This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Yang–Mills Monopoles in
Extremal Reissner–Nordström Black Hole Metric

Xiangqin Zhang
School of Mathematics and Statistics
Henan University
Kaifeng, Henan 475004, PR China
and
Henan University of Animal Husbandry and Economy
Zhengzhou, Henan 450044, PR China

Yisong Yang111Email address for correspondence: [email protected]
Courant Institute of Mathematical Sciences
New York University
New York, NY 10012, USA
Abstract

We show that the closed-form monotone solution linking two different vacuum states, for the exterior Yang–Mills wave equation over the extremal Reissner–Nordström spacetime found in a recent work of Bizoń and Kahl, is the unique solution to an associated domain-wall type energy minimization problem. We also derive the accompanying interior Yang–Mills wave equation within the same formalism. We obtain a few more closed-form solutions, regular and singular, among other solutions of oscillatory behavior, and discuss several interesting features of the solutions based on some energy consideration.

Keywords: Reissner–Nordström black hole metric, Yang–Mills fields, monopoles, domain-wall solitons, minimization problem, existence and uniqueness of solution.

PACS numbers: 02.30.Hq, 11.15.q, 11.27.+d, 12.39.Ba

MSC numbers: 34B40, 35J50, 81T13, 83C57

1 Introduction

The concept of magnetic monopoles was first conceived by P. Curie [9] and later theoretically formulated by Dirac [10] based on electromagnetic duality of the Maxwell theory. A remarkable feature of this concept is that its existence could lead to an explanation why electric charges are multiples of a unit charge [10, 13, 25, 27, 39]. Such a charge quantization phenomenon was then extended by Schwinger [28] to dyons, namely, particles of both electric and magnetic charges. In modern development, thought experiments with monopoles immersed in a type-II superconductor led Mandelstam [19, 20], Nambu [23], and t’ Hooft [33, 33] to come up with a linear confinement mechanism for monopoles and their colored, non-Abelian or the Yang–Mills, field theory versions, which has propelled enormous activities and progress towards an understanding [29, 30] of the unsettled puzzle of quark confinement [14]. Like that in the classical Maxwell theory, a monopole in the pure Yang–Mills theory, for example, the celebrated Wu–Yang monopole [1, 13], is of infinite energy [16]. There are several well-recognized paths which may be taken to overcome this infinity or divergence problem. These include adopting the model of the Born–Infeld nonlinear electrodynamics [4, 5], coupling the Yang–Mills fields with matter fields [1, 13, 17, 24, 25, 32, 35, 38, 40], namely, considering the Yang–Mills–Higgs theory, and hosting the Yang–Mills fields over curved spacetime backgrounds due to gravity. In the context of this last approach, if a gravitational black hole is present and fields are considered outside the event horizon so that possible singularities leading to energy divergence are concealed inside the horizon, as in the scenario of the cosmic censorship hypothesis [22, 37], the total energy of the fields outside the horizon could remain finite, as a result. The present study belongs to this category of investigation. Specifically our work is motivated by the recent paper of Bizoń and Kahl [3] who considered spherically symmetric SU(2)SU(2) Yang–Mills fields in the exterior of the event horizon of the extremal Reissner–Nordström black hole and obtained a countable family of static regular solutions to the Yang–Mills equations, given in terms of a scalar field sequence {Wn}\{W_{n}\}, say, among which the first two, W1W_{1} and W2W_{2}, are in closed forms. These solutions resemble the Bartnick–McKinnon solutions [2, 31] featured as soliton-like configurations connecting vacuum states at infinity and oscillating about the unit-charge monopole state in local domains. The main result of our present work is to show that, viewed as a connecting orbit between different vacuum states which passes through the unit-charge monopole state midway, W1W_{1} is the unique global energy minimizer of the problem. The solution W2W_{2}, however, enjoys a different signature – it exhibits itself as a local energy maximizer in a well-formulated sense. Besides, we shall also obtain a new closed-form solution which has two singular points (in fact, two singular shells), near which energy blow-up takes place. Furthermore, we adapt the formalism of [3] to consider interior monopoles (inside the event horizon) and derive a governing wave equation which may be viewed as a companion of the wave equation of Bizoń–Kahl [3] for exterior monopoles. For this interior monopole equation, we obtain three closed-form solutions, similar to W1W_{1}, W2W_{2}, and the third singular solution to the exterior monopole equation described above, among which two are regular everywhere away from the curvature singularity of the Reissner–Nordström spacetime and the third one is singular at one point (in fact, one shell). It is noted that this interior monopole equation possesses a family of oscillatory solutions as the exterior equation, as well.

The content of the rest of the paper is outlined as follows. In Section 2, we formulate the problems to be studied with a review of the Yang–Mills fields over an extremal Reissner–Nordström spacetime and the exterior monopole equation derived in [3], followed with an extension of such a formulation to get the interior, ‘accompanying’, monopole equation. Subsequently, we focus on static equations. In Section 3, we review the two closed-form regular solutions, W1W_{1} and W2W_{2}, obtained in [3], and present a third, singular, one, for the exterior equation. We obtain the exact values of the energies of these two regular solutions to be further explored in later sections. We then present two closed-form regular solutions and a third, singular, one, for the interior equation, similar to those for the exterior equation. In Section 4, we study several issues of the exterior solutions based on some energy consideration. In particular, we formulate a domain-wall type minimization problem that associates a distinguishing significance to W1W_{1} as the unique global minimizer of the problem. We also present a virial identity which displays an energy partition relation for finite-energy solutions. In Section 5, we first prove the existence of a solution to the domain-wall minimization problem formulated in Section 4. Note that, due to the behavior of the weight function in the energy functional, it is not a trivial matter to preserve the asymptotic limit of the weak limit of a minimizing sequence. Fortunately, such a difficulty may be overcome by a local convergence method leading to realizing the weak limit as a monotone solution to the exterior equation and then by an energy comparison. We next prove the uniqueness of a monotone solution to the exterior equation. Since the energy functional is not convex, such a uniqueness result is usually not ensured [11] and seems surprising in the present context. Consequently, we recognize that W1W_{1} is actually this energy minimizer whose energy calculated in Section 3 is thus the energy minimum of the formulated problem. We also illustrate that W2W_{2} is a local energy maximizer among a class of testing configurations, which seems to suggest that, energetically, the closed-form solution W2W_{2} of the exterior monopole equation has a certain ‘significance’ associated as well. In Section 6, we briefly comment on how to obtain oscillatory solutions for both exterior and interior monopole equations using their polar-variable representations.

