First-order partial differential equation

Last updated

In mathematics, a first-order partial differential equation is a partial differential equation that involves only first derivatives of the unknown function of n variables. The equation takes the form

Contents

Such equations arise in the construction of characteristic surfaces for hyperbolic partial differential equations, in the calculus of variations, in some geometrical problems, and in simple models for gas dynamics whose solution involves the method of characteristics. If a family of solutions of a single first-order partial differential equation can be found, then additional solutions may be obtained by forming envelopes of solutions in that family. In a related procedure, general solutions may be obtained by integrating families of ordinary differential equations.

General solution and complete integral

The general solution to the first order partial differential equation is a solution which contains an arbitrary function. But, the solution to the first order partial differential equations with as many arbitrary constants as the number of independent variables is called the complete integral. The following n-parameter family of solutions

is a complete integral if . [1] The below discussions on the type of integrals are based on the textbook A Treatise on Differential Equations (Chaper IX, 6th edition, 1928) by Andrew Forsyth. [2]

Complete integral

The solutions are described in relatively simple manner in two or three dimensions with which the key concepts are trivially extended to higher dimensions. A general first-order partial differential equation in three dimensions has the form

where Suppose be the complete integral that contains three arbitrary constants . From this we can obtain three relations by differentiation

Along with the complete integral , the above three relations can be used to eliminate three constants and obtain an equation (original partial differential equation) relating . Note that the elimination of constants leading to the partial differential equation need not be unique, i.e., two different equations can result in the same complete integral, for example, elimination of constants from the relation leads to and .

General integral

Once a complete integral is found, a general solution can be constructed from it. The general integral is obtained by making the constants functions of the coordinates, i.e., . These functions are chosen such that the forms of are unaltered so that the elimination process from complete integral can be utilized. Differentiation of the complete integral now provides

in which we require the right-hand side terms of all the three equations to vanish identically so that elimination of from results in the partial differential equation. This requirement can be written more compactly by writing it as

where

is the Jacobian determinant. The condition leads to the general solution. Whenever , then there exists a functional relation between because whenever a determinant is zero, the columns (or rows) are not linearly independent. Take this functional relation to be

Once is found, the problem is solved. From the above relation, we have . By summing the original equations , and we find . Now eliminating from the two equations derived, we obtain

Since and are independent, we require

The above two equations can be used to solve and . Substituting in , we obtain the general integral. Thus a general integral describes a relation between , two known independent functions and an arbitrary function . Note that we have assumed to make the determinant zero, but this is not always needed. The relations or, suffice to make the determinant zero.

Singular integral

Singular integral is obtained when . In this case, elimination of from works if

The three equations can be used to solve the three unknowns . Solution obtained by elimination of this way leads to what are called singular integrals.

Special integral

Usually, most integrals fall into three categories defined above, but it may happen that a solution does not fit into any of three types of integrals mentioned above. These solutions are called special integrals. A relation that satisfies the partial differential equation is said to a special integral if we are unable to determine from the following equations

If we able to determine from the above set of equations, then will turn out to be one of the three integrals described before.

Two dimensional case

The complete integral in two-dimensional space can be written as . The general integral is obtained by eliminating from the following equations

The singular integral if it exists can be obtained by eliminating from the following equations

If a complete integral is not available, solutions may still be obtained by solving a system of ordinary equations. To obtain this system, first note that the PDE determines a cone (analogous to the light cone) at each point: if the PDE is linear in the derivatives of u (it is quasi-linear), then the cone degenerates into a line. In the general case, the pairs (p,q) that satisfy the equation determine a family of planes at a given point:

where

The envelope of these planes is a cone, or a line if the PDE is quasi-linear. The condition for an envelope is

where F is evaluated at , and dp and dq are increments of p and q that satisfy F=0. Hence the generator of the cone is a line with direction

This direction corresponds to the light rays for the wave equation. To integrate differential equations along these directions, we require increments for p and q along the ray. This can be obtained by differentiating the PDE:

Therefore the ray direction in space is

The integration of these equations leads to a ray conoid at each point . General solutions of the PDE can then be obtained from envelopes of such conoids.

Definitions of linear dependence for differential systems

This part can be referred to of Courant's book. [3]

We assume that these equations are independent, i.e., that none of them can be deduced from the other by differentiation and elimination.

Courant, R. & Hilbert, D. (1962), Methods of Mathematical Physics: Partial Differential Equations, II, p.15-18

An equivalent description is given. Two definitions of linear dependence are given for first-order linear partial differential equations.

Where are independent variables; are dependent unknowns; are linear coefficients; and are non-homogeneous items. Let .

Definition I: Given a number field , when there are coefficients (), not all zero, such that ; the Eqs.(*) are linear dependent.

