Skip to main content
Journal of Research of the National Institute of Standards and Technology logoLink to Journal of Research of the National Institute of Standards and Technology
. 2007 Jun 1;112(3):163–173. doi: 10.6028/jres.112.013

Acoustic Eigenvalues of a Quasispherical Resonator: Second Order Shape Perturbation Theory for Arbitrary Modes

James B Mehl 1,2
PMCID: PMC4656004  PMID: 27110463

Abstract

The boundary-shape formalism of Morse and Ingard is applied to the acoustic modes of a deformed spherical resonator (quasisphere) with rigid boundaries. For boundary shapes described by r = a [1 − ε ℱ(θ, ϕ)], where ε is a small scale parameter and ℱ is a function of order unity, the frequency perturbation is calculated to order ε2. The formal results apply to acoustic modes whose angular dependence is designated by the indices and m. Specific examples are worked out for the radial ( = 0) and triplet ( = 1) modes, for prolate and oblate spheroids, and for triaxial ellipsoids. The exact eigenvalues for the spheroids, and eigenvalue determined with finite-element calculations, are shown to agree with perturbation theory through terms of order ε2. This work is an extension of the author’s previous papers on the acoustic eigenfrequencies of deformed spherical resonators, which were limited to the second-order perturbation for radial modes [J. Acoust. Soc. Am. 71, 1109-1113 (1982)] and the first order-perturbation for arbitrary modes [J. Acoust. Soc. Am. 79, 278–285 (1986)].

Keywords: acoustic cavity resonators, acoustic eigenvalues, quasispheres, shape perturbation theory

1. Introduction

Spherical acoustic resonators [1] have been successfully applied to measurement of the universal gas constant [2] and to gas thermometry [37]. The radial (0, n) acoustic modes are well-suited for high-accuracy work because they are non-degenerate, well separated, insensitive to visco-thermal boundary effects, and only weakly dependent on the details of small shape imperfections [1,8,9]. Nearly-spherical resonators, now referred to as quasispherical resonators [10,11], have been designed to facilitate the measurement of the electromagnetic resonances for determination of the mean radius of the quasisphere. The speed of sound can be determined from the combination of measured acoustic and electromagnetic resonance frequencies [12]. The acoustic and electromagnetic eigenvalues for quasispherical resonators must be evaluated using approximation methods. In this paper, boundary-shape perturbation theory [13] is used to calculate the acoustic eigenvalues.

An idealized rigid spherical resonator of radius a has acoustic modes with the acoustic pressure proportional to the eigenfunctions

Φnm=j(knr)ϒm(θ,ϕ), (1)

where j (ξ) is a spherical Bessel function and ϒℓm is a linear combination of spherical harmonics

ϒm=μ=αmμYμ (2)

with coefficients α chosen to make the ϒℓm real. The eigenvalues for the perfect spherical geometry are kℓn = ξℓn/a, where ξℓn is the nth root of j (ξℓn) = 0.

The bounding surface of a quasispherical resonator has the form

rs=a[1ε(θ,ϕ)], (3)

where ε is a scale parameter satisfying 0 < ε << 1 and F is a smooth, non-negative function of θ and ϕ. The = 0 eigenvalues of a quasisphere differ from those of a perfect sphere of the same volume by a fraction of order ε2 or higher [1, 8]. The non-radial acoustic modes of a perfect sphere occur in rnultiplets of degeneracy 2 + 1. Typically, this degeneracy is split to order ε in a quasisphere, but the mean eigenvalue of any multiplet differs from the corresponding eigenvalue of a perfect sphere of the same volume by an amount of order ε2 or higher [9]. The same is true of the electric and magnetic modes of a quasisphere with perfectly conducting walls [12].