2 Yang–Mills monopole equations in Reissner–Nordström metric

Consider the Reissner–Nordström metric [22, 37]

ds2=(12Mr+Q2r2)dt2+(12Mr+Q2r2)1dr2+r2(dθ2+sin2θdϕ2),\mathrm{d}s^{2}=-\left(1-\frac{2M}{r}+\frac{Q^{2}}{r^{2}}\right)\,\mathrm{d}t^{2}+\left(1-\frac{2M}{r}+\frac{Q^{2}}{r^{2}}\right)^{-1}\,\mathrm{d}r^{2}+r^{2}(\mathrm{d}\theta^{2}+\sin^{2}\theta\mathrm{d}\phi^{2}), (2.1)

describing a charged black hole of mass MM and charge QQ in the Schwarzschild spherical coordinates. At the critical situation

M=Q,M=Q, (2.2)

the metric (2.1) assumes the form

ds2=(rM)2r2dt2+r2(rM)2dr2+r2(dθ2+sin2θdϕ2).\mathrm{d}s^{2}=-\frac{(r-M)^{2}}{r^{2}}\mathrm{d}t^{2}+\frac{r^{2}}{(r-M)^{2}}\,\mathrm{d}r^{2}+r^{2}(\mathrm{d}\theta^{2}+\sin^{2}\theta\mathrm{d}\phi^{2}). (2.3)

known as the extremal Reissner–Nordström black hole metric with its event horizon at r=Mr=M. We will be interested in both the exterior region, r>Mr>M, and interior one, r<Mr<M, possibly extended to its curvature singularity at r=0r=0 as exhibited by the associated Kretschmann invariant given by

K=RαβμνRαβμν=8M2r8(7M212Mr+6r2),K=R_{\alpha\beta\mu\nu}R^{\alpha\beta\mu\nu}=\frac{8M^{2}}{r^{8}}(7M^{2}-12Mr+6r^{2}), (2.4)

composed from the Riemann tensor induced by the gravitational metric form (2.1).

2.1 Exterior region

We first consider the space region exterior to the event horizon, r>Mr>M. As in [3], we use the variables

τ=t4M,x=ln(rM1),\tau=\frac{t}{4M},\quad x=\ln\left(\frac{r}{M}-1\right), (2.5)

to recast (2.3) into

ds2\displaystyle\mathrm{d}s^{2} =\displaystyle= Ω(x)(dτ2+C4(x)(dx2+dθ2+sin2θdϕ2)),\displaystyle\Omega(x)\left(-\mathrm{d}\tau^{2}+C^{4}(x)(\mathrm{d}x^{2}+\mathrm{d}\theta^{2}+\sin^{2}\theta\,\mathrm{d}\phi^{2})\right), (2.6)
Ω(x)\displaystyle\Omega(x) =\displaystyle= 16M2(1+ex)2C(x)=cosh(x2).\displaystyle\frac{16M^{2}}{(1+\mathrm{e}^{-x})^{2}}\quad C(x)=\cosh\left(\frac{x}{2}\right). (2.7)

Thus the exterior region in terms of the radial variable, r>Mr>M, is converted in terms of the variable xx to the full line <x<-\infty<x<\infty, such that the ‘midway’ spot, x=0x=0, where the conformal factor C(x)C(x) is minimized, corresponds to the classical Schwarzschild radius

rs=2M,r_{\mbox{\small s}}=2M, (2.8)

which will have a specific meaningfulness in our study.

We consider the SU(2)SU(2) Yang–Mills theory over such a spacetime. Use σa\sigma_{a} (a=1,2,3a=1,2,3) to denote the Pauli matrices and set

ta=σa2i,a=1,2,3.t_{a}=\frac{\sigma_{a}}{2\mathrm{i}},\quad a=1,2,3. (2.9)

Then {ta}\{t_{a}\} generates SU(2)SU(2) and satisfies the commutator relation [ta,tb]=ϵabctc[t_{a},t_{b}]=\epsilon_{abc}t_{c}. As in [3], we choose the gauge field to be the 1-form

A=W(τ,x)ω+t3cosθdϕ,ω=t1dθ+t2sinθdϕ,A=W(\tau,x)\,\omega+t_{3}\cos\theta\,\mathrm{d}\phi,\quad\omega=t_{1}\,\mathrm{d}\theta+t_{2}\sin\theta\,\mathrm{d}\phi, (2.10)

where WW is a scalar field. Then we compute to get the Yang–Mills curvature field

F\displaystyle F =\displaystyle= dA+AA\displaystyle\mathrm{d}A+A\wedge A (2.11)
=\displaystyle= t1(Wτdτ+Wxdx)dθ+t2sinθ(Wτdτ+Wxdx)dϕ+t3sinθ(W21)dθdϕ.\displaystyle t_{1}(W_{\tau}\,\mathrm{d}\tau+W_{x}\,\mathrm{d}x)\wedge\mathrm{d}\theta+t_{2}\sin\theta(W_{\tau}\,\mathrm{d}\tau+W_{x}\,\mathrm{d}x)\wedge\mathrm{d}\phi+t_{3}\sin\theta(W^{2}-1)\,\mathrm{d}\theta\wedge\mathrm{d}\phi.\quad

Therefore, with x0=τ,x1=x,x2=θ,x3=ϕx^{0}=\tau,x^{1}=x,x^{2}=\theta,x^{3}=\phi and F=FμνatadxμdxνF=F^{a}_{\mu\nu}t_{a}\,\mathrm{d}x^{\mu}\wedge\mathrm{d}x^{\nu}, we can read (2.11) to obtain the nonzero and independent components of the 2-form FF to be