Definition II (differential linear dependence): Given a number field , when there are coefficients (), not all zero, such that , the Eqs.(*) are thought as differential linear dependent. If , this definition degenerates into the definition I.

The div-curl systems, Maxwell's equations, Einstein's equations (with four harmonic coordinates) and Yang-Mills equations (with gauge conditions) are well-determined in definition II, whereas are over-determined in definition I.

Characteristic surfaces for the wave equation

Characteristic surfaces for the wave equation are level surfaces for solutions of the equation

There is little loss of generality if we set : in that case u satisfies

In vector notation, let

A family of solutions with planes as level surfaces is given by

where

If x and x0 are held fixed, the envelope of these solutions is obtained by finding a point on the sphere of radius 1/c where the value of u is stationary. This is true if is parallel to . Hence the envelope has equation

These solutions correspond to spheres whose radius grows or shrinks with velocity c. These are light cones in space-time.

The initial value problem for this equation consists in specifying a level surface S where u=0 for t=0. The solution is obtained by taking the envelope of all the spheres with centers on S, whose radii grow with velocity c. This envelope is obtained by requiring that

This condition will be satisfied if is normal to S. Thus the envelope corresponds to motion with velocity c along each normal to S. This is the Huygens' construction of wave fronts: each point on S emits a spherical wave at time t=0, and the wave front at a later time t is the envelope of these spherical waves. The normals to S are the light rays.

Related Research Articles

<span class="mw-page-title-main">Laplace's equation</span> Second order partial differential equation

In mathematics and physics, Laplace's equation is a second-order partial differential equation named after Pierre-Simon Laplace, who first studied its properties. This is often written as

<span class="mw-page-title-main">Potential flow</span> Velocity field as the gradient of a scalar function

In fluid dynamics, potential flow describes the velocity field as the gradient of a scalar function: the velocity potential. As a result, a potential flow is characterized by an irrotational velocity field, which is a valid approximation for several applications. The irrotationality of a potential flow is due to the curl of the gradient of a scalar always being equal to zero.

In physics, charge conjugation is a transformation that switches all particles with their corresponding antiparticles, thus changing the sign of all charges: not only electric charge but also the charges relevant to other forces. The term C-symmetry is an abbreviation of the phrase "charge conjugation symmetry", and is used in discussions of the symmetry of physical laws under charge-conjugation. Other important discrete symmetries are P-symmetry (parity) and T-symmetry.

<span class="mw-page-title-main">Heat equation</span> Partial differential equation describing the evolution of temperature in a region

In mathematics and physics, the heat equation is a certain partial differential equation. Solutions of the heat equation are sometimes known as caloric functions. The theory of the heat equation was first developed by Joseph Fourier in 1822 for the purpose of modeling how a quantity such as heat diffuses through a given region.

The calculus of variations is a field of mathematical analysis that uses variations, which are small changes in functions and functionals, to find maxima and minima of functionals: mappings from a set of functions to the real numbers. Functionals are often expressed as definite integrals involving functions and their derivatives. Functions that maximize or minimize functionals may be found using the Euler–Lagrange equation of the calculus of variations.

<span class="mw-page-title-main">Green's function</span> Impulse response of an inhomogeneous linear differential operator

In mathematics, a Green's function is the impulse response of an inhomogeneous linear differential operator defined on a domain with specified initial conditions or boundary conditions.

<span class="mw-page-title-main">Korteweg–De Vries equation</span> Mathematical model of waves on a shallow water surface

In mathematics, the Korteweg–De Vries (KdV) equation is a mathematical model of waves on shallow water surfaces. It is particularly notable as the prototypical example of an exactly solvable model, that is, a non-linear partial differential equation whose solutions can be exactly and precisely specified. KdV can be solved by means of the inverse scattering transform. The mathematical theory behind the KdV equation is a topic of active research. The KdV equation was first introduced by Boussinesq and rediscovered by Diederik Korteweg and Gustav de Vries (1895).

<span class="mw-page-title-main">Path integral formulation</span> Formulation of quantum mechanics

The path integral formulation is a description in quantum mechanics that generalizes the action principle of classical mechanics. It replaces the classical notion of a single, unique classical trajectory for a system with a sum, or functional integral, over an infinity of quantum-mechanically possible trajectories to compute a quantum amplitude.

Geometrical optics, or ray optics, is a model of optics that describes light propagation in terms of rays. The ray in geometrical optics is an abstraction useful for approximating the paths along which light propagates under certain circumstances.

