Fact-checked by Grok 2 weeks ago

Orthogonal trajectory

In , orthogonal trajectories are families of curves that intersect a given family of curves at right angles, such that the tangent lines to the curves at each point of intersection are . To determine orthogonal trajectories, one starts with the describing the of the original family, typically of the form \frac{dy}{dx} = f(x, y), and replaces the slope with its negative , \frac{dy}{dx} = -\frac{1}{f(x, y)}, to obtain the equation for the orthogonal family, which is then solved. A classic example involves the family of circles centered at the , x^2 + y^2 = c^2, whose is x + y \frac{dy}{dx} = 0; the orthogonal trajectories are then the straight lines through the , y = mx. The concept originated in the late 17th century, when proposed the problem of finding curves that intersect a given family at a constant angle—specifically 90 degrees—to challenge English mathematicians, including , who solved it rapidly using methods involving tangents and , though his approach was later critiqued for not fully resolving the underlying . and his nephews Nicolaus I and Nicolaus II provided more elegant solutions, advancing the topic toward partial and related problems like reciprocal trajectories. Orthogonal trajectories hold significance in the of plane curves and have applications in , such as modeling the relationship between lines of equal potential and current flow lines in a conducting sheet, or between lines and surfaces in .

Introduction

Definition

Orthogonal trajectories arise in the context of a one-parameter of curves, also known as a pencil of curves, which consists of a collection of plane curves parameterized by a single variable, such as x^2 + y^2 = c for varying c > 0. Given such a family, the orthogonal trajectories form another one-parameter family of curves that intersect every member of the original family at right angles everywhere along their points of intersection. The geometric condition for orthogonality in the plane requires that, at each intersection point, the tangent lines to the two curves are perpendicular. For curves with defined slopes m_1 and m_2 (excluding vertical tangents), this perpendicularity holds if the product of the slopes satisfies m_1 m_2 = -1. A foundational example illustrates this concept: the family of concentric circles centered at the origin, given by x^2 + y^2 = c for c > 0. Differentiating implicitly yields the slope of the tangent to these circles as \frac{dy}{dx} = -\frac{x}{y}. For the orthogonal family, the slope m must satisfy m \left( -\frac{x}{y} \right) = -1, so m = \frac{y}{x}, corresponding to the family of straight lines through the origin, y = kx for varying k. These lines intersect the circles at right angles, as radial lines are perpendicular to the tangents of circles centered at the origin.

Historical Development

The concept of orthogonal trajectories emerged in the late , with early explorations by and in the 1690s, particularly in geometric discussions related to curve families in and . Leibniz, collaborating with the Bernoulli brothers in the 1690s, explored methods to construct curves to given families, applying differential techniques to problems in and . In 1698, extended this work by solving for orthogonal trajectories of single-parameter curve families using Leibnizian differentials. The played a pivotal role in advancing the theory during the 1690s and early 1700s, linking orthogonal trajectories to reciprocal trajectories through solutions in their publications. Their efforts, amid the Leibniz-Newton priority dispute, positioned orthogonal trajectories as a key test of proficiency. In 1716, Leibniz proposed the problem of finding curves that intersect a given family at a constant angle (specifically 90 degrees) as a challenge to English mathematicians, including , who solved it rapidly using fluxions involving tangents and ; however, his approach was later critiqued by for not fully resolving the underlying . A significant milestone came in 1718 with the joint dissertation by and his son Nicolaus II Bernoulli (1695–1726), published in the Acta Eruditorum of , which systematically addressed orthogonal trajectories via partial s and construction methods for families. Nicolaus II further contributed in 1720 by posing the problem of reciprocal orthogonal trajectories as a deliberate challenge to Newtonian scholars, highlighting ongoing continental advancements. By the mid-18th century, the topic evolved from these geometric origins into a core element of theory, influencing broader developments in and paving the way for later geometric applications.

