Fact-checked by Grok 2 weeks ago

Two-dimensional flow

Two-dimensional flow in describes the motion of a where the at every point is parallel to a fixed and remains uniform along directions perpendicular to that , effectively varying only in two spatial dimensions. This assumption simplifies analysis for scenarios where one dimension (such as width) is significantly larger than the others, allowing flow parameters like , , and to be treated as independent of the third coordinate. Key assumptions in two-dimensional flow models often include incompressibility (constant density), irrotationality (zero vorticity, enabling theory), and inviscidity (negligible viscous effects), which facilitate mathematical tractability. The field can be expressed using a \psi(x, y, t), where the components are v_x = \frac{\partial \psi}{\partial y} and v_y = -\frac{\partial \psi}{\partial x}, ensuring the flow is divergence-free for incompressible cases; for irrotational flows, \psi satisfies \nabla^2 \psi = 0. In , two-dimensional irrotational and incompressible flows are represented by analytic functions via the complex potential \Phi(z) = \phi + i\psi, where z = x + iy, linking to \Phi'(z). Applications of two-dimensional flow theory are prominent in , such as modeling airflow over airfoils or wings with high aspect ratios, and in hydrodynamics for analyzing flow around ship hulls or in channels. It also underpins computational simulations and experimental studies of boundary layers, vortices, and uniform flows combined with sources or sinks, providing foundational insights into more complex three-dimensional phenomena.

Fundamentals

Definition and assumptions

Two-dimensional flow in describes a type of motion in which the components vary only in two spatial directions, typically within a plane, while assuming uniformity or infinite extent in the third perpendicular direction, as seen in plane flow configurations. This simplification models scenarios where the flow is effectively confined between parallel planes or extends indefinitely without variation along the spanwise axis, allowing the at any point to be identical along lines normal to the flow plane. The analysis of two-dimensional flow relies on several key assumptions to reduce the complexity of the governing equations. These include treating the fluid as incompressible, meaning remains constant; inviscid, or , where and frictional effects are neglected; and often steady-state, with time-independent unless otherwise specified. Additionally, three-dimensional effects, such as spanwise variations, are disregarded to focus solely on planar . These assumptions stem from the Euler equations for fluids and enable tractable mathematical models for many engineering applications. The theoretical foundations of two-dimensional flow emerged in the within hydrodynamics and , building on Leonhard Euler's 18th-century equations for inviscid fluid motion. Pioneering work by figures such as and advanced the use of for such flows, laying the groundwork for analytical treatments in subsequent decades. This development was particularly influential in early design and hydrodynamic problems. Unlike three-dimensional flows, which involve variations in all spatial coordinates and often require numerical solutions due to their inherent complexity, two-dimensional flow benefits from reduced dimensionality that permits exact analytical solutions, frequently employing complex variable techniques to represent the velocity field. This distinction facilitates deeper insights into fundamental flow behaviors without the full intricacy of volumetric effects.

Velocity field in two dimensions

In two-dimensional flow, the is represented by a \vec{v} that varies with position in a , typically the xy-, and possibly time. In Cartesian coordinates, this is expressed as \vec{v} = u(x,y,t) \hat{i} + v(x,y,t) \hat{j}, where u and v are the horizontal and vertical components, respectively. These components describe the local motion at any point (x,y) within the domain. For flows exhibiting radial symmetry or involving circular geometries, such as those around cylinders, the velocity field is conveniently expressed in cylindrical (polar) coordinates as \vec{v} = v_r(r,\theta,t) \hat{e}_r + v_\theta(r,\theta,t) \hat{e}_\theta, where v_r is the radial component and v_\theta is the azimuthal (tangential) component, with no axial velocity in the z-direction for strictly two-dimensional cases. This representation facilitates analysis in problems where the flow depends on radial distance r and angular position \theta. To visualize the flow kinematics, pathlines, streamlines, and streaklines are used, each providing distinct insights into particle trajectories. A pathline is the actual path traced by an individual particle over time, governed by the ordinary s \frac{dx}{dt} = u(x,y,t) and \frac{dy}{dt} = v(x,y,t) in Cartesian coordinates. A streamline is an instantaneous curve tangent to the velocity vector at every point, defined by the \frac{dy}{dx} = \frac{v}{u} (for u \neq 0), or equivalently in polar coordinates \frac{r d\theta}{dr} = \frac{v_\theta}{v_r} (for v_r \neq 0). A streakline connects the positions of all particles that have passed through a fixed point in the flow at different times, revealing injection patterns but coinciding with streamlines only in steady flows. In two-dimensional flows, these lines lie within the plane and simplify compared to three-dimensional cases. The magnitude of the , or speed q, quantifies the intensity and is given by q = \sqrt{u^2 + v^2} in Cartesian coordinates or q = \sqrt{v_r^2 + v_\theta^2} in cylindrical coordinates. The direction of the is specified by the angle \alpha = \tan^{-1}(v/u) relative to the x- in Cartesian coordinates, indicating the local orientation. These kinematic descriptors assume the adheres to the standard two-dimensional approximations, such as uniformity in the third dimension.

