Fact-checked by Grok 2 weeks ago

Homogeneous differential equation

In mathematics, a homogeneous differential equation is a differential equation where the right-hand side is zero in the linear case or where the defining function is homogeneous of degree zero in the nonlinear first-order case. For linear ordinary differential equations (ODEs), homogeneity means the equation takes the form a_n(x) y^{(n)} + a_{n-1}(x) y^{(n-1)} + \dots + a_1(x) y' + a_0(x) y = 0, with no forcing term depending solely on the independent variable x. This structure ensures that the zero function is a solution and enables the application of the superposition principle: if y_1, y_2, \dots, y_n are linearly independent solutions, the general solution is their linear combination y = c_1 y_1 + c_2 y_2 + \dots + c_n y_n. For nonlinear ODEs, the \frac{dy}{dx} = f(x, y) is homogeneous if f(tx, ty) = f(x, y) for all t \neq 0, meaning f(x, y) = h(y/x) for some h. Such equations can be solved by the substitution v = y/x, which reduces the problem to a separable \frac{dv}{h(v) - v} = \frac{dx}{x}, integrable via standard techniques. Higher-order linear homogeneous ODEs with constant coefficients are solved by assuming solutions of the form y = e^{rx}, leading to a whose roots determine the form of the general solution—exponential for real roots, oscillatory for complex roots, and including factors for repeated roots. and are guaranteed by theorems stating that, given continuous coefficients and initial conditions, a unique solution exists on the interval where coefficients are defined. These equations arise in modeling physical systems without external forces, such as undamped harmonic oscillators or free decay processes.

Introduction

Definitions and Distinctions

In the context of differential equations, an ordinary differential equation (ODE) is an equation that relates a function of a single independent variable to its derivatives with respect to that variable. ODEs are classified by order, where the order is the highest derivative present; first-order ODEs involve only the first derivative, while higher-order ones include derivatives of order two or more. They are further categorized as linear or nonlinear: a linear ODE has the unknown function and its derivatives appearing to the first power with no products or nonlinear functions of them, whereas nonlinear ODEs include such terms. The term "homogeneous" in differential equations has distinct meanings depending on the context, often leading to confusion between nonlinear and linear cases. Central to the nonlinear interpretation is the concept of a : a function f(x, y) is homogeneous of degree k if f(tx, ty) = t^k f(x, y) for all t > 0 and suitable x, y. For a nonlinear ODE of the form \frac{dy}{dx} = f(x, y), it is called homogeneous if f(x, y) is a homogeneous function of degree zero, meaning f(tx, ty) = f(x, y). In contrast, for linear , homogeneity refers to the absence of a nonhomogeneous forcing term. A linear ODE \frac{dy}{dx} + P(x)y = Q(x) is homogeneous if Q(x) = 0, resulting in \frac{dy}{dx} + P(x)y = 0; otherwise, it is nonhomogeneous. For higher-order linear ODEs, such as a_n(x) y^{(n)} + \cdots + a_1(x) y' + a_0(x) y = g(x), homogeneity holds when g(x) = 0. This distinction highlights that nonlinear homogeneous ODEs rely on scaling properties of the right-hand side, while linear ones emphasize the of the equation without external inputs. To illustrate nonhomogeneous counterparts, consider a linear example: \frac{dy}{dx} = -2xy + x^2, where the term x^2 acts as the nonhomogeneous part, contrasting with the homogeneous version \frac{dy}{dx} = -2xy. Similarly, for second-order linear cases, y'' + y' + y = \sin x is nonhomogeneous due to \sin x, unlike y'' + y' + y = 0.

