This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy. Close this notification
Paper The following article is Open access

Derivation of Jacobian formula with Dirac delta function

, , and

Published 10 March 2021 © 2021 The Author(s). Published on behalf of the European Physical Society by IOP Publishing Ltd
, , Citation Dohyun Kim et al 2021 Eur. J. Phys. 42 035006 DOI 10.1088/1361-6404/abdca9

0143-0807/42/3/035006

Abstract

We demonstrate how to make the coordinate transformation or change of variables from Cartesian coordinates to curvilinear coordinates by making use of a convolution of a function with Dirac delta functions whose arguments are determined by the transformation functions between the two coordinate systems. By integrating out an original coordinate with a Dirac delta function, we replace the original coordinate with a new coordinate in a systematic way. A recursive use of Dirac delta functions allows the coordinate transformation successively. After replacing every original coordinate into a new curvilinear coordinate, we find that the resultant Jacobian of the corresponding coordinate transformation is automatically obtained in a completely algebraic way. In order to provide insights on this method, we present a few examples of evaluating the Jacobian explicitly without resort to the known general formula.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

A coordinate transformation or change of variables from a coordinate system to another in multi-dimensional integrals has widely been applied to a variety of fields in mathematics and physics. This transformation always involves a factor called the Jacobian, which is the determinant of the Jacobian matrix. The matrix elements of the Jacobian matrix are the first-order partial derivatives of the new coordinates with respect to the original coordinates. The formula for the change of variables from n-dimensional variables x1, x2, ..., xn to q1, q2, ..., qn is expressed in terms of the Jacobian $\mathcal{J}$:

Equation (1)

where the integrand f(x1, ..., xn ) is a function of the independent variables x1, ..., xn and F(q1, ..., qn ) = f[x1(q1, ..., qn ), ..., xn (q1, ..., qn )]. In general, the variables x1, ..., xn can be treated as the Cartesian coordinates of an n-dimensional Euclidean space and the variables q1, q2, ..., qn form a set of curvilinear coordinates representing the same Euclidean space. We assume that each curvilinear coordinate qi is uniquely defined by the Cartesian coordinates: qi = qi (x1, ..., xn ). We also assume that the inverse transformation is uniquely defined as xi = xi (q1, ..., qn ). Then the Jacobian $\mathcal{J}$ can be expressed as

Equation (2)

where $\mathcal{D}\mathcal{e}\mathcal{t}$ stands for the determinant.

In physics, the Jacobian appears frequently when one makes the change of variables between Cartesian and curvilinear coordinates in various physical quantities involving surface or volume integrals. However, in most physics textbook including classical mechanics and electromagnetism usually abstract descriptions are provided. In many textbooks of calculus or mathematical physics [1], the Jacobian formula is derived in the following way: first, one transforms the multi-variable differential volume by applying the change of variables. Next, one imposes a geometrical argument that the infinitesimal volume is invariant under the transformation [2]. The invariance of the volume can also be confirmed by applying Green's theorem to show that ${\int }_{S}\mathrm{d}{x}_{1}\enspace \mathrm{d}{x}_{2}{\enspace =\int }_{T}\mathcal{J}\mathrm{d}{q}_{1}\enspace \mathrm{d}{q}_{2}$, where S is a rectangular region in the x1 x2 plane and T is the corresponding region [3]. Although experienced teachers or researchers may follow the abstract logic in this kind of Jacobian derivation without difficulties, the concept is rather unclear or less intuitive to undergraduate physics-major students who are not familiar with advanced mathematical concepts of multi-variable calculus.

The Dirac delta function δ(x) is not a well-defined function but a distribution defined only through integration: ${\int }_{-\infty }^{\infty }\mathrm{d}x\delta \left(x\right)=1$. For any smooth function f(x), ${\int }_{-\infty }^{\infty }\mathrm{d}x\delta \left(x-y\right)f\hspace{-1pt}\left(x\right)=f\hspace{-1pt}\left(y\right)$. This property can be applied to changing the integration variable. Recently, some of us introduced an alternative proof of Cramer's rule by making use of Dirac delta functions [4]. It turns out that the method with a convolution of a coordinate vector with Dirac delta functions provides a systematic way to change integration variables from original coordinates to new coordinates. This change of variables enables us to reproduce Cramer's rule.

The Dirac delta function technique exploited in the derivation of Cramer's rule in reference [4] can be immediately applicable to the evaluation of the Jacobian for the coordinate transformation or change of variables after replacing the coordinate vector with an arbitrary function. However, the direct application of the approach in reference [4] is limited to a linear transformation between two coordinate systems because Cramer's rule applies only to the system of linear equations. Thus the method is not applicable to the transformation involving a set of curvilinear coordinates which are frequently used in physics.