Governing principles

Continuity and momentum equations

In two-dimensional , the enforces mass conservation by requiring that the divergence of the field vanishes. For a with constant \rho, this simplifies to the condition that the sum of the partial derivatives of the components with respect to their respective spatial coordinates is zero: \frac{\partial u}{\partial x} + \frac{\partial v}{\partial y} = 0, where u and v are the components in the x and y directions, respectively. This equation ensures that the net through any closed curve in the plane is zero, reflecting the incompressibility assumption that volume is preserved under deformation. The momentum conservation in two-dimensional is governed by Euler's equations, which describe the acceleration of particles under gradients. In the x-, the equation is \frac{\partial u}{\partial t} + u \frac{\partial u}{\partial x} + v \frac{\partial u}{\partial y} = -\frac{1}{\rho} \frac{\partial p}{\partial x}, and in the y-, it takes the analogous form \frac{\partial v}{\partial t} + u \frac{\partial v}{\partial x} + v \frac{\partial v}{\partial y} = -\frac{1}{\rho} \frac{\partial p}{\partial y}, where p is the and t is time. These nonlinear partial differential equations capture the convective and local acceleration terms balanced against the force per unit mass, assuming no viscous effects or body forces other than . For steady, along a streamline, integrating Euler's equations yields Bernoulli's , which expresses conservation of per unit . The states that \frac{p}{\rho} + \frac{1}{2} q^2 + g z = \constant, where q = \sqrt{u^2 + v^2} is the speed, g is , and z is the (often negligible in horizontal flows). This relation holds under the assumptions of constant , no , and steady conditions, allowing , , and to interchange without loss along the flow path. In irrotational two-dimensional flows, where the vanishes, a \phi exists such that the components are u = \partial \phi / \partial x and v = \partial \phi / \partial y. Substituting into the for leads to for the potential: \nabla^2 \phi = \frac{\partial^2 \phi}{\partial x^2} + \frac{\partial^2 \phi}{\partial y^2} = 0. This simplifies the analysis of potential flows by reducing the vector field to a that satisfies harmonic properties.

Vorticity in two-dimensional flows

In two-dimensional flows, is defined as the out-of-plane component of the of the velocity field, representing a scalar measure of the local of elements. For a velocity field with components u in the x- and v in the y-, the \omega (specifically \omega_z) is given by \omega = \frac{\partial v}{\partial x} - \frac{\partial u}{\partial y}. This expression quantifies the tendency of parcels to rotate about the z-axis perpendicular to the flow plane. The sign of indicates the direction of : positive \omega corresponds to counterclockwise , while negative \omega corresponds to , assuming the standard right-hand rule for the . Zero implies irrotational flow, where the velocity field has no local spinning motion. In cylindrical coordinates (r, \theta, z) for two-dimensional flow in the r-\theta plane, with v_r and azimuthal velocity v_\theta, the component \omega_z takes the form \omega = \frac{1}{r} \frac{\partial (r v_\theta)}{\partial r} - \frac{1}{r} \frac{\partial v_r}{\partial \theta}. This formulation is particularly useful for analyzing flows with radial symmetry, such as those involving vortices or swirling motions. For incompressible two-dimensional flows, the evolution of is governed by the , which describes how is advected by the flow and diffused by : \frac{D \omega}{Dt} = \nu \nabla^2 \omega, where \frac{D}{Dt} is the , \nu is the kinematic , and \nabla^2 is the Laplacian operator. This simplified neglects baroclinic terms, which arise from gradients and are absent in constant- flows; it highlights that is conserved along fluid particle paths in inviscid conditions but spreads through viscous . The incompressibility assumption relates to the , ensuring the velocity field is divergence-free.