In principle, a highly accurate measurement of the speed of sound in a gas can be made by measuring the acoustic and electromagnetic resonance frequencies of the same quasisphere. Geometric contributions to the error will then be of order ε2. If the shape is known, higher accuracy can be obtained if the theoretical coefficients of the ε2 perturbation terms can be calculated. This has already been achieved for the radial acoustic modes [8]. A corresponding theory for the electromagnetic modes, a subject of current research by the author, is much more complex, and is closely related to the theory of the second order shape perturbation theory for the non-radial acoustic modes, as developed in this paper. The results derived here will be useful in experimental studies of quasispheres which will compare the effects of shape on the acoustic and electromagnetic spectra. Also, the results enable the use of the lowest-frequency acoustic modes, the 11m triplet, to be used for high-accuracy work.

2. Formalism

Morse and Feshbach [13] (MF hereafter) present a formalism for calculating the eigenfrequencies of an acoustic cavity resonator C enclosed within an unperturbed cavity C0. Figure 1 illustrates the geometry; the surfaces enclosing C and C0 are designated S and S0, respectively, and the region between C and C0 is designated C′. The unperturbed cavity has a set of eigenfunctions ΦN and eigenvalues kN2, satisfying the Helmholtz equation

(2+kN2)ΦN=0 (4)

in C0 and the Neumann boundary condition

ΦNnn^ΦN=0 (5)

on S0. (For brevity, the subscript N in these equations, and other subscripts in capital letters, are used to represent sets of lower-case mode indices.)

Fig. 1.

Fig. 1

Perturbed cavity C (boundary S) within unperturbed cavity C0 (boundary S0). The region between C and C0 is designated C′.

The perturbed problem is defined by a surface S enclosed within S0, and enclosing a cavity C. The perturbed problem satisfies

(2+k2)Φ=0 (6)

in C and the Neumann boundary condition

Φn=0 (7)

on S. A second-order expression for the perturbed eigenvalue k2 is [MF Eq. (9.2.53)]

(k2kN2)NNN+PN[ANP(k2kN2)NNP][APN(k2kP2)NPN]NPP(k2kP2)APP, (8)

where

AQP=SΦPΦQndS (9)

and

NPQ=NQP=CΦPΦQdV. (10)

A more useful computational form for the integrals (9) can be obtained by applying the divergence theorem to ΦP∇ΦQ ± ΦQ∇ΦP in C to obtain

AQP+APQ=C[2ΦPΦQ(kP2+kQ2)ΦPΦQ]dV (11)

and

AQPAPQ=(kQ2kP2)CΦPΦQdV. (12)

A negative sign occurs in these expressions because the outward normal from C′ on S is n^. For Eq. (12), use was made of the the orthogonality of the unperturbed functions in C0 for pq:

CΦPΦQdV+CΦPΦQdV=0,QP. (13)

An expression for the diagonal terms follows directly from Eq. (11):

APP=C[kp2|ΦP|2|ΦP|2]dV. (14)

This is an integral over the region between S and S0 of a quantity proportional to the difference between the potential energy and the kinetic energy. The corresponding term for the perturbation of the electromagnetic modes has the same form [12]. An expression for the off-diagonal terms can be obtained from the sum of Eqs. (11) and (12):

APQ=C[kP2ΦPΦQΦPΦQ]dV. (15)

3. Deformed Spherical Resonator

Consider a deformed spherical resonator with a boundary surface S defined by Eq. (3).

When applied to a quasisphere with S defined by Eq. (3), the volume C′, and accordingly the integrals in Eqs. (14) and (15) are of order ε. Equation (8) can then be solved iteratively to obtain

k2kN2kN2=ANNkN2NNN+PN|ANP|2NNNNPPkN2(kN2+kP2)+O(ε3). (16)

To evaluate Eq. (16) to order ε2, the numerator of the first term on the right must be calculated to O(ε2), the denominator to O(ε), the coefficients ANP in the sum term to O(ε). The normalization constants in the denominator of the sum term need only be calculated to O(1).

The acoustic modes of a perfect spherical resonator occur in multiplets with (2 + 1)-fold degeneracy. Only the = 0 radial modes are nondegenerate. When calculating the perturbation series for nonradial modes, the coefficients α in Eq. (2) should be chosen to make the coefficients APQ zero for the modes with kP = kQ. Equation (12) shows that ANP = APN exactly for degenerate pairs, so the proper choice of coefficients α can be obtained by diagonalizing the submatrix [ANP] linking the multiplet terms.