The main goal of this work is to present a more intuitive derivation of the Jacobian involving any coordinate transformation and to demonstrate how it works with heuristic examples. We develop an alternative derivation of the Jacobian formula as well as the coordinate transformation by convolving a function with Dirac delta functions. Our derivation relies only on the direct integration of Dirac delta functions whose arguments involve the coordinate transformation rules. Hence, we expect that students who are familiar with the Dirac delta function can compute the Jacobian formula in any coordinate transformations by themselves without referring to a reference. The approach that we present in this paper is quite straightforward and requires mostly algebraic computation skills.

In this paper, we derive the Jacobian for a coordinate transformation or change of variables in the case of a non-linear transformation by convolving an arbitrary function with Dirac delta functions. While the basic strategy to perform the coordinate transformation is similar to that employed in reference [4], the integration of the original coordinates is rather involved because of the non-linear property of the transformation functions between the two coordinate systems. The integration of the original coordinates can be carried out by making use of Dirac delta functions. An extra factor containing partial derivatives of corresponding coordinate variables appears in front of the original integrand after that multiple integration. It turns out that the extra factor can be evaluated by making use of the chain rule of partial derivatives. A recursive use of Dirac delta functions enables us to achieve the coordinate transformation successively. After replacing every coordinate with new coordinates, we identify the resultant extra factor in the integrand with the Jacobian for the coordinate transformation or change of variables.

The derivation of the Jacobian formula presented in this paper is new to our best knowledge. It is remarkable that our derivation is free of borrowing abstract and advanced mathematical concepts unlike the other derivations available. Instead, we exploit a simple concept of integration of the one-dimensional Dirac delta function repeatedly in combination with a purely algebraic manipulation in reorganizing the extra factor by applying chain rules. This intuitive and systematic approach is expected to be pedagogically useful in upper-level mathematics or physics courses in practice of the recursive use of both the Dirac delta function and the chain rule of partial derivatives.

This paper is organized as follows: in section 2, we introduce notations that are frequently used in the rest of the paper, make a rough sketch of our strategy and present a formal derivation of the Jacobian factor for the coordinate transformation from the n-dimensional Cartesian coordinates to a set of curvilinear coordinates. Section 3 is devoted to the explicit evaluation of the Jacobians for a few examples. Conclusions are given in section 4 and a rigorous derivation of the chain rule for partial derivatives is given in appendices.

2. Derivation of the Jacobian

2.1. Strategy and notation

In this subsection, we present our strategy to derive the Jacobian for a coordinate transformation or change of variables from the Cartesian coordinates xi to the curvilinear coordinates qi with transformation functions

Equation (3)

for i = 1 through n, where n is a positive integer. We assume that the two sets of coordinates describe a single point uniquely and, therefore, the two sets of coordinates have a one-to-one correspondence although the transformation is in general non-linear. Thus the transformation in equation (3) is invertible: the inverse transformation from the curvilinear coordinates to the Cartesian coordinates exists. If the curvilinear coordinates are a linear combination of the Cartesian coordinates, then the linear transformation is invertible if the transformation matrix is non-singular: the determinant of the matrix is not vanishing. Then, the inverse transformation can be written as

Equation (4)

The basic strategy to derive the Jacobian with Dirac delta functions is the same as that for the derivation of Cramer's rule for a partial set of a coordinate transformation or change of variables given in reference [4]. One could immediately apply the approach in reference [4] to find the Jacobian as long as the transformation (3) is linear. In general, the transformation functions (3) are non-linear. Here we develop a generalized version of the approach in reference [4] in order to consider arbitrary curvilinear coordinates.

We define an n-dimensional integral In ,

Equation (5)

where dn x ≡ dx1...dxn is the n-dimensional differential volume element and the integrand f(x1, ..., xn ) is an arbitrary function of the Cartesian coordinates. We assume that every Cartesian coordinate is integrated over the region (−, ). We define the unity

Equation (6)

for i = 1 through n. Multiplying the unity ${\mathbb{1}}_{i}$ to the integral In , one can integrate out the integration variable xi by making use of the Dirac delta function keeping the qi integral unevaluated. By applying this process to In recursively from i = 1 through n, we complete the change of the integration variables from the Cartesian coordinates to the curvilinear coordinates. While we have suppressed the bounds of the integration for the curvilinear coordinate qi 's in equation (6), the curvilinear coordinates are assumed to be integrated over the entire region to cover the whole Euclidean space represented by the Cartesian coordinates by a single time.