Potential flow formulation

Velocity potential and stream function

In two-dimensional irrotational flows, the velocity field can be represented by a scalar \phi, defined such that the velocity vector \vec{v} is the of \phi, i.e., \vec{v} = \nabla \phi. This representation implies that the of the velocity is zero, \nabla \times \vec{v} = 0, which is the mathematical condition for irrotationality. For incompressible flows, where the divergence of the velocity is zero, \nabla \cdot \vec{v} = 0, the velocity potential satisfies , \nabla^2 \phi = 0, or in Cartesian coordinates, \frac{\partial^2 \phi}{\partial x^2} + \frac{\partial^2 \phi}{\partial y^2} = 0. Solutions to this equation enable analytical determination of the velocity components u = \frac{\partial \phi}{\partial x} and v = \frac{\partial \phi}{\partial y}. For two-dimensional incompressible flows, the velocity field can alternatively be described using a \psi, a such that the velocity components are given by u = -\frac{\partial \psi}{\partial y} and v = \frac{\partial \psi}{\partial x}. These relations automatically satisfy the for , \frac{\partial u}{\partial x} + \frac{\partial v}{\partial y} = 0, as they represent the of a vector normal to the flow plane. When the flow is also irrotational, the likewise satisfies , \nabla^2 \psi = 0, or \frac{\partial^2 \psi}{\partial x^2} + \frac{\partial^2 \psi}{\partial y^2} = 0. The lines of constant \psi, known as streamlines, trace the direction of the field at every point. In such flows, the lines (\phi = \constant) are everywhere perpendicular to the streamlines (\psi = \constant), forming an orthogonal network that aligns with the flow direction. This orthogonality arises from the of the gradients being zero, \nabla \phi \cdot \nabla \psi = 0. The \phi and \psi are mathematically linked through the Cauchy-Riemann conditions, \frac{\partial \phi}{\partial x} = -\frac{\partial \psi}{\partial y} and \frac{\partial \phi}{\partial y} = \frac{\partial \psi}{\partial x}, which ensure both functions are harmonic conjugates (up to sign) in the . These conditions stem directly from the shared definitions and confirm that the pair (\phi, \psi) represents a valid solution.

Complex potential representation

In two-dimensional potential flow, the complex potential w(z) provides a unified representation of the flow field using , where z = x + i y is the complex position variable, x and y are the Cartesian coordinates, and w(z) = \phi(x, y) - i \psi(x, y). Here, \phi is the and \psi is the , serving as the real and negative imaginary parts, respectively. The components are obtained from the derivative of the complex potential as \frac{dw}{dz} = u - i v, where u = \frac{\partial \phi}{\partial x} = -\frac{\partial \psi}{\partial y} and v = \frac{\partial \phi}{\partial y} = \frac{\partial \psi}{\partial x} are the Cartesian components. For w(z) to represent a valid , it must be an of z, meaning it is differentiable in the complex sense everywhere in the flow domain except possibly at singularities. This analyticity implies that \phi and \psi satisfy the Cauchy-Riemann equations, \frac{\partial \phi}{\partial x} = -\frac{\partial \psi}{\partial y} and \frac{\partial \phi}{\partial y} = \frac{\partial \psi}{\partial x}, which in turn guarantee that the flow is irrotational (\nabla \times \mathbf{V} = 0) and incompressible (\nabla \cdot \mathbf{V} = 0), with both potentials satisfying \nabla^2 \phi = 0 and \nabla^2 \psi = 0. In polar coordinates, the transformation z = r e^{i \theta} is particularly useful for flows involving circular or cylindrical geometries, allowing w(z) to be expressed in terms of r and \theta. The corresponding Cauchy-Riemann equations become \frac{\partial \phi}{\partial r} = -\frac{1}{r} \frac{\partial \psi}{\partial \theta} and \frac{1}{r} \frac{\partial \phi}{\partial \theta} = \frac{\partial \psi}{\partial r}, facilitating analysis of radial and angular variations in the flow. A key advantage of the complex potential formulation is its exploitation of conformal mapping, where analytic functions preserve local angles and orientations between curves, enabling the transformation of complex flow domains into simpler ones while maintaining the physical flow properties. This simplifies the solution of boundary value problems, such as mapping a circular boundary to an airfoil shape, as demonstrated in classical applications.