F021=12Wτ,F121=12Wx,F032=12sinθWτ,F132=12sinθWx,F233=12sinθ(W21).F^{1}_{02}=\frac{1}{2}W_{\tau},\quad F^{1}_{12}=\frac{1}{2}W_{x},\quad F^{2}_{03}=\frac{1}{2}\sin\theta W_{\tau},\quad F^{2}_{13}=\frac{1}{2}\sin\theta W_{x},\quad F^{3}_{23}=\frac{1}{2}\sin\theta(W^{2}-1). (2.12)

On the other hand, if we rewrite line element (2.6)–(2.7) as ds2=gμνdxμdxν\mathrm{d}s^{2}=g_{\mu\nu}\,\mathrm{d}x^{\mu}\mathrm{d}x^{\nu}, then the Yang–Mills action density reads

\displaystyle{\cal L} =\displaystyle= 14FμνagμαgνβFαβa|det(gαβ)|\displaystyle\frac{1}{4}\,F^{a}_{\mu\nu}g^{\mu\alpha}g^{\nu\beta}F^{a}_{\alpha\beta}\sqrt{|\det(g_{\alpha\beta})|} (2.13)
=\displaystyle= (C24Wτ2+14C2Wx2+18C2(1W2)2)sinθ,\displaystyle\left(-\frac{C^{2}}{4}W_{\tau}^{2}+\frac{1}{4C^{2}}W_{x}^{2}+\frac{1}{8C^{2}}(1-W^{2})^{2}\right)\sin\theta,

in view of (2.12). Hence we can rewrite the Yang–Mills action as

L=d4x=2πWdxdτ,L=\int{\cal L}\,\mathrm{d}^{4}x=2\pi\int{\cal L}_{W}\,\mathrm{d}x\mathrm{d}\tau, (2.14)

where

W=C22Wτ2+12C2Wx2+14C2(1W2)2,{\cal L}_{W}=-\frac{C^{2}}{2}W_{\tau}^{2}+\frac{1}{2C^{2}}W_{x}^{2}+\frac{1}{4C^{2}}(1-W^{2})^{2}, (2.15)

as obtained in [3], so that the Yang–Mills equation is the Euler–Lagrange equation of (2.15):

Wττ=C2(C2Wx)x+C4W(1W2),W_{\tau\tau}=C^{-2}(C^{-2}W_{x})_{x}+C^{-4}W(1-W^{2}), (2.16)

whose static limit reads

W′′tanh(x2)W+W(1W2)=0,W^{\prime\prime}-\tanh\left(\frac{x}{2}\right)W^{\prime}+W(1-W^{2})=0, (2.17)

where and in the sequel, we use the notation W=WxW^{\prime}=W_{x}, etc., interchangeably.

2.2 Interior region

We next consider the region that is interior to the event horizon, 0<r<M0<r<M. Thus, using the variables

τ=t4M,x=ln(1rM),0<r<M,\tau=\frac{t}{4M},\quad x=\ln\left(1-\frac{r}{M}\right),\quad 0<r<M, (2.18)

we have <x<0-\infty<x<0 and we see that the line element (2.3) becomes

ds2\displaystyle\mathrm{d}s^{2} =\displaystyle= Ω(x)(dτ2+S4(x)(dx2+dθ2+sin2θdϕ2)),\displaystyle\Omega(x)\left(-\mathrm{d}\tau^{2}+S^{4}(x)(\mathrm{d}x^{2}+\mathrm{d}\theta^{2}+\sin^{2}\theta\mathrm{d}\phi^{2})\right), (2.19)
Ω(x)\displaystyle\Omega(x) =\displaystyle= 16M2(1ex)2,S(x)=sinh(x2).\displaystyle\frac{16M^{2}}{(1-\mathrm{e}^{-x})^{2}},\quad S(x)=\sinh\left(\frac{x}{2}\right). (2.20)

Therefore, from (2.10) and (2.11), we analogously obtain the effective Yang–Mills action density

W=S22Wτ2+12S2Wx2+14S2(1W2)2,{\cal L}_{W}=-\frac{S^{2}}{2}W_{\tau}^{2}+\frac{1}{2S^{2}}W_{x}^{2}+\frac{1}{4S^{2}}(1-W^{2})^{2}, (2.21)

over the spatial interval <x<0-\infty<x<0, with the associated Euler–Lagrange equation

Wττ=S2(S2Wx)x+S4(1W2)W,W_{\tau\tau}={S^{-2}}(S^{-2}W_{x})_{x}+{S^{-4}}(1-W^{2})W, (2.22)

whose static limit reads

W′′coth(x2)W+W(1W2)=0.W^{\prime\prime}-\coth\left(\frac{x}{2}\right)W^{\prime}+W(1-W^{2})=0. (2.23)

The equations (2.22) and (2.23) are a new pair of equations, accompanying (2.16) and (2.17).

3 Exact solutions

In [3], the following two nontrivial exact solutions to the exterior Yang–Mills monopole equation (2.17) are obtained:

W1(x)=tanh(x2),W2(x)=2coshx262coshx+4+36.W_{1}(x)=\tanh\left(\frac{x}{2}\right),\quad W_{2}(x)=\frac{2\cosh x-2-\sqrt{6}}{2\cosh x+4+3\sqrt{6}}. (3.1)

Here we find that there is a third exact solution,

W3(x)=2coshx2+62coshx+436,W_{3}(x)=\frac{2\cosh x-2+\sqrt{6}}{2\cosh x+4-3\sqrt{6}}, (3.2)

which is apparently singular at

x±=±ln(362+232322)±1.104268225,x_{\pm}=\pm\ln\left(\frac{3\sqrt{6}}{2}+2\sqrt{3}-2-\frac{3\sqrt{2}}{2}\right)\approx\pm 1.104268225, (3.3)

giving rise to the corresponding singular radii

r±=(1+ex±)M,r_{\pm}=\left(1+\mathrm{e}^{x_{\pm}}\right)M, (3.4)

in the original radial variable, which are distributed about the Schwarzschild radius rsr_{\mbox{\small s}} following

M<r<rs=2M<r+,r+r+2=(1+cosh(x+))M>rs.M<r_{-}<r_{\mbox{\small s}}=2M<r_{+},\quad\quad\frac{r_{-}+r_{+}}{2}=\left(1+\cosh(x_{+})\right)M>r_{\mbox{\small s}}. (3.5)