We first compute ${\mathbb{1}}_{1}{\times}{I}_{n}$:

Equation (7)

By definition, we integrate over x1 by making use of the delta function δ[q1q1(x1, ..., xn )] keeping the q1 integral unevaluated:

Equation (8)

where x1 is expressed in terms of q1 and xj 's for j = 2 through n satisfying the condition that the argument of the delta function vanishes, q1q1(x1, ..., xn ) = 0. Because the explicit forms of f(xi ) and xi vary depending on the integration step, we adopt the notation f(q1, x2, ..., xn ) after the x1 integration. We will present more detailed explanations for this notation in the later part of this subsection. The extra factor ${\mathcal{G}}_{1}$ is the remnant of the integration of the delta function and its explicit form will be given later in this paper.

Next, we multiply ${\mathbb{1}}_{2}$ to In in equation (8) to find that

Equation (9)

where every x1 in the argument of the delta function as well as the integrand function is replaced with the expression in terms of q1 and xj 's for j = 2 through n. Performing the integration over x2 by making use of the delta function, we find

Equation (10)

After the x2 integration every x2 in the integrand is expressed in terms of q1, q2 and xj 's for j = 3 through n. Again, ${\mathcal{G}}_{2}$ is the remnant of the integration of the delta functions.

In this way, we integrate over xk for k = 1 through n successively. Finally, after the integration over xn we find that In reduces into the n-dimensional multiple integral over the curvilinear coordinates qi for i = 1 through n only:

Equation (11)

where ${\mathcal{G}}_{n}$ is the remnant of the integration of the delta functions. At this stage, the integrand acquires an additional factor 1/${\mathcal{G}}_{n}$ in front of the original integrand f in equation (5). This extra factor is identified with the Jacobian.

In an intermediate step, for example, where the integration over xj (1 ⩽ jn) is carried out, x1, ..., xj in the integrand must be replaced with the expressions in terms of q1, ..., qj and xj+1, ..., xn as

Equation (12)

for k = 1 through j. Each xk in equation (12) is determined by the condition that the argument of the corresponding Dirac delta function vanishes:

Equation (13)

One must keep in mind that the explicit form of each xk varies depending on the integration step as is displayed in equation (12). Thus, one must distinguish, for example, xk (q1, ..., qj , xj+1, ..., xn ) from xk (q1, ..., qj−1, xj , ..., xn ), where the former is the expression after the integration over xj and the latter is that after the integration over xj−1. This notation is also applied to the original integrand function f and the extra factor ${\mathcal{G}}_{i}$.

As an explicit example, we consider the coordinate transformation between the two-dimensional Cartesian coordinates and the polar coordinates with the transformation functions

Equation (14)

and the inverse transformation functions

Equation (15)

In an intermediate step, y can be expressed in terms of θ and x as y(θ, x) = x tan θ, which must be distinguished from y(θ, r) = r sin θ.

Since the dependence of a coordinate variable varies according to the integration step, one must take special care of dealing with their partial derivatives. In order to avoid such an ambiguity, we introduce a notation for the partial derivative with subscripts of variables that are held constant. In the above two-dimensional transformation, the partial derivative of θ with respect to y holding x fixed is denoted by

Equation (16)

while the partial derivative of θ with respect to y holding r fixed is represented by

Equation (17)

It is apparent from equations (16) and (17) that

Equation (18)

In a general case, for a variable qj (q1, ..., qi , xi+1, ..., xn ), we denote the partial derivative of qj with respect to xa holding q1, ..., qi , xi+1, ..., xn fixed with the subscript (i) as

Equation (19)

where i + 1 ⩽ jn and i + 1 ⩽ an. Similarly, for a variable xk (q1, ..., qi , xi+1, ..., xn ), the partial derivative of xk with respect to qb holding q1, ..., qi , xi+1, ..., xn fixed is denoted by

Equation (20)

where 1 ⩽ ki and 1 ⩽ bi. We denote the partial derivative of x with respect to xc holding q1, ..., qi , xi+1, ..., xn fixed by

Equation (21)

where 1 ⩽ i and i + 1 ⩽ cn. Finally, we express the partial derivative of q with respect to xc holding x1, ..., xn without subscript:

Equation (22)

where 1 ⩽ n and 1 ⩽ cn.