More precisely, the off-diagonal terms of this sub-matrix must be of order ε2. Consider the application of Eq. (8) to the multiplet components. The first-order perturbation shift of each component is of order ε. In an iterative solution of Eq. (8) the order of the terms in the denominator of the sum terms would be NPP = O(1), k2kN2=O(ε), and APP = O(ε). The numerator is the square of ANP(k2kN2)NPN. Both k2kN2 and NNP are of order ε, so if ANP = O(ε2) the numerator will be of order ε4, and the entire sum term of order ε3.

To get the first term in Eq. (16) to O(ε2), the numerator must be calculated to O(ε2), and the denominator to O(ε). The thickness of the integration volume C′ in Eq. (14) is of order ε, so the integrand of the numerator is needed to O(ε). Within C′ the radial derivative of the spherical Bessel function is of order ε so the function itself satisfies

j(knr)=j(ξn)+O(ε2). (17)

Integrals NPQ with PQ do not appear in Eq. (16), only normalization integrals for which the repeated indices are superfluous. The notation can hence be simplified by using an ordinary math font for N and a set of lower-case indices to designate the mode. In the new notation, the normalization integral in the denominator is

Nnm=dΩ0rs[j(knr)]2|ϒm|2r2dr, (18)

which can be evaluated as the difference between an integral from r = 0 to a and an integral from rS to a to obtain, for ℓnm ≠ 010,

Nnm=a32j[(ξn)]2[1(+1)ξn2]εa3[j(ξn)]2|ϒm|2dΩ+O(ε2). (19)

For the special case nℓm = 010, the eigenfunction is j0(k01r) = 1 and the eigenvalue is ξ01 = 0. The normalization constant is

N010=a33+O(ε), (20)

which differs from Eq. (19) by a factor of 2/3.

The function ℱ may itself depend on the scaling parameter ε; it is useful to make this explicit:

=0+ε1+O(ε2). (21)

The first term in Eq. (16) then has the form

Anmnmkn2Nnm=2ε[ξn2ϒm2(0+ε1ε02)|rϒm|2(0+ε1)]dΩξn2(+1)2εξn20|ϒm|2dΩ+O(ε3). (22)

(Note that, the operator r∇ appearing in this expression involves only angular derivatives.) The coefficients (15) in the perturbation series are only needed to O(ε), so the integrand is only needed to O(1); only the leading order of Eq. (18) is needed in the denominator. The sum in Eq. (16) simplifies to

nmnm|Anmnm|2NnmNnmkn2(kn2kn2)=nmnm4ε2|Bmm(n)|2ξn2(+1)ξn2[ξn2(+1)](ξn2ξn2) (23)

where

Bmm(n)=[ξn2ϒmϒmr2ϒmϒm]0dΩ. (24)

The sum over n′ in Eq. (23) is

Sn=v=11ξn2ξv2ξv2ξn2(+1), (25)

where the prime on the summation symbol indicates the omission of the terms with ν = ℓn. The sums are evaluated analytically in the Appendix using the technique of Ref [8]. The results for ′ ≠ 0 are

Sn={j(ξn)2ξnj(ξn),for,ξn23(+1)4[ξn2(+1)]2,for=, (26)

and, for ′ = 0,

Sn0={12ξn2+j0(ξn)2ξnj0ξn,for0,1/(4ξ0n2),for=0. (27)

The full sum in (16) is thus

nmnm|Anmnm|2NnmNnmkn2(kn2kn2)=4ε2ξn2(+1)m|Bmm(n)|2Sn+2ε2|Bm00(n)|2ξn2[ξn2(+1)], (28)

where the last term is 1/3 of the contribution from the 010-mode, which has a special normalization; 2/3 of the contribution of this term is included in the sum term.