For the interior Yang–Mills monopole equation (2.23), we likewise find the following three nontrivial exact solutions:

w1(x)\displaystyle w_{1}(x) =\displaystyle= coth(x2),\displaystyle\coth\left(\frac{x}{2}\right), (3.6)
w2(x)\displaystyle w_{2}(x) =\displaystyle= 2coshx+262coshx4+36,\displaystyle\frac{2\cosh x+2-\sqrt{6}}{2\cosh x-4+3\sqrt{6}}, (3.7)
w3(x)\displaystyle w_{3}(x) =\displaystyle= 2coshx+2+62coshx436,\displaystyle\frac{2\cosh x+2+\sqrt{6}}{2\cosh x-4-3\sqrt{6}}, (3.8)

for <x<0-\infty<x<0. The solutions w1w_{1} and w2w_{2} given in (3.6) and (3.7), respectively, are regular but w3w_{3} in (3.6) is singular at

x0=ln(2[1+3]+32[1+3]),x_{0}=-\ln\left(2[1+\sqrt{3}]+\frac{3}{\sqrt{2}}[1+\sqrt{3}]\right), (3.9)

whereby the irrelevant positive root is discarded.

We now compute the energies of these solutions.

Recall that the energy-momentum tensor induced from (2.13) is

Tμν=FμαagαβFνβa+14gμν(FγδagγαgδβFαβa),T_{\mu\nu}=-F^{a}_{\mu\alpha}g^{\alpha\beta}F^{a}_{\nu\beta}+\frac{1}{4}g_{\mu\nu}(F^{a}_{\gamma\delta}g^{\gamma\alpha}g^{\delta\beta}F^{a}_{\alpha\beta}), (3.10)

which via =T00=g00T00{\cal H}=T^{0}_{0}=g^{00}T_{00} gives rise to the Hamiltonian energy density associated with (2.6)–(2.7) and (2.12) to be

W=C22Wτ2+12C2Wx2+14C2(1W2)2.{\cal H}_{W}=\frac{C^{2}}{2}W_{\tau}^{2}+\frac{1}{2C^{2}}W_{x}^{2}+\frac{1}{4C^{2}}(1-W^{2})^{2}. (3.11)

Thus, evaluating the static total energy

E(W)=2π(12C2Wx2+14C2(1W2)2)dx,E(W)=2\pi\int_{-\infty}^{\infty}\left(\frac{1}{2C^{2}}W_{x}^{2}+\frac{1}{4C^{2}}(1-W^{2})^{2}\right)\,\mathrm{d}x, (3.12)

we obtain the exact value

E(W1)=8π5,E(W_{1})=\frac{8\pi}{5}, (3.13)

for the solution W1W_{1} given in (3.1). For the solution W2W_{2} given in (3.1), utilizing MAPLE, we obtain the exact value

E(W2)=2π(1219870423+30388736)tanh1(532+3)+157071962+132222883353685723+430790492,E(W_{2})=2\pi\frac{(12198704\sqrt{2}\sqrt{3}+30388736)\tanh^{-1}\left(\frac{5}{3\sqrt{2}+\sqrt{3}}\right)+15707196\sqrt{2}+13222288\sqrt{3}}{35368572\sqrt{3}+43079049\sqrt{2}}, (3.14)

which is approximately 2π(0.9664005272)2\pi(0.9664005272). These two exact results are consistent with those found in [3] based on numerical evaluation. Not surprisingly, the singular solution W3(x)W_{3}(x) is of infinite energy. In other words, this solution behaves singularly in both point-wise and energy-wise manners at x±x_{\pm}.

We now turn our attention to the interior solutions listed in (3.6)–(3.8). In view of (3.10), (2.12), and (2.19)–(2.20), we get the Hamiltonian energy density

W=S22Wτ2+12S2Wx2+14S2(1W2)2,{\cal H}_{W}=\frac{S^{2}}{2}W_{\tau}^{2}+\frac{1}{2S^{2}}W_{x}^{2}+\frac{1}{4S^{2}}(1-W^{2})^{2}, (3.15)

which renders the total energy for a static interior solution in the full region 0<r<M0<r<M to be

E=2π0(12S2Wx2+14S2(1W2)2)dx.E=2\pi\int_{-\infty}^{0}\left(\frac{1}{2S^{2}}W_{x}^{2}+\frac{1}{4S^{2}}(1-W^{2})^{2}\right)\,\mathrm{d}x. (3.16)

All the solutions given in (3.6)–(3.8) are of infinity energy due to the divergence of the energy density (3.15) as a consequence of the curvature singularity of the Reissner–Nordström black hole spacetime at r=0r=0 displayed by (2.4). Nevertheless, we may assume that we are in a situation that gravity is generated from a bulk of matter which occupies a finite region within the event horizon,

rrB,0<rB<M.r\leq r_{\mbox{\tiny B}},\quad 0<r_{\mbox{\tiny B}}<M. (3.17)

Then we can focus on the ‘regularized’ region

<x<xB<0,xB=ln(1rBM),rBr<M,-\infty<x<x_{\mbox{\tiny B}}<0,\quad x_{\mbox{\tiny B}}=\ln\left(1-\frac{r_{\mbox{\tiny B}}}{M}\right),\quad r_{\mbox{\tiny B}}\leq r<M, (3.18)

instead. It is seen that the solutions (3.6) and (3.7) are of finite energies over (3.18) for any 0<rB<M0<r_{\mbox{\tiny B}}<M and the solution (3.8) is of finite energy only when xBx_{\mbox{\tiny B}} is below its singular point x0x_{0} spelled out in (3.9). Thus, we are led to the finite-energy condition

xB<ln(2[1+3]+32[1+3])2.421226122,x_{\mbox{\tiny B}}<-\ln\left(2[1+\sqrt{3}]+\frac{3}{\sqrt{2}}[1+\sqrt{3}]\right)\approx-2.421226122, (3.19)

or equivalently,

rB>((3+33)2+43+2(3+33)2+43+4)M(0.9111873445)M.r_{\mbox{\tiny B}}>\left(\frac{(3+3\sqrt{3})\sqrt{2}+4\sqrt{3}+2}{(3+3\sqrt{3})\sqrt{2}+4\sqrt{3}+4}\right)M\approx(0.9111873445)M. (3.20)