In the coordinate transformation from (x1, ..., xn ) to (q1, ..., qn ), we define the function ${\mathcal{G}}_{k}$, which is relevant for a partial set of integral variables corresponding to a transformation from (x1, ..., xk ) to (q1, ..., qk ), as

Equation (23)

for k = 1 through n. Note that ${\mathcal{G}}_{n}$ is the inverse of the Jacobian ${\mathcal{J}}_{n}$ which is defined by

Equation (24)

2.2. 2-dimensional case

In this subsection, we consider a two-dimensional integral I2 for the integration of an arbitrary function f(x1, x2):

Equation (25)

The Cartesian coordinates x1 and x2 are transformed into curvilinear coordinates q1 and q2 with the transformation relations

Equation (26)

which are assumed to be invertible and non-singular. Thus, x1 and x2 can be expressed in terms of q1 and q2:

Equation (27)

We multiply the unities

Equation (28)

to I2 for i = 1, 2, sequentially. First, we multiply ${\mathbb{1}}_{1}$ in equation (28) to I2. Then, I2 can be expressed as

Equation (29)

The integration over x1 can be carried out by making use of the delta function for q1 as

Equation (30)

which leads to

Equation (31)

After the integration over x1, every x1 on the right-hand side of equation (30) and δ[q2q2(x1, x2)]f(x1, x2) in equation (28) must be replaced with x1(q1, x2) satisfying the condition that the argument of the delta function vanishes, q1q1(x1, x2) = 0. Then the delta function for q2 and f(x1, x2) can be expressed as δ[q2q2(q1, x2)] and f(q1, x2), respectively.

After multiplying ${\mathbb{1}}_{2}$ in equation (28) to I2 in equation (31), the integration over x2 can be performed by making use of the remaining delta function for q2 as

Equation (32)

After the integration over x2, every x2 on the right-hand sides of equations (30) and (32) and in f(q1, x2) is replaced with x2(q1, q2), which can be obtained from the condition that the argument of the delta function vanishes, q2q2(q1, x2) = 0. Then, we can express x1 in terms of q1 and q2 by replacing x2 with x2(q1, q2) in x1(q1, x2). We have obtained the last equality of equation (32) by making use of the chain rule for the partial derivatives. A rigorous proof of this formula is given in equation (A5) of appendix A.

After both x1 and x2 are integrated out, the integral I2 reduces into

Equation (33)

where the Jacobian for the change of variables is identified as $\mathcal{J}{=\mathcal{J}}_{2}=1{/\mathcal{G}}_{2}$. This completes the proof of the Jacobian for a two-dimensional coordinate transformation or change of variables.

2.3. 3-dimensional case

In this subsection, we extend the results in the previous subsection to the three-dimensional case. This is a special case of the n-dimensional coordinate transformation or change of variables, which we will prove in the next subsection. However, it is worthwhile to prove the three-dimensional case in detail for a pedagogical purpose.

We consider a three-dimensional integral I3 for an arbitrary function f(x1, x2, x3):

Equation (34)

The Cartesian coordinates xi for i = 1 through 3 are transformed into the curvilinear coordinates qi 's as

Equation (35)

The inverse transformation can be expressed as

Equation (36)

We carry out the change of variables by multiplying the unites

Equation (37)

for i = 1 through 3 to I3, sequentially. First, after multiplying ${\mathbb{1}}_{1}$ in equation (37) to I3, we find that I3 can be expressed as

Equation (38)

Similarly to the two-dimensional case, we perform the integration over x1 by making use of the Dirac delta function for q1 as

Equation (39)

which leads to

Equation (40)

After the integration over x1, every x1 in the integrand and remaining delta functions is replaced with x1(q1, x2, x3) satisfying the condition that the argument of the Dirac delta function vanishes, q1q1(x1, x2, x3) = 0. Then the delta function for q2 in equation (37) can be expressed as δ[q2q2(q1, x2, x3)].

After multiplying ${\mathbb{1}}_{2}$ in equation (37) to I3 in equation (40), the integration over x2 can be carried out by making use of the Dirac delta function for q2 as

Equation (41)

where the last equality comes from equation (B5). Then, equation (41) is expressed as

Equation (42)

x2(q1, q2, x3) is determined from the condition that the argument of the Dirac delta function vanishes, q2q2(q1, x2, x3) = 0. Substituting x2(q1, q2, x3) into x1(q1, x2, x3), we obtain x1 = x1(q1, q2, x3) and the argument of the Dirac delta function for q3 in equation (37) is expressed as δ[q3q3(q1, q2, x3)].