3.1 Reference Eigenvalues

In order to separate out the effects of shape from the effects of volume, the perturbed eigenvalues k2 will be compared with the eigenvalues (kℓnm)2 of a reference sphere of the same volume V as the perturbed sphere. The fractional difference equals

k2(knm)2(knm)2=(ka)2ξn2ξn2, (29)

where (ka′)2/ξ2ℓn, is the product of (a′/a)2 and the sum of 1 and the series on the right side of Eq. (16). The ratio of the volume V = 4π(a′)3/3 to the volume V0 = 4πa3/3 of the unperturbed sphere is

(aa)3=14π(1ε)3dΩ=13ε0+3ε2021+O(ε3). (30)

where the triangular brackets indicate an average over solid angle. The ratio of the squared radii is

(aa)=12ε0+ε2[02+2021]+O(ε3). (31)

The desired fractional difference is

(ka)2ξn2ξn2=Anmnmkn2Nnm+4ε2ξn2(+1)m|Bmm(n)|2Sn+2ε2|Bm00(n)|2ξn2[ξn2(+1)]2ε0+ε2[02+20221]2ε0Anmnmkn2Nnm+O(ε3), (32)

where the term coupling to the 01-mode has been made explicit.

3.2 Series Evaluation

Identification of the contributions to the coefficients Bmm(n) is facilitated by expressing the shape as

0=λμcλμYλμ. (33)

Equation (24) then involves linear combinations of terms of the forms

m|λμ|mYmYλμYmdΩ. (34)

and

=(+1)λ(λ+1)(+1)2m|λμ|m,Ym(YλμθYmθ+1sin2θYλμϕYmϕ)dΩ (35)

where Eq. (35) was obtained using the technique described in the Appendix of Ref [9]. Alternatively, Eq. (35) can be derived using the raising and lowering angular momentum operators (see, e.g. Ref [14]). The bracket expressions (34) vanish unless the following conditions are satisfied:

1.||λ+, (36)
2.m=μ+m, (37)
3.+λ+must be even. (38)

It is clear from Eqs. (2), (34), and (35) that Bmm(n) can be expressed as a linear combination of bracket ex pressions (34) with | m′ | ≤ ′ and | m |≤ ℓ. Accordingly, it is possible to identify the terms that can possibly contribute to non-varnishing values of Bmm(n), by applying the following rules:

  1. First look at the non-vanishing cλµ in Eq. (33).

  2. For the unperturbed mode index , look at each expansion-coefficient index λ and find the values of ′ satisfying the conditions of Eqs. (36) and (38).

  3. Note that the coefficients α in Eq. (2) are often non-zero only for µ′ = ±m. Consider the possible terms; then for each unperturbed mode index m, and expansion coefficient index µ find the value of m′ satisfying Eq. (37).

  4. Once the possible non-vanishing coefficients Bmm(n) are identified, computation of the values of the coefficients can be carried out using symbolic algebra software.

4. Examples

The second order perturbations of the = 0 radial modes and the three-fold degenerate = 1 modes are worked out in this section for prolate and oblate ellipsoids. and for triaxial ellipsoids.

The eigenfunctions of the unperturbed = 0 modes are

Φ0n0=j0(k0nr)ϒ00. (39)

The appropriate = 1 unperturbed eigenfunctions for any quasi-spherical resonator that has its major axes aligned with the x^, y^, and z^ directions are the product of j1(k1nr) and

ϒ11=Y11+Y112=34πsinθcosϕ=34πxrϒ10=Y10=34πcosθ=34πzrϒ1,1=Y11+Y112i=34πsinθcosϕ=34πyr, (40)

for which the submatrix with components A1nm,1nm is diagonal in mm′.

4.1 Prolate Spheroid

For a spheroid of semi-major axis a and semi-minor axes b = a/(1 + ε), with ε > 0, the radial coordinate is

r=a1+(2ε+ε2)sin2θ=a(1ε), (41)

with

=sin2θ+ε(12sin2θ32sin4θ)+o(ε2), (42)

for which

0=23,02=815,1=715. (43)

The shape function ℱ0 is an exact linear combination of Y00 and Y20. Accordingly. for the radial modes, the contributions to B00m(n) are limited to ′= 0 and ′= 2. For the = 1 modes, the contributions to B1mm(n) are limited to ′=1 and ′ = 3.