Historical Background

The concept of homogeneous differential equations emerged in the early amid the burgeoning study of differential equations by mathematicians associated with the and their contemporaries. Gabriele Manfredi explored the construction of homogeneous equations in his 1707 treatise De constructione aequationum differentialium, providing early methods for their that influenced subsequent developments. Shortly thereafter, Riccati advanced the field through his work on nonlinear equations that exhibited homogeneous properties, particularly in De usu motus tractorii (1752), where techniques for such forms were detailed using tractional motion. The term "homogeneous" itself was first applied to differential equations by in section 9 of his 1726 article De integrationibus aequationum differentialium, marking a key milestone in classifying equations based on their functional homogeneity. Leonhard Euler built upon these foundations in the mid-18th century, introducing homogeneous functions in the context of integrals and differential equations during his prolific period from approximately 1734 to 1755. In works such as his 1736 contributions to integrating Riccati's equation—a prototypical homogeneous form—Euler employed systematic methods involving series solutions and constant differentials, as outlined in Institutiones Calculi Differentialis (1755). These efforts laid groundwork for handling higher-order cases. By the late , advanced the classification and solution techniques for homogeneous equations, particularly through series expansions and applications to in papers published in the 1770s and his comprehensive Mécanique Analytique (1788), where integration methods for such equations were refined. In the , Carl Gustav Jacobi contributed significantly to solving homogeneous systems via his 1842–1843 lectures on , later published with extensions by contemporaries, emphasizing reductions to partial forms in . Simultaneously, played a pivotal role in formalizing linear homogeneous systems during the early 1800s, developing existence and uniqueness theorems in his 1820s analyses, such as those in Analyse Algébrique (1821), which provided rigorous frameworks for linear ordinary differential equations. By the , the concept of homogeneous differential equations had solidified within modern theory, benefiting from the integration of set-theoretic rigor in to establish general theorems, though without major shifts in the core classification established earlier.

Homogeneous Differential Equations

Form and Properties

A is called homogeneous if it can be expressed in the form \frac{dy}{dx} = f(x, y), where f is a of degree zero, meaning f(tx, ty) = f(x, y) for all t \neq 0. Equivalently, this takes the form \frac{dy}{dx} = g\left(\frac{y}{x}\right), where g is a of the single variable v = y/x. This equivalence follows from the property of homogeneous functions of degree zero, which depend only on the ratio y/x. Key properties of these equations include invariance under scaling of the variables: if (x, y(x)) is a solution curve, then so is (tx, ty(tx)) for any scalar t \neq 0, as the derivative \frac{d(ty)}{d(tx)} = \frac{dy}{dx} remains unchanged and f(tx, ty) = f(x, y). This scaling property relates directly to for homogeneous functions, which states that if f is homogeneous of degree zero, then x \frac{\partial f}{\partial x} + y \frac{\partial f}{\partial y} = 0. Additionally, such equations become separable upon the substitution y = vx, though the explicit process is distinct from solving. Geometrically, the solution curves to homogeneous equations are homogeneous in the plane, meaning they are invariant under radial scaling from the origin, and all nontrivial solutions pass through the origin. For example, the equation \frac{dy}{dx} = \frac{y}{x} has solutions y = cx for constant c, which are straight lines through the origin. In contrast to non-homogeneous first-order equations, which include an additive term independent of the ratio y/x (such as \frac{dy}{dx} = g(y/x) + h(x)), homogeneous equations lack this term, ensuring that the origin is always a solution point and that curves do not shift away from it under scaling.

Solution by Substitution

A first-order homogeneous ordinary differential equation takes the form \frac{dy}{dx} = g\left(\frac{y}{x}\right), where g is a of degree zero. The homogeneity implies that the equation is unchanged under of variables by a positive constant, suggesting that solutions may exhibit scaling invariance along rays from the origin in the xy-plane. This property motivates the v = \frac{y}{x}, which reduces the problem to analyzing the behavior in terms of the v, effectively transforming the equation into one involving the polar in a change to polar coordinates. To apply the substitution, express y = v x, where v = v(x). Differentiating with respect to x using the gives \frac{dy}{dx} = v + x \frac{dv}{dx}. Substituting into the original equation yields v + x \frac{dv}{dx} = g(v). Rearranging terms produces the separable equation x \frac{dv}{dx} = g(v) - v, or equivalently, \frac{dv}{g(v) - v} = \frac{dx}{x}. This separation leverages the homogeneity, as the right side depends only on x and the left on v, allowing without further coupling. Integrating both sides gives \int \frac{dv}{g(v) - v} = \int \frac{dx}{x} + C, where C is the constant of integration. The left integral generally requires specific techniques depending on g(v), such as partial fractions or trigonometric substitutions. Solving the resulting equation for v yields v(x), and back-substituting y = v x provides the implicit or explicit solution for y(x). This method assumes x \neq 0, as division by x is involved; solutions may need separate verification near x = 0. Additionally, singular solutions, such as constant ratios v = k where g(k) = k, must be checked separately, as they satisfy \frac{dv}{dx} = 0 and may represent envelopes or special cases like y = 0 for certain g. Consider the example \frac{dy}{dx} = \frac{x + y}{x - y}. Here, g(v) = \frac{1 + v}{1 - v}, confirming homogeneity. Substitute v = \frac{y}{x}, so \frac{dy}{dx} = v + x \frac{dv}{dx}, leading to v + x \frac{dv}{dx} = \frac{1 + v}{1 - v}. Simplifying gives x \frac{dv}{dx} = \frac{1 + v^2}{1 - v}, or \frac{1 - v}{1 + v^2} dv = \frac{dx}{x}. Integrating the left side: \int \frac{1 - v}{1 + v^2} dv = \int \frac{dv}{1 + v^2} - \int \frac{v \, dv}{1 + v^2} = \arctan v - \frac{1}{2} \ln(1 + v^2). The right side integrates to \ln |x|\ + C. Thus, \arctan v - \frac{1}{2} \ln(1 + v^2) = \ln |x| + C. Back-substituting v = \frac{y}{x} yields the implicit solution \arctan \left( \frac{y}{x} \right) - \frac{1}{2} \ln \left(1 + \left( \frac{y}{x} \right)^2 \right) = \ln |x| + C. An alternative explicit form for v can involve trigonometric identities, such as expressing the solution using \tan(\theta/2) substitutions during integration, leading to v = \frac{1 + \tan(\theta/2)}{1 - \tan(\theta/2)} for certain parameterizations, though the logarithmic-arctangent form is more direct.