Analytical Methods

In Cartesian Coordinates

To derive the orthogonal trajectories of a given family of curves in Cartesian coordinates, begin with the family represented implicitly as F(x, y, c) = 0, where c is a parameter. Differentiate this equation implicitly with respect to x to obtain F_x + F_y y' = 0, where y' = \frac{dy}{dx}. Solve for y' and eliminate the parameter c to yield the differential equation y' = f(x, y) describing the slopes of the original family. The orthogonal trajectories have slopes given by the negative reciprocal, y' = -\frac{1}{f(x, y)}, forming a new first-order ordinary differential equation. Solve this ODE to obtain the equation of the orthogonal family. Consider the family of concentric circles given by x^2 + y^2 = c. Differentiate implicitly: $2x + 2y y' = 0, so y' = -\frac{x}{y}. The orthogonal slopes satisfy y' = \frac{y}{x}. This separable equation rearranges to \frac{dy}{y} = \frac{dx}{x}. Integrating both sides yields \ln |y| = \ln |x| + k, or y = m x where m = \pm e^k is an arbitrary constant, representing radial lines through the . For the family of parabolas y = c x^2, differentiate to get y' = 2 c x. Substitute c = \frac{y}{x^2} to eliminate the : y' = \frac{2 y}{x}. The orthogonal slopes are y' = -\frac{x}{2 y}. Separate variables: $2 y \, dy = -x \, dx. Integrate to obtain y^2 = -\frac{x^2}{2} + k, or x^2 + 2 y^2 = k (with k > 0), a family of ellipses centered at the origin. This procedure assumes the curves in the family are differentiable with respect to x, non-singular (avoiding points where the tangent is undefined), and defined in the Euclidean plane. Singular cases, such as vertical tangents where f(x, y) is infinite, require careful handling, often by switching to parametric forms or considering the reciprocal slope directly.

In Polar Coordinates

In polar coordinates, the analytical method for finding orthogonal trajectories is adapted for families of curves exhibiting , such as spirals or circles centered at the . Consider a family of curves given by F(r, \varphi, c) = 0, where c is a . To derive the , differentiate implicitly with respect to the angular variable \varphi, treating r as a of \varphi, and eliminate the parameter c. This yields a of the form \frac{dr}{d\varphi} = r^2 f(r, \varphi), where f(r, \varphi) encapsulates the relation for the original family. For the orthogonal trajectories, the condition of perpendicularity requires that the product of the angular slopes be -1, leading to the replacement \frac{dr}{d\varphi} = -\frac{1}{f(r, \varphi)}, or, equivalently, \frac{d\varphi}{dr} = -f(r, \varphi). This new differential equation is then solved to obtain the orthogonal family. The procedure leverages the polar form's natural handling of angular dependence, simplifying the (ODE) compared to Cartesian equivalents for symmetric cases. A representative example illustrates this using the family of cardioids r = c (1 + \cos \varphi). Differentiating with respect to \varphi gives \frac{dr}{d\varphi} = -c \sin \varphi. Substituting the original yields \frac{1}{r} \frac{dr}{d\varphi} = -\frac{\sin \varphi}{1 + \cos \varphi}, or equivalently \frac{dr}{d\varphi} = -r \tan\left(\frac{\varphi}{2}\right), confirming f(r, \varphi) = -\frac{1}{r} \tan\left(\frac{\varphi}{2}\right). For orthogonality, replace with \frac{dr}{d\varphi} = r \cot\left(\frac{\varphi}{2}\right). Separating variables and integrating results in the orthogonal family r = d (1 - \cos \varphi), which is another cardioid family rotated by \pi radians. The polar approach relates to Cartesian coordinates through the transformations x = r \cos \varphi and y = r \sin \varphi, with the Cartesian slope \frac{dy}{dx} given by \frac{dy}{dx} = \frac{\frac{dr}{d\varphi} \sin \varphi + r \cos \varphi}{\frac{dr}{d\varphi} \cos \varphi - r \sin \varphi}. This method offers advantages in problems with circular or spiral , where Cartesian coordinates often result in complex, non-separable ODEs due to the mixing of radial and angular terms; polar coordinates naturally decouple these, facilitating analytical solutions.