Basic flow elements

Uniform flow

Uniform flow represents the simplest case of two-dimensional potential flow, characterized by a constant \vec{v} = U \hat{i}, where U is the constant speed in the positive x-direction. This flow is steady, irrotational, and incompressible, with no spatial variation in velocity magnitude or direction. The uniformity implies zero and zero , ensuring the flow aligns with the assumptions of . The for uniform flow is given by \phi = U x, from which the components are derived as u = \frac{\partial \phi}{\partial x} = U and v = \frac{\partial \phi}{\partial y} = 0. The corresponding is \psi = U y, satisfying the Cauchy-Riemann conditions with the potential and yielding the same components via u = \frac{\partial \psi}{\partial y} = U and v = -\frac{\partial \psi}{\partial x} = 0. These functions satisfy \nabla^2 \phi = 0 and \nabla^2 \psi = 0, confirming the irrotational nature. In complex variable formulation, the complex potential is w(z) = U z, where z = x + i y, combining the real part \phi = U x and imaginary part \psi = U y. The streamlines, defined by lines of constant \psi, are parallel straight lines perpendicular to the equipotential lines of constant \phi, resulting in horizontal flow paths along the x-direction.

Source and sink flows

In two-dimensional , a represents an idealized point from which emanates radially outward in the . The is irrotational and incompressible, with particles moving away from the origin along straight radial paths. The strength of the , denoted by m, is defined as the volume per unit depth perpendicular to the . The radial velocity component is given by v_r = \frac{m}{2\pi r}, while the tangential component is v_\theta = 0, resulting in a purely radial velocity field that diminishes inversely with distance r from the . The velocity potential for a source is \phi = \frac{m}{2\pi} \ln r, satisfying \nabla^2 \phi = 0 everywhere except at the origin. The corresponding stream function is \psi = \frac{m}{2\pi} \theta, where \theta is the polar angle. These functions ensure that the velocity components derive correctly as v_r = \frac{\partial \phi}{\partial r} = \frac{1}{r} \frac{\partial \psi}{\partial \theta} and v_\theta = \frac{1}{r} \frac{\partial \phi}{\partial \theta} = -\frac{\partial \psi}{\partial r}. In complex variable notation, the complex potential is w(z) = \frac{m}{2\pi} \ln z, where z = x + iy is the complex position, with \phi = \operatorname{Re}(w) and \psi = \operatorname{Im}(w). The streamlines, defined by \psi = \constant, are radial lines emanating from the origin at constant \theta. A flow is analogous to but directed inward toward the , representing fluid absorption. It is modeled by assigning a negative strength -m to expressions, yielding v_r = -\frac{m}{2\pi r} < 0, \phi = -\frac{m}{2\pi} \ln r, \psi = -\frac{m}{2\pi} \theta, and w(z) = -\frac{m}{2\pi} \ln z. The streamlines remain radial lines, but fluid motion converges toward the origin. This configuration approximates the two-dimensional limit of a line source or extending uniformly in the third .