Homogeneous Linear Ordinary Differential Equations

General Theory

A homogeneous linear differential equation of order n is expressed as a_n(x) y^{(n)}(x) + a_{n-1}(x) y^{(n-1)}(x) + \cdots + a_1(x) y'(x) + a_0(x) y(x) = 0, where a_n(x) \neq 0 and the coefficients a_i(x) are continuous functions on an I. This equation can be compactly written in operator form as L(y) = 0, where L is a linear of order n acting on the function y. Under the assumption that the coefficients a_i(x) are continuous on an open interval containing the initial point, the for such an equation possesses a unique on that interval. For first-order equations, this follows from the Picard-Lindelöf theorem, which guarantees existence and uniqueness when the functions involved satisfy Lipschitz conditions; this result extends to higher-order linear equations by reduction to a system of first-order equations. Additionally, the holds: if y_1 and y_2 are solutions to L(y) = 0, then any c_1 y_1 + c_2 y_2 (with constants c_1, c_2) is also a , reflecting the of the L. This generalizes to n solutions for an nth-order equation. The set of all solutions to the homogeneous equation L(y) = 0 forms a vector space over the real or complex numbers, with dimension exactly n. A basis for this solution space consists of a fundamental set of n linearly independent solutions \{y_1, y_2, \dots, y_n\}, such that the general solution is y(x) = c_1 y_1(x) + c_2 y_2(x) + \cdots + c_n y_n(x), where the c_i are arbitrary constants determined by initial conditions. In the context of nonhomogeneous linear equations L(y) = g(x), the solutions to the associated homogeneous equation provide the complementary function, which, when added to a particular solution, yields the general solution; this structure is essential in methods like variation of parameters. For a first-order homogeneous linear equation y' + p(x) y = 0, where p(x) is continuous, the general solution is y(x) = C \exp\left( -\int p(x) \, dx \right), with C an arbitrary constant; this illustrates the vector space structure, as scalar multiples span the one-dimensional solution space.