Numerical Methods

Runge-Kutta Integration

When analytical solutions for the orthogonal trajectories are unavailable, particularly for nonlinear equations of the form y' = -\frac{1}{f(x, y)}, numerical methods become essential. In such cases, the problem is reformulated as an (IVP) by selecting an initial point (x_0, y_0) on one of the given curves and integrating the orthogonal forward and backward to approximate the trajectory passing through that point. The fourth-order Runge-Kutta (RK4) method is a widely adopted explicit numerical for this purpose, providing a balance of accuracy and computational efficiency for ODEs like those arising in orthogonal trajectory problems. To apply RK4, consider the orthogonal ODE y' = g(x, y), where g(x, y) = -\frac{1}{f(x, y)}. Starting from (x_n, y_n), the next point (x_{n+1}, y_{n+1}) with step size h is computed iteratively as follows: \begin{align*} k_1 &= h \, g(x_n, y_n), \\ k_2 &= h \, g\left(x_n + \frac{h}{2}, y_n + \frac{k_1}{2}\right), \\ k_3 &= h \, g\left(x_n + \frac{h}{2}, y_n + \frac{k_2}{2}\right), \\ k_4 &= h \, g(x_n + h, y_n + k_3), \\ y_{n+1} &= y_n + \frac{k_1 + 2k_2 + 2k_3 + k_4}{6}, \\ x_{n+1} &= x_n + h. \end{align*} This process generates a sequence of points approximating the orthogonal curve, which can be plotted or interpolated for visualization. The method evaluates four times per step, estimating the via a weighted that matches the Taylor expansion up to fourth order. For illustration, consider the family of curves y = e^x + c, which has slope f(x, y) = e^x, yielding the orthogonal y' = -e^{-x}. Starting from an initial point such as (0, 1) on a member of the family (e.g., c = 0), RK4 with h = 0.1 can generate points along the : for instance, the first step yields y(0.1) \approx 0.9048, and subsequent steps trace the approximate curve y \approx e^{-x}, demonstrating how builds the orthogonal path even when an exact exists. This approach is particularly valuable for more complex families where g(x, y) is nonlinear and inseparable. Regarding accuracy, RK4 exhibits a local of O(h^5) per step, arising from the omission of higher-order terms in the , and a global error of O(h^4) over the integration interval, assuming sufficient of g(x, y). To enhance and control errors, especially for stiff or varying ODEs in problems, adaptive step-size variants like the Runge-Kutta-Fehlberg (RKF45) can be employed, which dynamically adjusts h based on embedded error estimates to maintain a specified .

Finite Difference Approximations

Finite difference approximations offer a , -based for computing orthogonal trajectories, enabling the batch generation of multiple curves across a domain for visualization or analysis when the given family's is known numerically. The approach begins by discretizing the into a uniform rectangular , where points serve as evaluation locations for the slope f(x, y) of the original family of curves. At each point, the orthogonal direction is determined by taking the vector to the local , effectively using the slope condition where the orthogonal increment satisfies \Delta y / \Delta x \approx -1 / f(x, y). This creates a representing the directions of the orthogonal trajectories. To trace the trajectories, the curves are propagated from initial points using finite difference schemes, such as the forward , which approximates the solution to the underlying first-order differential equation by stepping along the orthogonal directions: y_{n+1} = y_n + h \left( -\frac{1}{f(x_n, y_n)} \right), where h is the step size aligned with the grid spacing, and x_{n+1} = x_n + h. Central differences can also be applied for improved accuracy in estimating slopes between grid points. For locations not coinciding with mesh nodes, of the from surrounding grid values ensures continuity, allowing smoother propagation without aliasing. This method is particularly effective for handling non-uniform or data-driven slope fields derived from simulations. A representative example involves approximating orthogonal trajectories to a numerical of streamlines in fluid flow, where the given family's slopes are extracted from vectors on a . The field is then constructed, and trajectories are traced to produce lines, often visualized as overlaid plots to confirm right-angle intersections. Such computations facilitate the creation of orthogonal coordinate for further numerical simulations. While this grid-based distributes the computation spatially and can better manage singularities in the by avoiding direct path integration through problematic regions, it incurs a computational cost of O(n^2) operations for an n \times n due to the need to evaluate and store at all points. Improvements include higher-order schemes, such as bicubic methods, to yield smoother curves and reduce discretization errors, enhancing overall fidelity without proportionally increasing expense.