Over a given interval (3.18), the energy of the solution (3.6) is generically much greater than that of the solution (3.7), since the former is singular at its right boundary point, x=0x=0, and the latter is everywhere regular. As a comparison, we list our calculation of the energy

E(W)=2πxB(12S2Wx2+14S2(1W2)2)dx,E(W)=2\pi\int_{-\infty}^{x_{\mbox{\tiny B}}}\left(\frac{1}{2S^{2}}W_{x}^{2}+\frac{1}{4S^{2}}(1-W^{2})^{2}\right)\,\mathrm{d}x, (3.21)

for xB=1x_{\mbox{\tiny B}}=-1 for the solutions w1w_{1} and w2w_{2}, given in (3.6) and (3.7), respectively, as follows:

E(w1)=2π(3.273953196),E(w2)=2π(0.3220989496).E(w_{1})=2\pi(3.273953196),\quad E(w_{2})=2\pi(0.3220989496). (3.22)

4 Energy consideration of exterior monopoles

We now investigate the problem whether there is a nontrivial energetically stable exterior monopole solution linking two vacuum states, W=±1W=\pm 1, characterized by the boundary condition W()=±1W(-\infty)=\pm 1 and W()=±1W(\infty)=\pm 1. It is clear that the problem of linking the same vacuum state with W()=W()=±1W(-\infty)=W(\infty)=\pm 1 is not well posed without some further elaboration since the vacuum states themselves minimize the energy and may be approached arbitrarily by nontrivial field configurations. Thus it remains to consider the problem of linking two different vacuum states, say,

η=inf{E(W)|W()=1,W()=1},\eta=\inf\{E(W)\,|\,W(-\infty)=-1,W(\infty)=1\}, (4.1)

and ask whether this minimization problem has a solution, where E(W)E(W) is as given in (3.12). This is typically a domain wall problem [6, 7, 8, 26, 36], interpolating two ground state domains, W=±1W=\pm 1. However, it is not hard to see that this problem is not well posed either.

In fact, let u(x)u(x) be a smooth function satisfying

u(x)=1,x1;u(x)=1,x1.u(x)=-1,\quad x\leq-1;\quad u(x)=1,\quad x\geq 1. (4.2)

Then, using u(a)=u(xa)u_{(a)}=u(x-a) as a testing function, we have

E(u(a))\displaystyle E(u_{(a)}) =\displaystyle= πa1a+11C2(x)((u(a))2+12(u(a)21)2)dx\displaystyle\pi\int_{a-1}^{a+1}\frac{1}{C^{2}(x)}\left((u^{\prime}_{(a)})^{2}+\frac{1}{2}(u^{2}_{(a)}-1)^{2}\right)\,\mathrm{d}x (4.3)
\displaystyle\leq K0a1a+1dxC2(x)0as |a|.\displaystyle K_{0}\int_{a-1}^{a+1}\frac{\mathrm{d}x}{C^{2}(x)}\to 0\quad\mbox{as }|a|\to\infty.

Here K0>0K_{0}>0 is a constant depending on the properties of uu over [1,1][-1,1]. As a consequence, this shows the quantity η\eta in (4.1) is zero which is therefore not attainable.

On the other hand, since the boundary condition in (4.1) dictates that the unit magnetic charge monopole configuration, W=0W=0, is to occur somewhere as an intermediate state, we may conveniently impose the following additional ‘intermediate state condition’

W(0)=0.W(0)=0. (4.4)

It is interesting that this condition implies that the potential part of the energy density

W=12C2(x)Wx2+14C2(x)(1W2)2{\cal H}_{W}=\frac{1}{2C^{2}(x)}W_{x}^{2}+\frac{1}{4C^{2}(x)}(1-W^{2})^{2} (4.5)

is maximized at the spot where the intermediate state occurs, x=0x=0. In Figure 1, we plot the energy density (4.5) for the solutions W1W_{1} and W2W_{2} given in (3.1) as an illustration. It is interesting to notice that the former peaks sharply at x=0x=0 where the solution passes the intermediate state and the latter also peaks at the same spot which is not where the solution passes the intermediate state but is the midpoint between the two spots,

x0±=ln(1+62±6+462),x^{\pm}_{0}=\ln\left(1+\frac{\sqrt{6}}{2}\pm\frac{\sqrt{6+4\sqrt{6}}}{2}\right), (4.6)

where the solution passes the intermediate state. Such a property of energy concentration is typical for solitons in field theories.

Refer to caption
Figure 1: Plots of the energy distributions of the solutions linking two different vacuum states and the same vacuum state and the unit-charge monopole state viewed as an intermediate state. The former is depicted by a solid-line curve showing the concentration of energy at the spot where the solution passes through the intermediate state. The latter is drawn by a dash-line curve exhibiting the concentration of energy in the middle of the two spots where the solution passes through the intermediate state. Energetically both solutions demonstrate themselves as domain-wall solitons.

These pictures coincide with what we know about a domain wall soliton configuration in general.

Thus, we are led to modifying (4.1) into the minimization problem

η0inf{E(W)|W()=1,W()=1,W(0)=0}.\eta_{0}\equiv\inf\{E(W)\,|\,W(-\infty)=-1,W(\infty)=1,W(0)=0\}. (4.7)

We shall show that W1W_{1} given in (3.1) is the unique solution to this problem. Consequently, as a by-product, we have

η0=8π5,\eta_{0}=\frac{8\pi}{5}, (4.8)

in view of (3.13).