Constant Coefficient Cases

Homogeneous linear ordinary differential equations with constant coefficients take the form a_n y^{(n)} + a_{n-1} y^{(n-1)} + \cdots + a_1 y' + a_0 y = 0, where the a_i are constants and a_n \neq 0. The standard method to solve these equations involves assuming a solution of the form y = e^{rx}, where r is a constant to be determined. Substituting this assumed form into the differential equation yields the characteristic equation a_n r^n + a_{n-1} r^{n-1} + \cdots + a_1 r + a_0 = 0, a polynomial equation in r whose roots determine the form of the general solution. The nature of the roots of the characteristic equation dictates the structure of the solution. For an nth-order equation, there are n roots (counting multiplicities), which may be real, complex, or repeated. If all roots r_1, r_2, \dots, r_n are real and distinct, the general solution is the linear combination y(x) = \sum_{i=1}^n C_i e^{r_i x}, where the C_i are arbitrary constants determined by initial conditions. This follows from the superposition principle, which guarantees that linear combinations of solutions are also solutions. When the characteristic equation has repeated roots, the solutions must be modified to ensure linear independence. For a root r of multiplicity k, the corresponding terms in the general solution include e^{rx}, x e^{rx}, \dots, x^{k-1} e^{rx}. For example, consider the second-order equation y'' - 12y' + 36y = 0. The is r^2 - 12r + 36 = 0, or (r - 6)^2 = 0, yielding a repeated r = 6. The general is y(x) = (C_1 + C_2 x) e^{6x}. Complex roots occur in conjugate pairs for equations with real coefficients. If the roots are \alpha \pm i\beta (with \beta \neq 0), the corresponding real solutions are e^{\alpha x} \cos(\beta x) and e^{\alpha x} \sin(\beta x). For a second-order equation, the general solution is then y(x) = e^{\alpha x} (A \cos(\beta x) + B \sin(\beta x)). A classic example is the damped equation y'' + 2y' + 2y = 0, modeling lightly damped vibrations. The r^2 + 2r + 2 = 0 has roots r = -1 \pm i, so the general solution is y(x) = e^{-x} (A \cos x + B \sin x). For distinct real roots, consider the second-order equation y'' - 3y' + 2y = 0. The r^2 - 3r + 2 = 0 factors as (r - 1)(r - 2) = 0, giving r = 1 and r = 2. The general is y(x) = C_1 e^{x} + C_2 e^{2x}. In all cases, the functions forming the general solution are linearly independent, which can be verified using the determinant. For two solutions y_1 and y_2, the is W(y_1, y_2) = y_1 y_2' - y_2 y_1'; if W \neq 0 at some point, the solutions are linearly independent over the interval. For the exponential solutions above, the is nonzero, confirming the basis for the solution space.

Variable Coefficient Cases

Homogeneous linear ordinary differential equations with variable coefficients present significant challenges compared to their constant coefficient counterparts, as no general closed-form solution exists in terms of elementary functions for arbitrary functions. Instead, solutions often rely on , expansions, or numerical approximations, particularly when the coefficients introduce singularities or complex behavior. One key approach for solving these equations around regular singular points is the method of Frobenius, which extends the power series method by assuming solutions of the form y = x^r \sum_{k=0}^{\infty} a_k x^k, where r is determined by an indicial equation derived from the lowest-order terms. This technique yields solutions valid in intervals excluding the singular points, provided the coefficients satisfy analyticity conditions near the point of expansion. A notable class amenable to exact solutions is the Cauchy-Euler equation, given by x^2 y'' + a x y' + b y = 0, where a and b are constants. Substituting y = x^m (for x > 0) transforms the equation into the algebraic m^2 + (a - 1)m + b = 0. The roots m_1 and m_2 determine the general solution: if distinct, y = c_1 x^{m_1} + c_2 x^{m_2}; if repeated (m_1 = m_2 = m), then y = (c_1 + c_2 \ln x) x^m. For complex roots m = \alpha \pm i \beta, the solutions involve x^\alpha \cos(\beta \ln x) and x^\alpha \sin(\beta \ln x). Consider the example y'' - \frac{1}{x} y' + \frac{1}{x^2} y = 0, which, upon multiplying by x^2, becomes the Cauchy-Euler form x^2 y'' - x y' + y = 0 with a = -1 and b = 1. The characteristic equation is m^2 - 2m + 1 = 0, yielding the repeated root m = 1, so the general solution is y = (c_1 + c_2 \ln x) x. Certain equations with variable coefficients admit solutions in terms of special functions. The Bessel equation, x^2 y'' + x y' + (x^2 - \nu^2) y = 0, has solutions involving the Bessel functions of the first kind J_\nu(x) and second kind Y_\nu(x), which are linearly independent and form the general solution. Similarly, the Legendre equation, (1 - x^2) y'' - 2x y' + n(n+1) y = 0, for integer n \geq 0 has polynomial solutions known as Legendre polynomials P_n(x), alongside a second solution Q_n(x) that is non-polynomial. These functions arise in applications such as quantum mechanics and heat conduction. When closed-form expressions via are unavailable, numerical methods or asymptotic approximations are typically employed to obtain solutions, especially for equations with irregular singular points or non-standard coefficients.