Isogonal Trajectories

Isogonal trajectories generalize the concept of orthogonal trajectories by considering families of curves that intersect a given family at a constant angle \alpha \neq 90^\circ. To derive the differential equation for isogonal trajectories, suppose the given family of curves satisfies the differential equation \frac{dy}{dx} = f(x, y), where f(x, y) represents the slope m of the tangent to the given curves. Let g(x, y) be the slope of the isogonal trajectory, such that the angle \alpha between the tangents satisfies \tan \alpha = \left| \frac{g - f}{1 + f g} \right|. Solving for g yields the formula g = \frac{f + \tan \alpha}{1 - f \tan \alpha}, assuming the appropriate orientation for the angle. Thus, the differential equation for the isogonal trajectories is \frac{dy}{dx} = \frac{f(x, y) + \tan \alpha}{1 - f(x, y) \tan \alpha}. This framework reduces to the orthogonal case when \alpha = 90^\circ, as \tan 90^\circ \to \infty, simplifying the expression to g = -\frac{1}{f}. A classic example illustrates the power of this method: consider the family of concentric circles centered at the , given by x^2 + y^2 = c^2, with slope f(x, y) = -\frac{x}{y}. For \alpha = [45](/page/45)^\circ, where \tan [45](/page/45)^\circ = 1, the slope of the isogonal trajectories becomes g = \frac{-\frac{x}{y} + 1}{1 - (-\frac{x}{y})(1)} = \frac{y - x}{y + x}. The resulting is \frac{dy}{dx} = \frac{y - x}{y + x}. To solve it, substitute t = \frac{y}{x}, leading to x \frac{dt}{dx} = \frac{t - 1}{t + 1} - t, but more directly by separation: \frac{1 + t}{1 + t^2} \, dt = -\frac{dx}{x}. Integrating gives \ln(1 + t^2) + 2 \arctan t = -2 \ln x + k, or in polar coordinates r = a e^{b \varphi} with b = -1, confirming the trajectories are logarithmic spirals.

Reciprocal Trajectories

Reciprocal trajectories constitute a of curves for which the original given serves as the orthogonal trajectories, typically realized through geometric transformations such as circle inversion (point inversion) or polar reciprocity with respect to a conic section. This duality arises because inversion preserves , ensuring that curves orthogonal to the original map to curves that intersect the transformed at right . The concept emerged in the late 17th century as a byproduct of investigations into orthogonal trajectories by the brothers, particularly Bernoulli's 1698 solution for single-parameter families of curves, with further development in their joint dissertation linking it to pole-polar relations in conics. These relations, central to , treat points and lines dually via the polar of a point with respect to a conic, transforming curve families while maintaining envelope and locus properties. Mathematically, if families C_1 and C_2 are with respect to an inversion, the orthogonal trajectories of C_1 coincide with C_2 under the inversion mapping; for instance, the polar of a of conics (a linear family sharing common points or lines) yields another of conics. Key properties of trajectories include angle preservation under inversion, which facilitates their role in projective ity by interchanging points and tangents without altering incidence structures, distinguishing them from isogonal trajectories that generalize to arbitrary fixed angles rather than emphasizing families.

Applications

Curved Coordinate Systems

Orthogonal trajectories play a fundamental role in the construction of non-Cartesian coordinate systems, where one family of curves serves as the primary coordinate lines—such as surfaces—and the orthogonal family acts as the conjugate coordinates, such as field lines, ensuring that the coordinate grid intersects at right angles everywhere. This diagonalizes the , simplifying the expressions for differential operators like the and Laplacian in the new system. In such systems, solutions to \nabla^2 \phi = 0 separate into independent functions along each coordinate direction, as the condition aligns with the requirements for additive . A prominent example is elliptic coordinates, defined by confocal ellipses and hyperbolas that form mutually orthogonal trajectories. In two dimensions, the elliptic coordinates (\mu, \nu) are given by x = a \cosh \mu \cos \nu and y = a \sinh \mu \sin \nu, where the ellipses correspond to constant \mu and the hyperbolas to constant \nu, with a as the focal distance. The scale factors for these coordinates are identical due to the symmetry: h_\mu = h_\nu = a \sqrt{\sinh^2 \mu + \sin^2 \nu}, which determine the arc length elements ds^2 = h_\mu^2 d\mu^2 + h_\nu^2 d\nu^2. To construct such a , one begins with a given family of curves, such as the confocal ellipses parameterized by a , and solves the orthogonal trajectory problem to obtain the conjugate family, forming a complete orthogonal grid that covers the plane without singularities except at the foci. This process ensures the is diagonal in the new coordinates, with off-diagonal terms vanishing due to the perpendicularity of the vectors. These orthogonal systems offer significant advantages in solving partial differential equations (PDEs) over domains exhibiting natural symmetry, such as elliptical boundaries, by enabling in equations like the \nabla^2 \psi + k^2 \psi = 0, which reduces the PDE to ordinary differential equations along each coordinate. For instance, in polar coordinates—a simpler orthogonal system derived similarly from circles and rays—the Laplacian separates into radial and angular parts, illustrating the broader utility for problems with .