Let WW be a finite-energy critical point of the energy functional (3.12). Set Wλ(x)=W(λx)W_{\lambda}(x)=W(\lambda x). Then, from ddλE(Wλ)|λ=1=0\frac{\mathrm{d}}{\mathrm{d}\lambda}E(W_{\lambda})|_{\lambda=1}=0, we get the virial identity

1C2(x)([1+xtanh(x2)]Wx2+x2tanh(x2)(W21)2)dx=(W21)22C2(x)dx,\int_{-\infty}^{\infty}\frac{1}{C^{2}(x)}\left(\left[1+x\tanh\left(\frac{x}{2}\right)\right]W_{x}^{2}+\frac{x}{2}\,\tanh\left(\frac{x}{2}\right)(W^{2}-1)^{2}\right)\,\mathrm{d}x=\int_{-\infty}^{\infty}\frac{(W^{2}-1)^{2}}{2C^{2}(x)}\,\mathrm{d}x, (4.9)

which is an energy partition relation. This identity demonstrates that the ‘potential energy’ is much greater than the ‘elastic energy’ for an exterior monopole. For example, for the exact solution W=W1W=W_{1} in (3.1), we have

(W21)24C2(x)dx=815,Wx22C2(x)dx=415,\int_{-\infty}^{\infty}\frac{(W^{2}-1)^{2}}{4C^{2}(x)}\,\mathrm{d}x=\frac{8}{15},\quad\int_{-\infty}^{\infty}\frac{W_{x}^{2}}{2C^{2}(x)}\,\mathrm{d}x=\frac{4}{15}, (4.10)

and, for W=W2W=W_{2} given in (3.1), these quantities are approximately 0.87651545750.8765154575 and 0.089885070220.08988507022, respectively.

5 Existence and uniqueness of exterior energy minimizer

The symmetry of the functional (3.12) indicates that the problem (4.7) amounts to considering the minimization problem

I0inf{I(W)01C2(x)((W)2+12(W21)2)dx|W(0)=0,W()=1},I_{0}\equiv\inf\left\{I(W)\equiv\int_{0}^{\infty}\frac{1}{C^{2}(x)}\left((W^{\prime})^{2}+\frac{1}{2}(W^{2}-1)^{2}\right)\,\mathrm{d}x\,\bigg{|}\,W(0)=0,W(\infty)=1\right\}, (5.1)

over the set of admissible functions which are absolutely continuous on all compact subintervals of the half line [0,)[0,\infty).

Let {𝒲n}\{{\cal W}_{n}\} be a minimizing sequence of (5.1) satisfying

limnI(𝒲n)=I0;I(𝒲n)I0+1n,n=1,2,.\lim_{n\to\infty}I({\cal W}_{n})=I_{0};\quad I({\cal W}_{n})\leq I_{0}+\frac{1}{n},\quad n=1,2,\dots. (5.2)

By modifying the sequence {𝒲n}\{{\cal W}_{n}\} if necessary, we may assume that {𝒲n}\{{\cal W}_{n}\} enjoys the property that each function 𝒲n{\cal W}_{n} satisfies 0𝒲n(x)10\leq{\cal W}_{n}(x)\leq 1. Furthermore, for each 𝒲n{\cal W}_{n}, we may modify it such that it minimizes the partial functional

In(W)0n1C2(x)((W)2+12(W21)2)dx,I_{n}(W)\equiv\int_{0}^{n}\frac{1}{C^{2}(x)}\left((W^{\prime})^{2}+\frac{1}{2}(W^{2}-1)^{2}\right)\,\mathrm{d}x, (5.3)

among functions fulfilling the boundary condition W(0)=0,W(n)=𝒲n(n)W(0)=0,W(n)={\cal W}_{n}(n) since the functional is differentiable and weakly lower semicontinuous over the Sobolev space W1,2(0,n)W^{1,2}(0,n), say (cf. Theorem 2 on page 448 in Evans[11] specifically as well as the studies in [12, 35, 40] in other contexts). Therefore 𝒲n{\cal W}_{n} when restricted to the interval [0,n][0,n] is a critical point of InI_{n}. As a consequence, we have

0n1C2(x)(𝒲nw+(𝒲n21)𝒲nw)dx=0,wC01(0,n).\int_{0}^{n}\frac{1}{C^{2}(x)}\left({\cal W}^{\prime}_{n}w^{\prime}+({\cal W}_{n}^{2}-1){\cal W}_{n}w\right)\,\mathrm{d}x=0,\quad\forall w\in C_{0}^{1}(0,n). (5.4)

Extracting a suitable subsequence if necessary, for example, using a diagonal subsequence argument, we may assume that {𝒲n}\{{\cal W}_{n}\} is weakly convergent in the Sobolev space W1,2(0,a)W^{1,2}(0,a), with the weighted measure C2(x)dxC^{-2}(x)\,\mathrm{d}x, for any a>0a>0. Let 𝒲{{\cal W}} denote such a weak limit which is well defined over [0,)[0,\infty). Fix a>0a>0. Neglecting the first few terms of the sequence if necessary, we may also assume n>an>a. Thus (5.4) gives us

0a1C2(x)(𝒲nw+(𝒲n21)𝒲nw)dx=0,wC01(0,a).\int_{0}^{a}\frac{1}{C^{2}(x)}\left({\cal W}^{\prime}_{n}w^{\prime}+({\cal W}_{n}^{2}-1){\cal W}_{n}w\right)\,\mathrm{d}x=0,\quad\forall w\in C_{0}^{1}(0,a). (5.5)

Using the compact embedding from the Sobolev space into the space C[0,a]C[0,a] and letting nn\to\infty in (5.5), we arrive at

0a1C2(x)(𝒲w+(𝒲21)𝒲w)dx=0,wC01(0,a),\int_{0}^{a}\frac{1}{C^{2}(x)}\left({{\cal W}}^{\prime}w^{\prime}+({{\cal W}}^{2}-1){{\cal W}}w\right)\,\mathrm{d}x=0,\quad\forall w\in C_{0}^{1}(0,a), (5.6)

which leads us to conclude by the standard elliptic regularity theory [12, 18] that 𝒲{{\cal W}} is a classical solution to the Euler–Lagrange equation of the functional II. Hence 𝒲{\cal W} satisfies

(𝒲C2(x))+1C2(x)(1𝒲2)𝒲=0,𝒲(0)=0.\left(\frac{{\cal W}^{\prime}}{C^{2}(x)}\right)^{\prime}+\frac{1}{C^{2}(x)}(1-{\cal W}^{2}){\cal W}=0,\quad{\cal W}(0)=0. (5.7)

Since {𝒲n}\{{\cal W}_{n}\} is a sequence with values confined in [0,1][0,1], so is its limit 𝒲{\cal W}. However, since 𝒲{\cal W} satisfies (5.7), we see that 𝒲0{\cal W}\equiv 0 or 0<𝒲(x)<10<{\cal W}(x)<1 for x>0x>0 because 𝒲0{\cal W}\equiv 0 and 𝒲1{\cal W}\equiv 1 are two equilibria of the differential equation in (5.7). We now show that the former does not occur.