References

  1. [1]
    Homogeneous Ordinary Differential Equation - Wolfram MathWorld
    A linear ordinary differential equation of order n is said to be homogeneous if it is of the form a_n(x)y^((n))+a_(n-1)(x)y^((n-1))+...+a_1(x)y
  2. [2]
    [PDF] Homogeneous Equations A function f(x, y) is said to be ...
    A first-order differential equation y. /. (x) = f(x, y) is called homogeneous if f(x, y) is homogeneous of degree 0. Caution In the context of linear ...<|control11|><|separator|>
  3. [3]
    [PDF] Homogeneous Linear Equations — The Big Theorems - UAH
    We are discussing general homogeneous linear differential equations. If the equation is of second order, it will be written as ay′′ + by′ + cy = 0 .
  4. [4]
    Ordinary Differential Equation -- from Wolfram MathWorld
    An ordinary differential equation (frequently called an "ODE," "diff eq," or "diffy Q") is an equality involving a function and its derivatives.
  5. [5]
    Homogeneous Function -- from Wolfram MathWorld
    A homogeneous function is a function that satisfies f(tx,ty)=t^nf(x,y) for a fixed n. Means, the Weierstrass elliptic function, and triangle center functions ...
  6. [6]
    [PDF] The History of Differential Equations, 1670–1950
    Differential equations have been a major branch of pure and applied mathematics since their inauguration in the mid 17th century. While their history has ...
  7. [7]
    De integraionibus aequationum differentialium - - aupiff
    Sep 11, 2020 · Author Johann Beroulli, June 1726. Bernoulli 1 Bernoulli 1_2. 1. When any first degree differential equation has been reduced to p d x = q d y ...
  8. [8]
    Joseph-Louis Lagrange (1736 - 1813) - Biography - MacTutor
    In papers which were published in the third volume, Lagrange studied the integration of differential equations and made various applications to topics such as ...
  9. [9]
    Memorial speech for Carl Gustav Jacob Jacobi - MacTutor
    His great winter lecture in 1842/43 on the integration of the differential equations, which, with Borchardts postscript and minor changes, was published as ...Missing: 19th | Show results with:19th
  10. [10]
    Differential Equations - Substitutions - Pauls Online Math Notes
    Nov 16, 2022 · y′=F(yx) y ′ = F ( y x ). First order differential equations that can be written in this form are called homogeneous differential equations.
  11. [11]
    [PDF] Chapter 1 Introduction and first-order equations
    In this introductory chapter we define ordinary differential equations, give examples showing how they are used and show how to find solutions of some.
  12. [12]
    [PDF] Chapter 14 First Order Differential Equations
    A differential equation is homogeneous if it has no terms that are functions of the independent variable alone. Thus an inhomogeneous equation is one in which ...
  13. [13]
    [PDF] First order differential equations - Purdue Math
    First order differential equations have the general form dy/dt = f(t,y) or dy/dx = f(x,y). Linear forms include dy/dt + p(t)y = g(t) or P(t) dy/dt + Q(t)y = G( ...Missing: ordinary | Show results with:ordinary<|control11|><|separator|>
  14. [14]
    [PDF] Using Substitution Homogeneous and Bernoulli Equations
    result, this is a homogeneous differential equation. We will substitute y = ux. By the product rule, . Making these substitutions we obtain. Now this ...Missing: method | Show results with:method
  15. [15]
    [PDF] HIGHER-ORDER LINEAR ORDINARY DIFFERENTIAL EQUATIONS I
    Aug 21, 2012 · When f(t) = 0 the equation is said to be homogeneous; otherwise it is said to be nonhomogeneous. Definition. We say that y = Y (t) is a solution ...
  16. [16]
    [PDF] General Theory of nth Order Linear Differential Equations
    The amazing Principle of Superposition (only valid for homogeneous (linear) diff.eqs.!) that we already know is applicable for first and second-order linear ...<|control11|><|separator|>
  17. [17]
    [PDF] Existence and uniqueness theorem for nth order linear ODEs.
    Existence and uniqueness theorem for nth order linear ODEs. The general nth order linear ODE if there is given by. (1) dny dtn. + an−1(t) dn−1y dtn−1. + ...
  18. [18]
    [PDF] The Existence and Uniqueness Theorem for Linear Systems
    Theorem 2 Existence and uniqueness theorem for linear systems. If the entries of the square matrix A(t) are continuous on an open interval I containing t0, then ...Missing: ordinary | Show results with:ordinary
  19. [19]