4.1.1 = 0 Modes

The non-vanishing coefficients are

B0000(n)=23ξ0n2,B0020(n)=2515ξ0n2. (44)

Evaluation of Eqs. (79) and (78) yields

S0n0=14ξ0n2,S0n2=16, (45)

where the latter was obtained using the condition j0(ξ0n)=j1(ξ0n)=0 and recurrence relations for the spherical Bessel functions. Evaluation of the perturbation series (32) yields

(ka)2ξ0n2ξ0n2=8ξ0n2ε2135+O(ε3), (46)

in agreement with Eq. (30) of Ref [8].

4.1.2 = 1 Modes

Equation (22) is

A1n01n0(k1n)2N1n0=4(ξ1n24)5(ξ1n22)ε2(9ξ1n4126ξ1n2+440)175(ξ1n22)2+O(ε3)A1n±1,1n±1(k1n)2N1n±1=4(ξ1n23)5(ξ1n22)ε2(6ξ1n4259ξ1n2+270)175(ξ1n22)2ε2+O(ε3). (47)

The non-vanishing coefficients are

B10,10(n)=25(ξ1n24),B10,30(n)=22135(ξ1n24),B1,±1,1,±1(n)=25(2ξ1n23),B1,±1,3,±1(n)=21435(ξ1n24). (48)

From Eqs. (74) and (72), the required sums are

S1n1=ξ1n264(ξ1n22)2,S1n3=ξ1n2510(ξ1n24), (49)

the latter following from recurrence relations for the spherical Bessel functions and the condition j1(ξ1n)=0. Substitution of Eqs. (43) and (47)–(49) into Eq. (32) yields

(ka)2ξ1n2ξ1n2=8(ξ1n2+1)15(ξ1n22)ε+4(54ξ1n8+373ξ1n6+495ξ1n46924ξ1n2+2980)7875(ξ1n22)ε2+O(ε3),m=0, (50)

and

(ka)2ξ1n2ξ1n2=4(ξ1n2+1)15(ξ1n22)ε+2(72ξ1n8961ξ1n6+3285ξ1n43282ξ1n22560)7875(ξ1n22)ε2+O(ε3),m=±1. (51)

The scalar Helmholtz equation separates in spheroidal coordinates, so the acoustical eigenvalues can be determined by direct numerical calculations [15]. The eigenvalues calculated numerically for a series of values of ε are compared with Eqs. (50) and (51) in Figs. 2 and 3.

Fig. 2.

Fig. 2

Comparison of perturbation series (lines) for prolate spheroid with exact numerical solutions (symbols).

Fig. 3.

Fig. 3

The absolute difference between the exact numerical eigenvalues knum2 and the predictions kpert2 of Eqs. (50) and (51) plotted as a fraction of k2ln ≡ (ξ1n/a′)2, as functions of ε. The dashed line, intended as a guide to the eye, is proportional to ε 3. The plots show that the differences are approximately proportional to ε3. The numerical values exceed the perturbation values for the 111 mode over the full displayed range, and for the 140 mode for ε < 0.025; for all other cases the difference is negative.

The average eigenvalue perturbation for the 1n-triplet is, from Eqs. (50) and (51),

(ka)2ξ1n2ξ1n21n=8(3ξ1n814ξ1n6+90ξ1n4243ξ1n2+10)1125(ξ1n22)3ε2+O(ε3), (52)

which has no linear term, consistent with the general results derived in Ref [9].

4.2 Oblate Spheroid

For an oblate spheroid of semi-major axis a and semi-minor axes b = a/(1 + ε), ε > 0, the radial coordinate is

r=a1+(2ε+ε2)cos2θ=a(1ε), (53)

with

=cos2θ+ε(12cos2θ32cos4θ)+O(ε2), (54)

for which

0=13,02=15,1=25. (55)

The perturbation calculations for the = 0 and = 1 modes parallel those for the prolate spheroid and will not be reproduced in detail here. The final expression for the eigenvalue perturbations for the radial modes is exactly the same as the result for the prolate spheroid [Eq. (46)]. For the = 1 modes the fractional perturbations are