Finally, after multiplying ${\mathbb{1}}_{3}$ in equation (37) to I3 in equation (42), we integrate over x3 by taking into account the delta function for q3 as

Equation (43)

where the last equality comes from equation (B11). After the integration over x3, every x3 in the integrand and the right-hand sides of equations (39), (41) and (43) is replaced with x3(q1, q2, q3) which is determined from the condition that the argument of the delta function vanishes, q3q3(q1, q2, x3) = 0. Substituting x3(q1, q2, q3) into x1(q1, q2, x3) and x2(q1, q2, x3), we can obtain the expressions for x1 = x1(q1, q2, q3) and x2 = x2(q1, q2, q3), respectively. Then, we can express the integrand f(x1, x2, x3) in terms of q1, q2 and q3 and the integral I3 is expressed as

Equation (44)

where the Jacobian for the change of variables is identified as $\mathcal{J}{=\mathcal{J}}_{3}=1{/\mathcal{G}}_{3}$. This completes the proof of the Jacobian for a three-dimensional coordinate transformation or change of variables.

2.4.  n -dimensional case

In this subsection, we consider the n-dimensional integral In for a function f(x1, ..., xn ) defined in equation (5) by multiplying the unities ${\mathbb{1}}_{i}$ in equation (6) to In in equation (5) sequentially. Then, we integrate In , which is multiplied by the unity, over xi for i = 1 through n successively by making use of the Dirac delta function

Equation (45)

After the integration of all xi variables, the corresponding Jacobian formula is obtained by employing mathematical induction.

First, the integration over x1 can be carried out from equation (7) and the result for the integration over x1 can easily be generalized from the two-dimensional version in equation (30) as

Equation (46)

which leads to

Equation (47)

After the x1 integration, every x1 in the integrand of equation (7) and ${\mathcal{G}}_{1}$ in equation (46) is replaced with

Equation (48)

The constraint equation coming from the convolution with the Dirac delta function in equation (46) is

Equation (49)

Then, we carry out the integration over xi for i = 1 through n − 1 by multiplying ${\mathbb{1}}_{i}$ to In in equation (47) sequentially. We assume that, after the integration over xi for i = 1 through n − 1, the result of the integration of Dirac delta functions is

Equation (50)

which leads to

Equation (51)

For any ji every xj in the integrand of equation (5) and ${\mathcal{G}}_{j}$'s in equation (50) is replaced with

Equation (52)

after the integrations over x1 through xi . There are i constraint equations coming from the convolution with Dirac delta functions in equation (50):

Equation (53)

where j runs from 1 through i.

After multiplying ${\mathbb{1}}_{i+1}$ to In in equation (51), we integrate out one more Cartesian coordinate xi+1 to find that

Equation (54)

where 1 ⩽ ji and 1 ⩽ ki + 1. The proof of the last equality can be found in equation (C8) in appendix C. In combination with equations (50) and (54), we find that

Equation (55)

For any ji + 1 every xj in the integrand of equation (7) and ${\mathcal{G}}_{j}$'s in equation (55) is replaced with

Equation (56)

after the integrations over x1 through xi+1. There are i + 1 constraint equations coming from the convolution with Dirac delta functions in equation (55):

Equation (57)

where j runs from 1 through i + 1. According to mathematical induction, this proves that the assumption in equation (50) with the constraints (52) and (53) is true for all i = 1 through n.

Finally, the integral In can be expressed in terms of q1, ..., qn as

Equation (58)

where 1 ⩽ kn. This completes the proof of the Jacobian $\mathcal{J}{=\mathcal{J}}_{n}=1{/\mathcal{G}}_{n}$ for an n-dimensional coordinate transformation or change of variables.

3. Application

Since the proof of the Jacobian formula in the previous section is rather abstract, readers who are not familiar with the notation might be confused. In this section, we present a few explicit examples of deriving the Jacobian without resort to the general formula derived in the previous section. We expect that the explicit examples will help readers to understand the method presented in the previous section more intuitively and to apply it to a specific change of variables.

3.1. Spherical coordinates

In this subsection, we consider the change of variables from the three-dimensional Cartesian coordinates (x, y, z) to the spherical coordinates (r, θ, ϕ). The Cartesian coordinates can be expressed in terms of the radius r, the polar angle θ and the azimuthal angle ϕ as

Equation (59)

We use cos θ instead of θ as the integration variable and reorganize the order of multiple integrations in order to simplify the computation. That is, (x1, x2, x3) in section 2.3 corresponds to (z, y, x) while (q1, q2, q3) corresponds to (cos θ, ϕ, r), respectively. However, it turns out that the integral is invariant under this reordering. We also note that one can use, for instance, sin θ instead of cos θ as an integration variable and it will not alter the result of the integration. Conventionally, the arctangent function is defined in the region $\left[-\frac{\pi }{2},\frac{\pi }{2}\right]$. The period of the tangent function is π, while the azimuthal angle ranges from 0 to 2π. Thus the angle ϕ for x > 0 is set to be $\mathrm{arctan}\enspace \frac{y}{x}\in \left[-\frac{\pi }{2},\frac{\pi }{2}\right]$ while that for x < 0 is set to be $\pi +\mathrm{arctan}\enspace \frac{y}{x}\in \left[\frac{\pi }{2},\frac{3\pi }{2}\right]$ in order to make the transformation function invertible in the entire range. Then the inverse transformation of equation (59) is expressed as