Vortex and multipole flows

Irrotational vortex

The irrotational vortex, also known as a potential or free vortex, represents a fundamental element of two-dimensional incompressible characterized by circulatory motion around a central with constant circulation \Gamma. In this flow, the velocity field consists solely of a tangential component, with no radial motion, given by v_\theta = \frac{\Gamma}{2\pi r} and v_r = 0, where r is the radial distance from the . This velocity distribution ensures that the circulation around any closed contour encircling the equals \Gamma, independent of the contour's size, while the flow remains purely azimuthal and decays inversely with radius. Despite its circulatory nature, the irrotational vortex satisfies the condition of zero everywhere except at the origin, where the manifests as a singularity to account for the concentrated . This irrotationality outside allows the to be described using a \phi and \psi, which are harmonic conjugates in two dimensions. The is \phi = -\frac{\Gamma}{2\pi} \theta, where \theta is the polar angle, and the is \psi = -\frac{\Gamma}{2\pi} \ln r. These functions yield the velocity components through their gradients: v_r = \frac{\partial \phi}{\partial r} = \frac{\partial \psi}{\partial \theta} = 0 and v_\theta = \frac{1}{r} \frac{\partial \phi}{\partial \theta} = -\frac{\partial \psi}{\partial r} = \frac{\Gamma}{2\pi r}. In complex variable formulation, the irrotational vortex is elegantly represented by the complex potential w(z) = -\frac{i \Gamma}{2\pi} \ln z, where z = x + i y is the complex position. Here, the real part of w(z) corresponds to \phi and the imaginary part to \psi, confirming the orthogonal nature of equipotential lines (rays from the origin) and streamlines. The streamlines, defined by \psi = constant, form concentric circles centered at the origin, illustrating the purely rotational, non-divergent character of the flow. This representation underscores the vortex's role as a building block for more complex potential flows via superposition, while highlighting the infinite velocity and singular behavior at r = 0.

Doublet

A two-dimensional doublet arises as the limiting case of a source-sink pair in , where a source and sink each of strength m are separated by a small d along the x-axis, and the limit is taken as d \to 0 while keeping the product \mu = m d constant; here, \mu represents the strength of the doublet. For an x-directed doublet at the origin, the velocity potential \phi and stream function \psi are \phi = -\frac{\mu}{2\pi} \frac{x}{x^2 + y^2}, \psi = \frac{\mu}{2\pi} \frac{y}{x^2 + y^2}. These satisfy and the Cauchy-Riemann conditions, ensuring the flow is irrotational and incompressible. The velocity components in Cartesian coordinates are u = \frac{\mu}{2\pi} \frac{x^2 - y^2}{(x^2 + y^2)^2}, \quad v = \frac{\mu}{2\pi} \frac{2 x y}{(x^2 + y^2)^2}. These expressions describe the dipole-like velocity distribution, with singularities at the . In notation, the potential is w(z) = -\frac{\mu}{2\pi z}, where z = x + i y. The streamlines, given by \psi = \ constant, consist of a family of circles tangent to the x-axis at the , reflecting the symmetric structure of the flow.

Applications and combinations

Superposition of basic flows

In two-dimensional potential flow, the superposition principle allows the combination of fundamental flow solutions to construct more complex velocity fields, as the governing Laplace's equation for the velocity potential \phi or stream function \psi is linear. This linearity implies that if \phi_1 and \phi_2 are solutions satisfying \nabla^2 \phi = 0, then the total potential \phi = \phi_1 + \phi_2 also satisfies the equation, enabling the algebraic addition of individual flows without altering their irrotational and incompressible properties. Similarly, in complex potential representation, the total w(z) = w_1(z) + w_2(z) preserves analyticity, facilitating analytical solutions for composite flows. A key example is the superposition of uniform flow and a source flow, which models the Rankine half-body—a streamlined shape open at the downstream end, useful for approximating blunt-nosed bodies in aerodynamics. The uniform flow has potential \phi_U = U x, where U is the free-stream velocity, while the two-dimensional source at the origin has \phi_s = \frac{m}{2\pi} \ln r with strength m. The combined potential is \phi = U x + \frac{m}{2\pi} \ln r, where r = \sqrt{x^2 + y^2}. The corresponding stream function \psi = U y + \frac{m}{2\pi} \theta reveals a stagnation streamline \psi = 0 that forms the half-body boundary, with the stagnation point located at x = -m/(2\pi U) on the upstream axis; far downstream, the body half-width asymptotes to m/(2U). Another basic superposition involves a source and a sink of equal strength m separated by a small distance, approximating a dipole (or doublet) flow when the separation approaches zero while maintaining constant product of strength and separation. The source potential \phi_s = \frac{m}{2\pi} \ln r_1 and sink \phi_k = -\frac{m}{2\pi} \ln r_2, where r_1 and r_2 are distances from the points, yield a combined field that, in the limit, produces \phi_d = \frac{\mu \cos \theta}{2\pi r} with dipole moment \mu, representing a flow with radial symmetry useful as a building block for more advanced configurations. The extends superposition to enforce boundary conditions, such as no normal flow through a , by placing mirror-image singularities outside the domain of interest. For a straight at y=0 and a source of strength m at (0, a), an image source of strength m is placed at (0, -a); the total potential \phi = \frac{m}{2\pi} \ln (r_1 r_2) ensures \partial \phi / \partial y = 0 on the wall by , where r_1 and r_2 are distances to the source and image, respectively. This technique simplifies problems involving solid boundaries, such as flow in channels or near flat plates, by transforming them into unbounded flows with added singularities.