(ka)2ξ1n2ξ1n2=8(ξ1n2+1)15(ξ1n22)ε+4(54ξ1n8677ξ1n6+3645ξ1n46924ξ1n21220)7875(ξ1n22)3ε2+O(ε3),m=0, (56)
(ka)2ξ1n2ξ1n2=4(ξ1n2+1)15(ξ1n22)ε+2(72ξ1n8+89ξ1n6+135ξ1n43282ξ1n2+1640)7875(ξ1n22)ε2+O(ε3),m=±1. (57)

Exact solutions for the oblate spheroid [16] were calculated and compared with Eqs. (56) and (57). The agreement, like the corresponding agreement for the prolate spheroid, is very good. The plots resemble Figs. 2 and 3.

The mode average is exactly the same as Eq. (52) for the prolate spheroid.

4.3 Triaxial Ellipsoid

The surface of the triaxial ellipsoid defined by

x2(1+ε2)2+y2+z2(1+ε1)2=a2(1+ε1)2(1+ε2)2 (58)

can be expressed in the form of Eq. (3) with

ε0=ε1sin2θ+ε2(cos2θ+sin2θsin2ϕ) (59)

and

ε21=ε12(32cos2θ1)sin2θ+ε1ε2(sin2ϕ3cos2θcos2ϕ)sin2θ+ε22(1+52sin2θcos2ϕ32sin4θcos4ϕ) (60)

The shape ℱ0 is an exact linear combination of Y00, Y20, and Y2,±2. The non-vanishing values of Bmm(n) are accordingly limited to the same values of ℓ′ as for the spheroids.

4.3.1 = 0 Modes

The non-vanishing coefficients are

εB0000(n)=23ξ0n2(ε1+ε2),εB0000(n)=515ξ0n2(2ε1ε2). (61)

Evaluation of the perturbation series (32) yields

(ka)2ξ0n2ξ0n2=8ξ0n2135(ε12ε1ε2+ε22)+O(ε3). (62)

The correctness of this result was checked by calculating the radial-mode eigenvalues of a triaxial ellipsoid using a finite-element method. The parameters ε1 and ε2 were varied, with the ratio held constant at ε1/ε2 = 2. Figure 4 shows that the difference between the finite-element results and Eq. (62) is cubic in ε 3.

Fig. 4.

Fig. 4

The differences between the values of kfem2 determined with the finite-element method for the 02 radial mode and the predictions kpert2 of Eq. (62) for an ellipsoid with ε2 = ε1/2. The line proportional to ε13 was fit to the plotted points.

4.3.2 = 1 Modes

The non-vanishing coefficients Bmm(n) are, with ξ =ξ1n for brevity,

B1,1,1,1(n)=25(2ξ23)(ε1+ε2),B1,0,1,0(n)=25[(ξ24)ε1+(2ε23)ε2],B1,1,1,1(n)=25[(2ξ23)ε1+(ξ24)ε2],B1,1,3,3(n)=370(ξ24)ε2,B1,1,3,1(n)=1470(ξ24)(4ε1ε2),B1,0,3,0(n)=2135(ξ24)(2ε1ε2),B1,1,3,1(n)=1470(ξ24)(4ε13ε2),B1,0,3,2(n)=135(ξ24)ε2,B1,1,3,3(n)=370(ξ24)ε2. (63)

The fractional perturbations for the 1nm modes have the form

(ka)2ξ1n2ξ1n2=4(ξ2+1)(p1mε1+p2mε2)15(ξ22)+p11mε12+p12mε1ε2+p22mε227875(ξ22)3, (64)

where

p1,1=p1,1=p2,0=p2,1=1,p1,0=p2,1=2,p11,1=2(72ξ8961ξ6+3285ξ43282ξ22560),p12,1=8(27ξ876ξ6+1035ξ43462ξ2440),p22,1=4(54ξ8+373ξ6+495ξ46924ξ22980),p11,0=p22,1,p12,0=p12,1,p22,0=p11,1,p11,1=p11,1,p12,1=8(9ξ8142ξ6180ξ4+1821ξ2670),p22,1=p11,1. (65)