Equation (60a)

Equation (60b)

Equation (60c)

We consider a three-dimensional integral J3 with an arbitrary integrand f(x, y, z)

Equation (61)

We multiply the unities

Equation (62a)

Equation (62b)

Equation (62c)

to J3 without changing the value of the integral sequentially.

Equation (63)

Here, the Heaviside step function Θ(x) is defined by

First, we integrate out the x1 = z coordinate by multiplying ${\mathbb{1}}_{1}$ in equation (62a) to J3 in equation (61). The integration over z can be performed as

Equation (64)

After the integration over z, every z in the integrand of J3 is replaced with the expression in terms of cos θ, y and x: $z=\sqrt{{x}^{2}+{y}^{2}}/\mathrm{tan}\enspace \theta $, where we omit the simple conversion between trigonometric functions here and after. The integrand of the remaining double integral over x and y is a function of θ, x and y:

Equation (65)

This can also be obtained by substituting cos θ, y and x into equation (39).

Next, we integrate out the x2 = y coordinate by multiplying ${\mathbb{1}}_{2}$ in equation (62b) into J3 in equation (65). The integration over y can be performed as

Equation (66)

where we have used the identity

Equation (67)

Equation (66) can also be obtained by substituting cos θ, ϕ and x into equation (41). After the integration over y, every y in the integrand of J3 is replaced with the expression in terms of cos θ, ϕ and x. The integrand of the remaining integral over x is a function of cos θ, ϕ and x:

Equation (68)

where y = x tan ϕ. After the integrations over both z and y, y = x tan ϕ and z = x/(cos ϕ  tan θ).

Finally, after multiplying ${\mathbb{1}}_{3}$ in equation (62c) into J3 in equation (68), the integration over x3 = x can be performed as

Equation (69)

This can also be obtained by substituting cos θ, ϕ and r into equation (43). Here, the radius r is non-negative because the Dirac delta function requires that $r=\sqrt{{x}^{2}+{y}^{2}+{z}^{2}}$ and x2 + y2 + z2 is non-negative. After the integration over all of the Cartesian coordinates, x, y and z are expressed as in equation (59).

Substituting equation (59) into (64), (66), (69), and f(x, y, z), we find that J3 can be expressed in terms of the spherical coordinates as

Equation (70)

The overall factor r2  sin θ is identified with the Jacobian

Equation (71)

This exactly reproduces the result which can be obtained by applying the formula (44) [1].

3.2.  n -dimensional polar coordinates

In this subsection, we extend the three-dimensional case to the n-dimensional coordinate transformation from the n-dimensional Cartesian coordinates (x1, ..., xn ) to the n-dimensional polar coordinates (r, θ1, θ2, ..., θn−2, ϕ). Here, r is the radius and there are n − 2 polar angles θi 's and a single azimuthal angle ϕ. The corresponding transformation functions are expressed as

Equation (72)

We reorganize the order of integrations of the Cartesian coordinates as (xn , xn−1, ..., x1) for simplicity and the corresponding curvilinear coordinates are reorganized as (cos θ1, cos θ2, ..., cos θn−2, ϕ, r). We consider an n-dimensional integral Jn with an arbitrary integrand f(x1, x2, ..., xn ):

Equation (73)

where dn x = dx1dx2...dxn . We multiply the unities

Equation (74a)

Equation (74b)

Equation (74c)

for i = 1 through n to Jn , sequentially, keeping the integral invariant. Θ(x) in equation (74b) is the Heaviside step function defined in equation (60b). There are numerous ways to perform the multiple integrations over the n Cartesian coordinates. Our strategy to integrate out xi 's is as follows: according to the integrand of the right-hand side of equation (74a), the integration over xni+1 in the integral Jn provides the constraint to the polar angle θi . Thus we choose to integrate over from xn to x3 to express them in terms of the polar angles from θ1 through θn−2. Then we integrate out x2 to express xn through x2 in terms of the n − 2 polar angles and the azimuthal angle ϕ by making use of equation (74b). As the last step, we integrate out x1 to determine all of the Cartesian coordinates in terms of the spherical polar coordinates by making use of (74c).