In fact, by using weak convergence, we have

Ia(𝒲)lim infnIa(𝒲n)limnI(𝒲n)=I0,I_{a}({\cal W})\leq\liminf_{n\to\infty}I_{a}({\cal W}_{n})\leq\lim_{n\to\infty}I({\cal W}_{n})=I_{0}, (5.8)

for any a>0a>0. Thus, letting aa\to\infty in (5.8), we get

I(𝒲)I0.I({\cal W})\leq I_{0}. (5.9)

Recall the result

I0I(tanhx2)=45.I_{0}\leq I\left(\tanh\,\frac{x}{2}\right)=\frac{4}{5}. (5.10)

Besides, we also have I(0)=1I(0)=1. Hence 𝒲0{\cal W}\not\equiv 0. Consequently, 𝒲{\cal W} satisfies 0<𝒲(x)<10<{\cal W}(x)<1 for x>0x>0.

Since 𝒲{\cal W} is bounded, we deduce that there is a sequence {xn}\{x_{n}\}, xnx_{n}\to\infty as nn\to\infty, such that 𝒲(xn)0{\cal W}^{\prime}(x_{n})\to 0 as nn\to\infty. Using this result as the boundary condition and integrating the differential equation in (5.7), we have

𝒲(x)C2(x)=x1C2(y)(1𝒲2(y))𝒲(y)dy.\frac{{\cal W}^{\prime}(x)}{C^{2}(x)}=\int^{\infty}_{x}\frac{1}{C^{2}(y)}(1-{\cal W}^{2}(y)){\cal W}(y)\,\mathrm{d}y. (5.11)

Hence 𝒲(x)>0{\cal W}^{\prime}(x)>0 for x>0x>0. In particular,

limx𝒲(x)=L,\lim_{x\to\infty}{\cal W}(x)=L, (5.12)

for some L(0,1]L\in(0,1]. Thus, if L<1L<1 in (5.12), we infer from (5.11) that

limx𝒲(x)=(1L2)L>0,\lim_{x\to\infty}{\cal W}^{\prime}(x)=(1-L^{2})L>0, (5.13)

which is false. Therefore L=1L=1 in (5.12). This establishes that 𝒲{\cal W} lies in the admissible space of the problem (5.1). Thus I(𝒲)I0I({\cal W})\geq I_{0}. Combining this result with (5.9), we arrive at

I(𝒲)=I0.I({\cal W})=I_{0}. (5.14)

In other words, the existence of a solution to the minimization problem (5.1) follows.

We next prove that, actually,

𝒲(x)=tanh(x2).{\cal W}(x)=\tanh\left(\frac{x}{2}\right). (5.15)

Thus, in particular, equality in (5.10) holds. We achieve this goal by showing that a positive solution to the initial value problem (5.7), with a limiting value at infinity, is unique.

In fact, let 𝒲1{\cal W}_{1} and 𝒲2{\cal W}_{2} be two such solutions. Then w=𝒲1𝒲2w={\cal W}_{1}-{\cal W}_{2} satisfies

(wC2(x))+wC2(x)(1𝒲12𝒲22𝒲1𝒲2)=0.\left(\frac{w^{\prime}}{C^{2}(x)}\right)^{\prime}+\frac{w}{C^{2}(x)}(1-{\cal W}_{1}^{2}-{\cal W}_{2}^{2}-{\cal W}_{1}{\cal W}_{2})=0. (5.16)

On the other hand, for any differentiable function hh, we have

C2(x)(𝒲1𝒲1h2C2(x))+𝒲12(h)2+(𝒲1𝒲2+𝒲22)𝒲12h2\displaystyle C^{2}(x)\left(\frac{{\cal W}_{1}{\cal W}^{\prime}_{1}h^{2}}{C^{2}(x)}\right)^{\prime}+{{\cal W}_{1}^{2}(h^{\prime})^{2}}+{({\cal W}_{1}{\cal W}_{2}+{\cal W}_{2}^{2}){\cal W}_{1}^{2}h^{2}}
=([𝒲1h])2(1𝒲12𝒲22𝒲1𝒲2)(𝒲1h)2.\displaystyle=([{\cal W}_{1}h]^{\prime})^{2}-(1-{\cal W}_{1}^{2}-{\cal W}_{2}^{2}-{\cal W}_{1}{\cal W}_{2})({\cal W}_{1}h)^{2}. (5.17)

Now substituting h=w𝒲1h=\frac{w}{{\cal W}_{1}} into (5), we have

(w)2C21C2(1𝒲12𝒲22𝒲1𝒲2)w2\displaystyle\frac{(w^{\prime})^{2}}{C^{2}}-\frac{1}{C^{2}}(1-{\cal W}_{1}^{2}-{\cal W}_{2}^{2}-{\cal W}_{1}{\cal W}_{2})w^{2}
=(𝒲1w2C2𝒲1)+(w𝒲1w𝒲1)2C2+(𝒲1𝒲2+𝒲22)w2C2.\displaystyle=\left(\frac{{\cal W}_{1}^{\prime}w^{2}}{C^{2}{\cal W}_{1}}\right)^{\prime}+\frac{(w^{\prime}{\cal W}_{1}-w{\cal W}_{1}^{\prime})^{2}}{C^{2}}+\frac{({\cal W}_{1}{\cal W}_{2}+{\cal W}_{2}^{2})w^{2}}{C^{2}}. (5.18)

Moreover, using the Cauchy–Kovalevskaya theorem, we know that the solution of (5.7) has the asymptotic form

𝒲(x)=bx+cx3+O(x5),x0.{\cal W}(x)=bx+cx^{3}+\mbox{O}(x^{5}),\quad x\sim 0. (5.19)

As a consequence of (5.19), we may integrate (5) and drop the resulting boundary terms to arrive at