These equations have the appropriate limits when either εl or ε2 is zero. The mode average is

(ka)2ξ1n2ξ1n21n=8(3ξ1n814ξ1n6+90ξ1n4243ξ1n2+10)1125(ξ1n22)3(ε12ε1ε2+ε22)+O(ε3), (66)

The correctness of Eqs. (64) and (65) was tested by calculating the modes of a triaxial ellipsoid using the finite-element method. The parameters εl and ε2 were varied, with the ratio held constant at εl/ε2 = 2. This choice corresponds to a uniform splitting of the triplet (the case considered in Ref [10]) as shown in Fig. 5.

Fig. 5.

Fig. 5

Fractional eigenvalue perturbations for the 11 m modes: the points are numerical results determined with the finite-element method, the lines represent Eqs. (64)–(65) for an ellipsoid with ε2 = ε1/2. The key identifies the lines by the value of the index m.

Figures 5 and 6 show the close agreement of the eigenvalues determined with the finite-element method and Eqs. (64) (65).

Fig. 6.

Fig. 6

The differences between the values of kfem2 determined with the finite-element method for the 11 m modes and the predictions kpert2 of Eqs. (64)–(65) for an ellipsoid with ε2 = ε1/2. The lines proportional to ε13 were fit to the plotted points. The key identifies the lines by the value of the index m.

5. Concluding Remarks

The formalism derived in this article can be applied, in principle, to arbitrary quasi-spherical resonators whose shape can be represented by Eq. (3). Section 3.2 lists the general principles that determine the possible contributions to the general series (32). Once the possible terms are identified, the use of symbolic algebra software can be used to calculate the terms. For increasingly complex shapes, this process should be programmed so as to minimize human error.

Acknowledgments

The author extends his thanks to Michael Moldover for encouraging this work and for useful conversations. This work was supported by the National Institute of Standards and Technology through a contract with K. T. Consulting, Inc.

Biography

About the author: James B. Mehl has maintained a strong collaboration with NBS/NIST since 1979. He was a Guest Researcher at the NIST Chemical Science and Technology Laboratory. He is now retired from the University of Delaware, where he served as Professor of Physics Department Chair, and Associate Dean. The National Institute of Standards and Technology is an agency of the Technology Admininstration, U.S. Department of Commerce.

6. Appendix. Evaluation of Infinite Sums

The sum Sn defined by Eq. (25) can be evaluated explicitly. For the case ′ ≠ 0, consider the contour integral

I=C1ξ2ξn2ξ2ξ2(+1)j(ξ)j(ξ)dξ, (67)

where C is a rectangular contour with corners at (± xN, ±y0), with y0 > 0 and xN sufficiently large that N zeros of j(ξ) lie along the positive real axis within C. The integrand is bounded on C, so the integral approaches zero as N → ∞; the sum of all residues within the contour is also zero in this limit.

The integrand has poles at ±ξℓn, (+1), and ±ξℓ′ν for ν =1,2,…N. When ℓ′ the poles are all first order, with residues

(±ξn)=1ξn2(+1)j(ξn)2ξnj(ξn), (68)

and

(±(+1))=1ξn2(+1), (69)
(±ξν)=1ξn2ξν2ξν2ξν2(+1). (70)

The sum of all residues ℛ(±ξℓ′µ) is 2Sn, so the condition that the sum of all residues is zero implies

Sn+(ξn)+((+1))=0, (71)

or

Sn=j(ξn)2ξnj(ξn),. (72)

When ℓ' = the pole at ξℓn is second order, with residue

(±ξn)=3ξn2(+1)4[ξn2(+1)]2. (73)

and the sum of the series is

Sn=ξn23(+1)4[ξln2(+1)]2. (74)

Next consider the case ′ = 0. The second factor in the integrand of the contour integral is then unity; the residues are

(±ξn)=1ξn2j0(ξn)2ξn2j0(ξn),0, (75)

and

(±ξν)=1ξn2ξν2. (76)