Physical Interpretations

In , surfaces, which are loci of constant , are everywhere orthogonal to the lines, as the is the negative of the potential, and the is to level surfaces of the potential. This orthogonality implies that the trajectories of the lines form a family of curves to the family of curves. A classic example is the field due to a point charge, where the surfaces are concentric spheres centered at the charge, and the orthogonal field line trajectories are radial straight lines emanating from the charge. In , particularly for irrotational incompressible flows, streamlines—curves tangent to the field—are orthogonal to lines defined by the , since the is the of the potential. This property holds because the level sets of the potential are perpendicular to its . An illustrative case is the of a uniform stream past a , which produces a dipole-like field; here, the lines form a family of circles, while the orthogonal streamlines curve around the symmetrically. Similar interpretations arise in other physical contexts. In steady-state heat conduction, isotherms (surfaces of constant temperature) are orthogonal to the trajectories of lines, as the heat flux is proportional to the negative temperature , analogous to the electric case. Gravitational fields exhibit the same structure, with field lines orthogonal to equipotential surfaces, mirroring due to the shared form of for the potentials. Mathematically, these physical interpretations stem from the condition that two scalar potentials and have orthogonal level sets if their gradients satisfy \nabla \phi \cdot \nabla \psi = 0, ensuring the directions normal to the respective level surfaces are perpendicular. In two dimensions, this orthogonality ties to : if and are real and imaginary parts of an , they satisfy the Cauchy-Riemann equations, guaranteeing harmonic conjugates whose level curves form orthogonal families.