Flow around airfoils and cylinders

In two-dimensional potential flow, the flow past a circular cylinder of radius a in a uniform stream of speed U at infinity is obtained by superposing a uniform flow and a doublet. The complex potential for this configuration is given by w(z) = U \left( z + \frac{a^2}{z} \right), where z is the complex coordinate in the plane perpendicular to the cylinder axis. This representation yields streamlines that form closed loops around the cylinder, with no penetration through its surface, and the flow remains symmetric fore and aft, resulting in zero lift and zero drag in the inviscid approximation. To generate lift, an irrotational vortex of circulation \Gamma is added to the uniform flow plus doublet potential, modeling the Magnus effect observed when a cylinder rotates in a fluid stream. The modified complex potential becomes w(z) = U \left( z + \frac{a^2}{z} \right) - \frac{i \Gamma}{2\pi} \log z, which introduces asymmetry in the velocity distribution around the cylinder. This circulation shifts higher velocities to one side and lower to the other, producing a pressure difference that yields a lift force per unit length of L = \rho U \Gamma, where \rho is the fluid density; the force acts perpendicular to the free-stream direction. The Joukowski airfoil extends this approach by using conformal to transform the around a circular into around an -shaped body. The is defined as z = \zeta + \frac{b}{\zeta}, where \zeta is the complex coordinate in the circle plane, z is in the plane, and b controls the asymmetry, producing a cusped trailing edge when the circle is offset appropriately. Applying the potential in the \zeta-plane and transforming yields the over the , with circulation added to satisfy the of smooth off the trailing edge. The lift on such airfoils, and more generally on any two-dimensional body in with circulation \Gamma, is given by the Kutta-Joukowski theorem: L = \rho U \Gamma per unit span, directed perpendicular to the oncoming flow. This theorem derives from the Blasius formula for the force on a body, which expresses the complex force as a contour integral of the complex velocity dw/dz around the body: F_x - i F_y = \frac{i \rho}{2} \oint \left( \frac{dw}{dz} \right)^2 dz, where evaluation for circulatory flows simplifies to the linear relation with \Gamma. The theorem underscores that lift in inviscid theory arises solely from circulation, independent of body shape details beyond enforcing the .