There is a single pole at ξ01 = 0, so the sum of all residues is

ν=(±ξν)+2(±ξn)=[2Sn0+1ξn2][2ξn2+j0(ξn)ξnj0(ξn)]=0, (77)

which implies

Sn0=12ξn2+j0(ξn)2ξnj0(ξn). (78)

When ℓ = 0 the (distinct) poles at ±ξ0n are second order; the residues are 3/(4ξ0n2), so the sum is

S0n0=14ξ0n2. (79)

7. References

  • 1.Moldover Michael R, Mehl James B, Greenspan Martin. Gas-filled spherical resonators: Theory and experiment. J Acoust Soc Am. 1986;79:253–272. [Google Scholar]
  • 2.Moldover Michael R, Trusler JP Martin, Edwards Terry, Mehl James B, Davis Richard S. Measurement of the universal gas constant using an acoustic resonator. J Res Natl Inst Technol. 1986;93:85–144. [Google Scholar]
  • 3.Moldover Michael R, Boyes Steven J, Meyer Christopher W, Goodwin Anthony RH. Thermodynamic temperatures of the triple points of mercury and gallium and in the interval 217 K to 303 K. J Res Natl Inst Stand Technol. 1999;104:11–46. [Google Scholar]
  • 4.Ewing Michael B, Trusler JP Martin. Primary acoustic thermometry between T = 90 K and T = 300 K. J Chem Thermodynam. 2000;32:1229–1255. [Google Scholar]
  • 5.Strouse Gregory F, Defibaugh Dana R, Moldover Michael R, Ripple Dean C. Progress in primary acoustic thermometry at NIST, 273 K to 505 K. In: Ripple Dean C., editor. Temperature: Its Measurement and Control in Science and Industry, Vol. 7, AIP Conference Proceedings. Melville, NY: 2003. pp. 31–36. [Google Scholar]
  • 6.Benedetto G, Gavioso RM, Spagnolo R, Marcarino P, Merlone A. Acoustic measurements of the thermodynamic temperature between the triple point of mercury and 380 K. Metrologia. 2004;41:7498. [Google Scholar]
  • 7.Pitre Laurent, Moldover Michael R, Tew Weston L. Acoustic thermometry: new results from 273 K to 77 K and progress towards 4 K. Metrologia. 2006;43:142–162. [Google Scholar]
  • 8.Mehl James B. Resonance frequencies of deformed spherical resonators. J Acoust Soc Am. 1982;71:1109–1113. [Google Scholar]
  • 9.Mehl James B. Resonance frequencies of deformed spherical resonators II. J Acoust Soc Am. 1986;79:278–285. [Google Scholar]
  • 10.Mehl James B, Moldover Michael R, Pitre Laurent. Designing quasi-spherical resonators for acoustic thermometry. Metrologia. 2004;41:295–304. [Google Scholar]
  • 11.May Eric F, Pitre Laurent, Mehl James B, Moldover Michael R, Schmid James W. Quasi-spherical cavity resonators for metrology based on the relative dielectric permittivity of gases. Rev Sci Inst. 2004;75:3307–3317. [Google Scholar]
  • 12.Mehl James B, Moldover Michael R. Measurement of the ratio of the speed of sound to the speed of light. Phys Rev A. 1986;34:3341–3344. doi: 10.1103/physreva.34.3341. [DOI] [PubMed] [Google Scholar]
  • 13.Morse PM, Feshbach H. Methods of Theoretical Physics, Part II. McGraw-Hill; New York: 1953. pp. 1052–1064. [Google Scholar]
  • 14.Jackson John D. Classical Electrodynamics, Third Edition, chapt 9. Wiley; New York: 1999. [Google Scholar]
  • 15.Chang Christopher TM. Natural resonant frequency of a prolate acoustical resonator. J Acoust Soc Am. 1971;49:611–614. [Google Scholar]
  • 16.Chang Christopher TM. Natural resonant frequency of an oblate acoustical resonator. J Acoust Soc Am. 1972;51:1–5. [Google Scholar]

Articles from Journal of Research of the National Institute of Standards and Technology are provided here courtesy of National Institute of Standards and Technology

RESOURCES