References

  1. [1]
  2. [2]
    [PDF] Orthogonal trajectories, MATH 3410 - Differential equations for ...
    Jan 29, 2018 · Orthogonal trajectories are of interest in the geometry of plane curves, and also in certain parts of applied mathematics.Missing: definition | Show results with:definition
  3. [3]
    Charles Bossut on Leibniz and Newton Part 2 - MacTutor
    The problem of orthogonal trajectories gave birth to that of reciprocal trajectories, which was proposed at the end of the dissertation of the Bernoullis.
  4. [4]
    Chapter 4 Nicolaus I Bernoulli and Orthogonal Trajectories
    This chapter discusses the contributions to partial differential calculus and the orthogonal trajectory problem made by Nicolaus I. Bernoulli.
  5. [5]
    [PDF] ES.1803 S24: Reading: Topic 1: Introduction to Differential Equations
    Given a one-parameter family of plane curves, its orthogonal trajectories are another one- parameter family of curves, each one of which is perpendicular to all ...
  6. [6]
    [PDF] Separable Equations - Penn Math
    Nov 30, 2009 · Each member of the family y = mx of straight lines through the origin is an orthogonal trajectory of the family x2 + y2 = r2 of concentric ...
  7. [7]
    The Newton–Leibniz Calculus Controversy, 1708–1730
    A general method for finding orthogonal trajectories was sought in the first decades of the eighteenth century. Leibniz, Jacob and Johann Bernoulli, but ...Missing: dissertation | Show results with:dissertation
  8. [8]
    [PDF] Two Leibnizian Manuscripts of 1690 Concerning Differential Equations
    In the late 169Os, he worked with the. Bernoullis on a problem important for optics, that of orthogonal trajectories, finding a family of curves that cut a ...
  9. [9]
    Why is differentiation under the integral sign named the Leibniz rule?
    Jan 4, 2019 · The rule appears in a Leibniz's 1697 letter to Bernoulli, as a side result in their long correspondence on the problem of orthogonal trajectories.
  10. [10]
    BIG HISTORY OF DIFFERENTIAL EQUATIONS
    Sep 15, 2017 · In 1698 John Bernoulli solved the problem of determining the orthogonal trajectories of single parameter family of curves. In 1701 James ...<|control11|><|separator|>
  11. [11]
  12. [12]
  13. [13]
    [PDF] DIFFERENTIAL EQUATION - Uttarakhand Open University
    ... Trajectories. 4.7.1 Self Orthogonal family of curves. 4.7.2 Orthogonal trajectories in Cartesian Coordinates. 4.7.3 Orthogonal trajectories in Polar ... cardioids ...
  14. [14]
    MATHEMATICA TUTORIAL, Part 1.2: Orthogonal Trajectories
    Orthogonal Trajectories. Two lines ℓ1andℓ2, with slopes m1 and m2, respectively, are orthogonal or perpendicular if their slopes satisfy the relationship m1m2=− ...
  15. [15]
    [PDF] 2.4 Some Applications 1. Orthogonal Trajectories
    In this section we give some examples of applications of first order differential equations. ... orthogonal trajectories. A procedure for finding a family of ...
  16. [16]
    3.3: The Runge-Kutta Method
    ### Summary of Fourth-Order Runge-Kutta Method
  17. [17]
    [PDF] I. INTRODUCTION The numerical solution of partial differential ...
    Any standard text on the subject of finite-difference methods will provide formulas of ... Several methods for the construction of orthogonal trajectories ...
  18. [18]
    MATHEMATICA TUTORIAL, Part 1.3: Euler Methods
    Orthogonal trajectories · Population models · Pursuit · Additional applications ... differential equation y' = f(x,y) by the finite difference. y′(xn)≈yn+1 ...
  19. [19]
    isogonal trajectory - PlanetMath
    Mar 22, 2013 · We want to determine the isogonal trajectories of this family, ie the curves ι ι intersecting all members of the family under a given angle.Missing: formula | Show results with:formula
  20. [20]
    The isogonal trajectories of a family of parabolas, y^2 = 4ax that ...
    Jan 4, 2020 · Let the trajectories cut the curve of a given family at an angle α, where tan α = k. The slope dy/dx = tanϕ of the tangent to a member of the ...Missing: formula | Show results with:formula
  21. [21]
    example of isogonal trajectory - PlanetMath
    Mar 22, 2013 · An example of an isogonal trajectory is the family of logarithmic spirals, given by r=Ce−φ in polar coordinates.Missing: mathematics | Show results with:mathematics
  22. [22]
    [PDF] The Theory Of Plane Curves,vol.i,ed.2
    POLAR RECIPROCAL AND OTHER DERIVED CURVES. Polar Reciprocal Curves defined. 135 ... ... Polar Reciprocal in homogeneous co-ordinates. 137 ... Page 18. CONTENTS.
  23. [23]
    [PDF] CHAPTER - Purdue Math
    Jan 1, 2010 · The family of orthogonal trajectories for a family of circles that are centered at the origin is another family of circles centered at the ...
  24. [24]
    [PDF] Lecture 23: Curvilinear Coordinates (RHB 8.10)
    If ei · ej = δij, then the ui are orthogonal curvilinear coordinates. For Cartesian coordinates, the scale factors are unity and the unit vectors ei reduce to ...
  25. [25]
    [PDF] METHODS OF ANALYSIS IN CHEMICAL AND NUCLEAR ...
    Dec 16, 2014 · Elliptic Coordinates, 93. Non- homogeneous differential equations ... General Approach to Orthogonal Curvilinear Coordinates x = x(u1 ...
  26. [26]
    [PDF] 3.1 Energy in the field - MIT
    Feb 8, 2005 · A very handy way of putting this is to say that electric field lines are orthogonal to equipotential surfaces. (If you are a hiker and you ...
  27. [27]
    [PDF] ECE 329: Fields and Waves I
    A spherical surface at r represents an equipotential surface. ▫ The equipotential surface is orthogonal to the electric field lines. Equipotential Surfaces.
  28. [28]
    [PDF] Chapter 7 Solution of the Partial Differential Equations
    If the gradients are orthogonal then the equipotential lines and the streamlines are also orthogonal, with the exception of stagnation points where the ...
  29. [29]
    [PDF] The heat and wave equations in 2D and 3D
    We call these lines the “heat flow lines” or the “orthogonal trajectories”, and draw these as dashed lines in the figure below. Note that lines of symmetry ...
  30. [30]
    [PDF] Gravity
    potential, also called equipotential surfaces. The field lines are always orthogonal to the equipotential surfaces, and if they are plotted with constant ...
  31. [31]
    Exact Differential Equations and Harmonic Functions - ADS
    ... Cauchy-Riemann equations and the non-vanishing of the first partial derivatives are sufficient for any two curves to be orthogonal trajectories of each other.