References

  1. [1]
    Two-Dimensional Flow - Richard Fitzpatrick
    Two-Dimensional Flow. Fluid motion is said to be two-dimensional when the velocity at every point is parallel to a fixed plane, and is the same everywhere on ...
  2. [2]
    Types of Fluid Flows – Introduction to Aerospace Flight Vehicles
    Two-Dimensional Flow · Dynamic Pressure · Airfoil Forces and Moments · Aerodynamic Coefficients · Representative Aerodynamic Coefficients · Lift Characteristics ...
  3. [3]
    [PDF] 6 Two dimensional hydrodynamics and complex potentials
    In this section we will exploit this connection to look at two dimensional hydrodynamics, i.e. fluid flow. Since static electric fields and steady state ...
  4. [4]
    Two dimensional flow - Wärtsilä
    Fluid motion can be said to be a two-dimensional flow when the flow velocity at every point is parallel to a fixed plane.
  5. [5]
    7: Two Dimensional Hydrodynamics and Complex Potentials
    Sep 5, 2021 · Here are the assumptions about the flow, we'll discuss them further below: (1) The flow is stationary, (2) The flow is incompressible, and (3) ...
  6. [6]
    Dimensional Flow - an overview | ScienceDirect Topics
    To summarize, two-dimensional flow is fluid motion where the velocity at all points is parallel to a given plane. We have already seen how the principle of ...
  7. [7]
    Potential Flow Theory – Introduction to Aerospace Flight Vehicles
    History. The origins of the potential flow theory can be traced back to the 18th century, with contributions from Daniel Bernoulli and Leonhard Euler.
  8. [8]
    [PDF] FROM NEWTON'S MECHANICS TO EULER'S EQUATIONS
    The Euler equations of hydrodynamics, which appeared in their present form in the 1750s, did not emerge in the middle of a desert.
  9. [9]
    [PDF] Kinematics of fluids - The Open University
    A flow is said to be two-dimensional if the flow parameters depend on only two space coordinates. Using Cartesian coordinates (x, y and z) or cylindrical polar ...
  10. [10]
    [PDF] Potential Flow Theory - MIT
    The velocity vector can then be transformed into. Cartesian coordinates at point P using equations (4.22) and (4.23). Figure 4: Velocity vector at point P due ...
  11. [11]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics 1.7 ...
    Dec 15, 2021 · In this section, we'll present three forms of flow kinematics: streamlines, pathlines, and streaklines. Each of these lines provides different ...
  12. [12]
    [PDF] The Velocity Field - MSU College of Engineering
    Often one sees pictures of "streamlines" made visible by the injection of smoke or dye into a flow as is shown in Fig. 4.3. Actually, such pictures show ...
  13. [13]
    Euler Equations
    The Euler equations of fluid dynamics in two-dimensional, steady form and incompressible form. On this slide we have two versions of the Euler Equations ...
  14. [14]
    Bernoulli's Equation
    Bernoulli's Equation. The Bernoulli equation states that,. where. points 1 and 2 lie on a streamline,; the fluid has constant density,; the flow is steady, ...
  15. [15]
    [PDF] Vorticity
    The fluid is a liquid of constant density ρ with a free surface given by z = Z(r) in cylindrical polar coordinates, see figure 3.3. The pressure above the free ...
  16. [16]
    [PDF] Fluid Mechanics
    In contrast, in fluid mechanics the vorticity ω is thought of as ... by working in cylindrical polar coordinates and introducing a stream function Ψ(r, θ).
  17. [17]
    [PDF] The Vorticity equation
    • For 2D flows, the vorticity transport equation. Dω. Dt. = ν∇2ω together with the equation for the vorticity in terms of the streamfunction ω = −∇2ψ and u ...
  18. [18]
    Velocity Potentials and Stream Functions
    We conclude that, for two-dimensional, irrotational, incompressible flow, the velocity potential and the stream function both satisfy Laplace's equation.
  19. [19]
    [PDF] Velocity Potential
    Therefore, in sketches of incompressible planar potential flow these lines, streamlines and equipotentials, form an orthogonal net that can be visualized as ...
  20. [20]
    Introduction - Richard Fitzpatrick
    Introduction. This chapter describes the use of complex analysis to facilitate calculations in two-dimensional, incompressible, irrotational fluid dynamics.
  21. [21]
    [PDF] CHAPTER 10 ELEMENTS OF POTENTIAL FLOW ( )# !2U
    Two-dimensional potential flows can be constructed from any analytic function of a complex variable, W z( ). From the Cauchy-Riemann conditions (10.76) !2".
  22. [22]
    [PDF] Potential Flows
    The flow is called potential flow or irrotational. Two-dimensional potential flow: u = ∂φ. ∂x v = ∂φ.
  23. [23]
    [PDF] Hydrodynamics
    MOTION OF A LIQUID IN TWO DIMENSIONS. 59. Stream-function ... Hydrodynamics. Tidal oscillations of a rotating sheet of water. Plane sheet ...