    [PDF] SECOND ORDER LINEAR DIFFERENTIAL EQUATIONS
    Theorem 4 (Superposition Principle). (i) If y1 and y2 are solutions of the homogeneous equation, then so is any linear com- bination y = c1y1 +c2y2 (c1,c2 ∈ R).
  20. [20]
    [PDF] Linear ODE
    The dimension of the vector space of solutions of (nLH) (a subspace of Cn(I)) is n, i.e., dim N(L) = n, where N(L) denotes the null space of L : Cn(I) → C(I).
  21. [21]
    [PDF] Higher Order Linear Differential Equations - Penn Math
    Jul 28, 2015 · It is called the solution space. The dimension of the solutions space is n. Being a vector space, the solution space has a basis. 1y1(x),y2(x) ...
  22. [22]
    [PDF] Theory of higher order differential equations - Purdue Math
    Definition 2. Samy T. Higher order. Differential equations. 7 / 52. Page 8. Homogeneous equations. General homogeneous linear equation: ... Complementary function ...
  23. [23]
    Linear Differential Equations - Pauls Online Math Notes
    Aug 1, 2024 · In this section we solve linear first order differential equations, i.e. differential equations in the form y' + p(t) y = g(t).
  24. [24]
    2.2: Constant Coefficient Equations - Mathematics LibreTexts
    May 23, 2024 · Solution. The characteristic equation for this problem is r 2 − r − 6 = 0 . The roots of this equation are found as r = − 2 , 3 .Example 2 . 2 . 1 · Example 2 . 2 . 2 · Example 2 . 2 . 4 · Example 2 . 2 . 5
  25. [25]
    9.2: Higher Order Constant Coefficient Homogeneous Equations
    Jul 20, 2020 · In this section we consider the homogeneous constant coefficient equation of n-th order.
  26. [26]
    3.3: Repeated Roots and Reduction of Order - Mathematics LibreTexts
    Oct 19, 2025 · Solution. The characteristic equation is. r 2 − 12 ⁢ r + 36 = 0. or. ( r − 6 ) 2 = 0. We have only the root r = 6 which gives the solution.Example 3 . 3 . 1 : repeated roots · Definition: General Solution · Example 3 . 3 . 2
  27. [27]
    17.3: Applications of Second-Order Differential Equations
    Sep 1, 2025 · Solve a second-order differential equation representing damped simple harmonic motion. ... constant-coefficient differential equation. This ...
  28. [28]
    3.6: Linear Independence and the Wronskian - Math LibreTexts
    Oct 19, 2025 · Hence, if the Wronskian is nonzero at some , only the trivial solution exists. Hence they are linearly independent.Missing: coefficient | Show results with:coefficient
  29. [29]
    7.3: Singular Points and the Method of Frobenius - Math LibreTexts
    Feb 23, 2025 · While behavior of ODEs at singular points is more complicated, certain singular points are not especially difficult to solve.Examples · Method of Frobenius · Example \(\PageIndex{3... · Bessel Functions
  30. [30]
    Variable Coefficient ODEs
    In this section of the course, we shall consider a number of strategies for solving variable coefficient ODEs like the one presented above.
  31. [31]
    [PDF] 12d. Regular Singular Points
    Oct 22, 2012 · The method we study here was discovered by the German mathematician. F. G. Frobenius in the 1870's and is therefore often called the Frobenius.
  32. [32]
    Differential Equations - Euler Equations - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will discuss how to solve Euler's differential equation, ax^2y'' + bxy' +cy = 0. Note that while this does not involve a ...
  33. [33]
    2.5: Cauchy-Euler Equations - Mathematics LibreTexts
    May 23, 2024 · Therefore, the general solution is found as y ⁡ ( x ) = ( c 1 + c 2 ⁢ ln ⁡ | x | ) ⁢ x r . For one root, r 1 = r 2 = r , the general solution is ...
  34. [34]
    [PDF] 7.4 Cauchy-Euler Equation
    Solution: The characteristic equation 2r(r − 1) + 4r + 3 = 0 can be obtained as follows: 2x2y00 + 4xy0 + 3y = 0. Given differential equation. Page 3. 552.
  35. [35]
    Bessel Differential Equation - an overview | ScienceDirect Topics
    The linear combination of the Bessel functions of the first and second kinds represents a complete solution of the Bessel equation: y ( x ) = C 1 J α ( x ) + C ...
  36. [36]
    Legendre Differential Equation -- from Wolfram MathWorld
    The Legendre differential equation is the second-order ordinary differential equation (1-x^2)(d^2y)/(dx^2)-2x(dy)/(dx)+l(l+1)y=0, which can be rewritten d/
  37. [37]
    7.2: Legendre Polynomials - Mathematics LibreTexts
    May 23, 2024 · These polynomial solutions are the Legendre polynomials, which we designate as y ⁡ ( x ) = P n ⁡ ( x ) . Furthermore, for n an even integer, P n ...