First, after multiplying ${\mathbb{1}}_{i}$ in equation (74a) into Jn , the integration over xni+1 for i = 1 through n − 2 can be performed as

Equation (75)

where one can obtain the same results from equations (46) and (54) taking care of the order of integration. After the integration over xni+1, we make the replacement

Equation (76)

Then, Jn is expressed as

Equation (77)

where every xi in the last factor for i = 3 through n is replaced by that in equation (76).

After multiplying ${\mathbb{1}}_{n-1}$ in equation (74b) into Jn in equation (77), the integration over x2 can be carried out in a similar manner as is done in equation (66). The result is

Equation (78)

where x2 = x1  tan ϕ after the integration. This can also be obtained from equation (54) while keeping the results in equations (75) and (76). Then, xni+1 for i = 1 through n − 1 can be expressed as

Equation (79)

Then, we find that

Equation (80)

where every xi in the last factor for i = 2 through n is replaced by that in equation (79).

Finally, after multiplying ${\mathbb{1}}_{n}$ in equation (74c) to Jn in equation (80), the integration over x1 can be performed like equation (69) and we obtain

Equation (81)

where we have omitted the replacements of x2 = x1  tan ϕ and x3 through xn that can be obtained from equation (79) on the left-hand side. This result can also be obtained from equation (54) while keeping the results in equations (75), (76), (78) and (79). After integrating out all of the Cartesian coordinates, we reproduce the expression for every xi that is given in equation (72).

Combining all of the results listed above, we find that the n-dimensional coordinate transformation or change of variables is carried out as

Equation (82)

The extra factor rn−1  sinn−2 θ1...sin2 θn−3  sin θn−2 in front of the original integrand f is identified with the Jacobian

Equation (83)

This exactly reproduces the result in references [5, 6], which can be obtained by applying the general formula (58).

4. Conclusions

We have derived the general formula for the Jacobian of the transformation from the n-dimensional Cartesian coordinates to arbitrary curvilinear coordinates by making use of Dirac delta functions, whose arguments correspond to the transformation functions between the two coordinate systems. The multiplication of the trivial identities (6) to the original integral enables us to integrate out the original integration variables corresponding to the Cartesian coordinates systematically. By making use of the chain rule for the partial derivatives, we can carry out the integration over the Cartesian coordinates successively and end up with the integral expressed in terms of the curvilinear coordinates. Then, the Jacobian can be read off by comparing the integrands of the resultant integral with the original one. It turns out that the formula derived in this paper exactly reproduces the Jacobian for the coordinate transformation or change of variables.

We have presented a few examples, where we have integrated out the Cartesian coordinates by making use of Dirac delta functions explicitly without resort to the general formula for the Jacobian derived in this paper. We find that the formulas obtained in these explicit examples are exactly the same as those in the general formula (58). Since the derivation of the Jacobian in the general case that makes use of the chain rule of the partial derivatives is rather abstract, we expect that these examples will give insights on understanding the derivation concretely.

To our best knowledge, this derivation of the Jacobian factor by making use of Dirac delta functions for the coordinate transformation or change of variables from the n-dimensional Cartesian coordinates to the curvilinear coordinates is new. Although there are several ways to derive the Jacobian available in textbooks [13], our derivation could be pedagogically useful in upper-level mathematics or physics courses in practice using Dirac delta functions successively. Compared with the methods popular in the textbook level, our method is more intuitive because we have employed only the explicit calculation of elementary single-dimensional integrals without relying on abstract geometrical interpretations or more abstract Green's theorem with which undergraduate physics-major students are not usually familiar. Furthermore, a detailed derivation of the chain rule for the partial derivatives, which is employed to prove the Jacobian formula, should be a nice working example with which one can understand a rigorous usage of the partial derivatives with multi-dimensional variables without ambiguity.

Acknowledgments

As members of the Korea Pragmatist Organization for Physics Education (KPOP $\mathcal{E}$), the authors thank the remaining members of KPOP $\mathcal{E}$ for useful discussions. This work is supported in part by the National Research Foundation of Korea (NRF) under the BK21 FOUR program at Korea University, Initiative for science frontiers on upcoming challenges, and by grants funded by the Korea government (MSIT), Grant No. NRF-2017R1E1A1A01074699 (J.L.) and No. NRF-2020R1A2C3009918 (J.E. and D.K.). The work of C.Y. is supported by Basic Science Research Program through the National Research Foundation (NRF) of Korea funded by the Ministry of Education (2020R1I1A1A01073770). All authors contributed equally to this work.

