc. A
479
(
2269
),
20220311
(
2023
).
29.
R. A.
Palmer
and
F. T.
Smith
, “
A body in nonlinear near-wall shear flow: Impacts, analysis and comparisons
,”
J. Fluid Mech.
904
,
A32
(
2020
).
30.
F. T.
Smith
and
A. S.
Ellis
, “
On interaction between falling bodies and the surrounding fluid
,”
Mathematika
56
,
140
(
2010
).
31.
F. T.
Smith
and
P. L.
Wilson
, “
Body-rock or lift-off in flow
,”
J. Fluid Mech.
735
,
91
(
2013
).
32.
M. A.
Jones
and
F. T.
Smith
, “
Fluid motion for car undertrays in ground effect
,”
J. Eng. Math.
45
(
3
),
309
334
(
2003
).
33.
E. M.
Jolley
and
F. T.
Smith
, “
A heavy body translating in a boundary layer:‘crash’,‘fly away’and ‘bouncing’ responses
,”
J. Fluid Mech.
936
,
A37
(
2022
).
34.
R. A.
Palmer
and
a F. T.
Smith
, “
When a small thin two-dimensional body enters a viscous wall layer
,”
Eur. J. Appl. Math.
31
(
6
),
1002
1028
(
2020
).
35.
F. T.
Smith
,
E. M.
Jolley
, and
R. A.
Palmer
, “
On modelling fluid/body interactions, impacts and lift-offs
,”
Acta Mech. Sin.
39
(
5
),
323019
(
2023
).
36.
K.
Liu
and
F. T.
Smith
, “
Collisions, rebounds and skimming
,”
Phil. Trans. R. Soc. A
372
(
2020
),
20130351
(
2014
).
37.
S. J.
Harris
,
R. A.
Palmer
, and
N. R.
McDonald
, “
Modelling floral and arthropod electrostatics using a two-domain AAA-least squares algorithm
,” arXiv:2411.02195 (
2024
).
38.
E. R.
Hunting
,
S. J.
England
,
K.
Koh
,
D. A.
Lawson
,
N. R.
Brun
, and
D.
Robert
, “
Synthetic fertilizers alter floral biophysical cues and bumblebee foraging behavior
,”
PNAS Nexus
1
(
5
),
1
6
(
2022
).
39.
E. R.
Hunting
,
J.
Matthews
,
P. F.
de Arróyabe Hernáez
,
S. J.
England
,
K.
Kourtidis
,
K.
Koh
,
K.
Nicoll
,
R. G.
Harrison
,
K.
Manser
,
C.
Price
et al, “
Challenges in coupling atmospheric electricity with biological systems
,”
Int. J. Biometeorol.
65
,
45
58
(
2021
).
Published open access through an agreement withJISC Collections