<span class="mw-page-title-main">Nonlinear Schrödinger equation</span>

In theoretical physics, the (one-dimensional) nonlinear Schrödinger equation (NLSE) is a nonlinear variation of the Schrödinger equation. It is a classical field equation whose principal applications are to the propagation of light in nonlinear optical fibers and planar waveguides and to Bose–Einstein condensates confined to highly anisotropic cigar-shaped traps, in the mean-field regime. Additionally, the equation appears in the studies of small-amplitude gravity waves on the surface of deep inviscid (zero-viscosity) water; the Langmuir waves in hot plasmas; the propagation of plane-diffracted wave beams in the focusing regions of the ionosphere; the propagation of Davydov's alpha-helix solitons, which are responsible for energy transport along molecular chains; and many others. More generally, the NLSE appears as one of universal equations that describe the evolution of slowly varying packets of quasi-monochromatic waves in weakly nonlinear media that have dispersion. Unlike the linear Schrödinger equation, the NLSE never describes the time evolution of a quantum state. The 1D NLSE is an example of an integrable model.

In differential geometry, the four-gradient is the four-vector analogue of the gradient from vector calculus.

In theoretical physics, Nordström's theory of gravitation was a predecessor of general relativity. Strictly speaking, there were actually two distinct theories proposed by the Finnish theoretical physicist Gunnar Nordström, in 1912 and 1913 respectively. The first was quickly dismissed, but the second became the first known example of a metric theory of gravitation, in which the effects of gravitation are treated entirely in terms of the geometry of a curved spacetime.

<span class="mw-page-title-main">Bring radical</span> Real root of the polynomial x^5+x+a

In algebra, the Bring radical or ultraradical of a real number a is the unique real root of the polynomial

<span class="mw-page-title-main">Routhian mechanics</span> Formulation of classical mechanics

In classical mechanics, Routh's procedure or Routhian mechanics is a hybrid formulation of Lagrangian mechanics and Hamiltonian mechanics developed by Edward John Routh. Correspondingly, the Routhian is the function which replaces both the Lagrangian and Hamiltonian functions. Routhian mechanics is equivalent to Lagrangian mechanics and Hamiltonian mechanics, and introduces no new physics. It offers an alternative way to solve mechanical problems.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

In mathematics, the spectral theory of ordinary differential equations is the part of spectral theory concerned with the determination of the spectrum and eigenfunction expansion associated with a linear ordinary differential equation. In his dissertation Hermann Weyl generalized the classical Sturm–Liouville theory on a finite closed interval to second order differential operators with singularities at the endpoints of the interval, possibly semi-infinite or infinite. Unlike the classical case, the spectrum may no longer consist of just a countable set of eigenvalues, but may also contain a continuous part. In this case the eigenfunction expansion involves an integral over the continuous part with respect to a spectral measure, given by the Titchmarsh–Kodaira formula. The theory was put in its final simplified form for singular differential equations of even degree by Kodaira and others, using von Neumann's spectral theorem. It has had important applications in quantum mechanics, operator theory and harmonic analysis on semisimple Lie groups.

In fluid dynamics, the Oseen equations describe the flow of a viscous and incompressible fluid at small Reynolds numbers, as formulated by Carl Wilhelm Oseen in 1910. Oseen flow is an improved description of these flows, as compared to Stokes flow, with the (partial) inclusion of convective acceleration.

The quantum cylindrical quadrupole is a solution to the Schrödinger equation, where is the reduced Planck constant, is the mass of the particle, is the imaginary unit and is time.

Lagrangian field theory is a formalism in classical field theory. It is the field-theoretic analogue of Lagrangian mechanics. Lagrangian mechanics is used to analyze the motion of a system of discrete particles each with a finite number of degrees of freedom. Lagrangian field theory applies to continua and fields, which have an infinite number of degrees of freedom.

<span class="mw-page-title-main">Causal fermion systems</span> Candidate unified theory of physics

The theory of causal fermion systems is an approach to describe fundamental physics. It provides a unification of the weak, the strong and the electromagnetic forces with gravity at the level of classical field theory. Moreover, it gives quantum mechanics as a limiting case and has revealed close connections to quantum field theory. Therefore, it is a candidate for a unified physical theory. Instead of introducing physical objects on a preexisting spacetime manifold, the general concept is to derive spacetime as well as all the objects therein as secondary objects from the structures of an underlying causal fermion system. This concept also makes it possible to generalize notions of differential geometry to the non-smooth setting. In particular, one can describe situations when spacetime no longer has a manifold structure on the microscopic scale. As a result, the theory of causal fermion systems is a proposal for quantum geometry and an approach to quantum gravity.

References

  1. Garabedian, P. R. (1964). Partial Differential Equations. New York: Wiley. OCLC   527754.
  2. Forsyth, A. R. (1928). A treatise on differential equations.
  3. Courant, R. & Hilbert, D. (1962). Methods of Mathematical Physics: Partial Differential Equations. Vol. II. New York: Wiley-Interscience. ISBN   9783527617241.

Further reading