  24. [24]
    [PDF] Lecture 11
    A potential vortex is irrotational everywhere except as the origin, where the vorticity is infinite. A delta function of vorticity. The circulation of any ...
  25. [25]
    Dipole (doublet flow) - MIT
    Dipole (doublet flow). A Dipole is a superposition of a sink and a source with the same strength. \begin{figure} \centering\ ...
  26. [26]
    7.3 Potential Flow Theory - Fluid Mechanics - Fiveable
    Doublet flow: $\phi = -\frac{\mu}{2\pi} \frac{x}{x^2 + y^2}$ and $\psi ... 2 + (-\frac{\partial \psi}{\partial x})^2}$; Stagnation points occur where ...
  27. [27]
    Method of Images - Richard Fitzpatrick
    We deduce that two complex velocity potentials, corresponding to distinct, two-dimensional, irrotational, incompressible flow patterns, can be superposed to ...
  28. [28]
    [PDF] Planar Rankine Half-Bodies
    If we superimpose a planar source on a uniform stream we can create streamlines which can be replaced by a solid body so as to generate the potential flow ...
  29. [29]
    V. Potential Flows – Intermediate Fluid Mechanics
    The sign convention is such that a counterclockwise rotation results in a downward force, and a clockwise rotation results in a upward force for flow along the ...
  30. [30]
    6: Potential Flows - Engineering LibreTexts
    May 1, 2022 · In two dimensional flow a source or a sink of flow is possible, since it implies that flow enters or leaves a given two dimensional plane. We ...
  31. [31]
    [PDF] Lecture Notes for 436-351 Thermofluids 2 Unit 1: Potential Flow
    Mar 10, 2003 · the origin. The stream function and velocity potential for this flow in cartesian coordinates can be obtained by substituting z = x + iy ...
  32. [32]
    [PDF] Fluids – Lecture 15 Notes - MIT
    The resulting flow is a doublet with strength κ. κ=const. In polar coordinates this is the equation for circles of diameter d, centered on x, y = (0, ±d/2).
  33. [33]
    [PDF] Magnus Effect
    The “Magnus Effect” is the lift produced by a rotating cylinder in a uniform stream. It can be predicted using the appropriate potential flow solution for a ...
  34. [34]
    10.3.1.1: Adding Circulation to a Cylinder - Engineering LibreTexts
    Mar 5, 2021 · The circulation mimics the Magnus's effect and hence it is used in representative flow. ... flow of with vortex can represent the viscous flow.
  35. [35]
    [PDF] Solution to the Magnus Effect - MIT OpenCourseWare
    Equation (29) is true for irrotational flow around any 2D object and not only cylinders. This equations is known as the Kutta-Zhukhousky lift theorem.
  36. [36]
    [PDF] LESSON 08A: CIRCULATION AND LIFT JM Cimbala
    • Discuss the physical significance: Spinning cylinders and spheres (the Magnus effect). Potential Flow Around a Circular Cylinder: Superposition of a Vortex ...
  37. [37]
    Joukowsky Airfoil - Complex Analysis
    The Joukowski map defines a one-to-one conformal mapping from the exterior of the unit circle, onto the exterior of the line segment.Missing: zeta + b/ zeta
  38. [38]
    Classic Airfoil Theory – Introduction to Aerospace Flight Vehicles
    Be able to explain conformal transformations and apply them to map potential flows about a circular cylinder onto airfoil-shaped boundaries. Understand and ...
  39. [39]
    On conformal mapping and the Joukowski transform - ResearchGate
    Sep 29, 2024 · This essay discusses conformal mapping and the classical Joukowski transform derived by Nikolai Zhukovsky in the 1910s.
  40. [40]
    Kutta-Joukowski Lift Theorem - HyperPhysics
    Kutta-Joukowski Lift Theorem. Two early aerodynamicists, Kutta in Germany and Joukowski ... L = ρGV. where ρ is the air density, V is the velocity of air ...Missing: rho Gamma Blasius
  41. [41]
    [PDF] Why airplanes fly, and ships sail - Purdue Math
    L = −iρv∞C. This is called the Kutta–Joukowski theorem. In words: The lifting force is the product of the density, velocity and circulation, and directed ...Missing: rho | Show results with:rho
  42. [42]
    [PDF] 21 Classical aerofoil theory
    and we have proved Blasius' lemma. 21.4 Kutta-Joukowski theorem. We now use Blasius' lemma to prove the Kutta-Joukowski lift theorem. For flow around a plane ...Missing: formula | Show results with:formula
  43. [43]
    Blasius Theorem - an overview | ScienceDirect Topics
    The result (6.62) is called the Kutta-Zhukhovsky lift theorem, and it plays a fundamental role in aero- and hydrodynamics. As described in Chapter 14, the ...
  44. [44]
    Ideal Lift of a Spinning Ball | Glenn Research Center - NASA
    Aug 28, 2025 · The Kutta-Joukowski lift theorem for a single cylinder states the lift per unit length L is equal to the density rho (ρ) of the air times the strength of the ...Missing: U Blasius