Appendix A.: 2-dimensional case

In this section, we consider a two-dimensional coordinate transformation from the Cartesian coordinates (x1, x2) to the curvilinear coordinates (q1, q2). The total differentials of q1 and x1 can be expressed as

Equation (A1a)

Equation (A1b)

where q1 = q1(x1, x2) in equation (A1a) and x1 = x1(q1, x2) in equation (A1b), respectively. Note that the definition of the partial derivative with a subscript are given in section 2.1: ${\left(\partial {x}_{1}/\partial {q}_{1}\right)}_{\left(1\right)}$, ${\left(\partial {x}_{1}/\partial {x}_{2}\right)}_{\left(1\right)}$, and ${\left(\partial {q}_{1}/\partial {x}_{1}\right)}_{\left(2\right)}$ are defined in equations (20)–(22), respectively.

Replacing dx1 in equation (A1a) with the right-hand side of equation (A1b), we obtain

Equation (A2)

Comparing the coefficients of the differentials on both sides, we find that

Equation (A3a)

Equation (A3b)

We note that equation (A3a) is trivial. By making use of equation (A3b), we find that the factor in the denominator of equation (32) ${\left(\partial {q}_{2}/\partial {x}_{2}\right)}_{\left(1\right)}$ can be expressed as

Equation (A4)

which leads to

Equation (A5)

Appendix B.: 3-dimensional case

In this section, we consider a three-dimensional coordinate transformation from the Cartesian coordinates (x1, x2, x3) to the curvilinear coordinates (q1, q2, q3). First, we consider the total differentials of q1 and x1 which can be expressed as

Equation (B1a)

Equation (B1b)

where q1 = q1(x1, x2, x3) in equation (B1a) and x1 = x1(q1, x2, x3) in equation (B1b), respectively. Replacing dx1 in equation (B1a) with the right-hand side of equation (B1b), we obtain

Equation (B2)

Comparing the coefficients of the differentials on both sides, we find that

Equation (B3a)

Equation (B3b)

Equation (B3c)

where equation (B3a) is trivial.

By making use of equation (B3b), we find that the factor in the denominator of equation (41) ${\left(\partial {q}_{2}/\partial {x}_{2}\right)}_{\left(1\right)}$ can be expressed as

Equation (B4)

which leads to

Equation (B5)

In order to prove equation (B11), we take into account the total differentials of q1, x1, q2 and x2. The total derivatives can be expressed as

Equation (B6a)

Equation (B6b)

Equation (B6c)

Equation (B6d)

where q1 = q1(x1, x2, x3) in equation (B6a), q2 = q2(x1, x2, x3) in equation (B6b), x1 = x1(q1, q2, x3) in equation (B6c), and x2 = x2(q1, q2, x3) in equation (B6d), respectively. Substituting equations (B6c) and (B6d) into equations (B6a) and (B6b), we obtain

Equation (B7)

Comparing both sides of equation (B7), we find two relevant non-trivial equations:

Equation (B8)

By making use of Cramer's rule, we find that

Equation (B9)

Then, by making use of equation (B9), we find that the factor ${\left(\partial {q}_{3}/\partial {x}_{3}\right)}_{\left(2\right)}$ in the denominator of equation (43) can be expressed as

Equation (B10)

which leads to

Equation (B11)

Appendix C.: n -dimensional case

The total differentials of q1, ..., qi , x1, ..., xi can be expressed as

Equation (C1)

The partial derivative of qi+1 with respect to xi+1 holding q1, ..., qi , xi+2, ..., xn fixed is

Equation (C2)

Substituting dx1, ..., dxi into dqj , we obtain

Equation (C3)

where 1 ⩽ ji. Because the coefficient of dxi+1 in (C3) should be 0, we obtain the following equation:

Equation (C4)

By making use of Cramer's rule, we find that

Equation (C5)

where

Equation (C6)

and ${\left({\mathbb{J}}_{\left[i{\times}i\right]}^{-1}\right)}^{\left(j\right)}\left(\partial {\mathbb{q}}_{i}\right)$ is identical to ${\mathbb{J}}_{\left[i{\times}i\right]}^{-1}$ except that the jth column is replaced with $\partial {\mathbb{q}}_{i}$. Substituting (C5) into (C2), we obtain

Equation (C7)

Here, the ij minor ${\mathcal{M}}_{ij}\left(\mathbb{A}\right)$ of an n × n square matrix $\mathbb{A}$ is the determinant of a matrix whose ith row and jth column are removed from $\mathbb{A}$. Hence, equation (C7) leads to

Equation (C8)

Please wait… references are loading.