0((w)2C21C2(1𝒲12𝒲22𝒲1𝒲2)w2)dx\displaystyle\int_{0}^{\infty}\left(\frac{(w^{\prime})^{2}}{C^{2}}-\frac{1}{C^{2}}(1-{\cal W}_{1}^{2}-{\cal W}_{2}^{2}-{\cal W}_{1}{\cal W}_{2})w^{2}\right)\,\mathrm{d}x
=01C2(x)((w𝒲1w𝒲1)2+(𝒲1𝒲2+𝒲22)w2)dx.\displaystyle=\int_{0}^{\infty}\frac{1}{C^{2}(x)}\left({(w^{\prime}{\cal W}_{1}-w{\cal W}_{1}^{\prime})^{2}}+{({\cal W}_{1}{\cal W}_{2}+{\cal W}_{2}^{2})w^{2}}\right)\,\mathrm{d}x. (5.20)

However, by virtue of (5.16), we may rewrite the left-hand side of (5) as

0((w)2C21C2(1𝒲12𝒲22𝒲1𝒲2)w2)dx\displaystyle\int_{0}^{\infty}\left(\frac{(w^{\prime})^{2}}{C^{2}}-\frac{1}{C^{2}}(1-{\cal W}_{1}^{2}-{\cal W}_{2}^{2}-{\cal W}_{1}{\cal W}_{2})w^{2}\right)\,\mathrm{d}x
=(wwC2)00w([wC2]+wC2(1𝒲12𝒲22𝒲1𝒲2))dx=0.\displaystyle=\left(\frac{ww^{\prime}}{C^{2}}\right)_{0}^{\infty}-\int_{0}^{\infty}w\left(\left[\frac{w^{\prime}}{C^{2}}\right]^{\prime}+\frac{w}{C^{2}}(1-{\cal W}_{1}^{2}-{\cal W}_{2}^{2}-{\cal W}_{1}{\cal W}_{2})\right)\,\mathrm{d}x=0. (5.21)

Combining (5) and (5), we conclude with w0w\equiv 0.

Hence the existence and uniqueness of an exterior energy-minimizing monopole solution linking different vacuum states at x±x\pm\infty and passes through the unit-charge monopole state at x=0x=0 is established which is the exact solution W1W_{1} stated in (3.1) and found in [3] which determines the minimum energy, Emin=8π5E_{\min}=\frac{8\pi}{5}, as stated in (3.13).

It is interesting to note that a rich family of even and odd solutions of (2.17) linking the same and different vacuum states is obtained in [3] which are characterized by the number nn of zeros they possess. When n=1n=1, their calculated energy coincides with our exact value; when n2n\geq 2, their calculated energies are all significantly greater than that with n=1n=1.

Motivated by the construction in this section and the even solutions with multiple zeros linking to the same vacuum state, W()=W()=±1W(-\infty)=W(\infty)=\pm 1, obtained in [3], we may impose the following minimization problem

σ0inf{E(W)|W()=W()=1,W(0)0},\sigma_{0}\equiv\inf\{E(W)\,|\,W(-\infty)=W(\infty)=1,W(0)\leq 0\}, (5.22)

over the admissible set of functions which are absolutely continuous on all compact subintervals of the real line. Thus, with the function (5.15) as a testing function, i.e., choosing W(x)=|tanh(x2)|W(x)=|\tanh(\frac{x}{2})| for all xx, we have the estimate

σ08π5.\sigma_{0}\leq\frac{8\pi}{5}. (5.23)

With the problem (5.22) in mind, we consider the exact solution W2W_{2} given in (3.1) whose value at x=0x=0 is

W2(0)=h0,h0=13+6,W_{2}(0)=-h_{0},\quad h_{0}=\frac{1}{3+\sqrt{6}}, (5.24)

and we ask what this ‘height’ h0h_{0} means with regard to the problem (5.22). So we take the trial function

Wh(x)=2coshxd2coshx+4+36,d2+3(2+6)h,0<h<1,W_{h}(x)=\frac{2\cosh x-d}{2\cosh x+4+3\sqrt{6}},\quad d\equiv 2+3(2+\sqrt{6})h,\quad 0<h<1, (5.25)

which satisfies Wh(0)=hW_{h}(0)=-h and is suggested by the form of W2(x)W_{2}(x) in (3.1). Hence we can compute the energy E(Wh)E(W_{h}). Surprisingly, we find that E(Wh)E(W_{h}) is maximized at h=h0h=h_{0} as illustrated in Figure 2.

Refer to caption
Figure 2: A plot of the normalized energy E(Wh)2π\frac{E(W_{h})}{2\pi} against d=2+3(2+6)hd=2+3(2+\sqrt{6})h for 16<h<15\frac{1}{6}<h<\frac{1}{5} such that the critical height h0=13+6h_{0}=\frac{1}{3+\sqrt{6}} corresponds to d0=12+563+64.45d_{0}=\frac{12+5\sqrt{6}}{3+\sqrt{6}}\approx 4.45 where the energy peaks with the signature value, 0.96640.9664, first obtained in [3].

Therefore, the quantity h0h_{0} may be regarded as a ‘canonical height’ at which the energy confined over the profile family of functions of the form (5.25) is maximized and brings forth a critical point of the energy functional.

6 Remarks on oscillatory solutions

In [3], a family of solutions of (2.17) with prescribed numbers of zeros are obtained. In this section, we revisit such a construction of oscillatory solutions from the viewpoint of recasting the equation into a form of its ‘polar variable’ representation such that the appearance of multiple zeros of the solutions becomes somewhat more transparent as we now show.

For our purpose, we use the following polar variable anzatz

W=ρcosω,W=ρsinω,W=\rho\cos{\omega},\quad W^{\prime}=\rho\sin{\omega}, (6.1)

to represent a solution of (2.17). Thus, we see that the pair (ρ,ω)(\rho,{\omega}) satisfies the coupled system of equations

ρ\displaystyle\rho^{\prime} =\displaystyle= tanh(x2)ρsin2ω+ρ3cos3ωsinω,\displaystyle\tanh\left(\frac{x}{2}\right)\,\rho\sin^{2}{\omega}+\rho^{3}\cos^{3}{\omega}\sin{\omega}, (6.2)
ω\displaystyle{\omega}^{\prime} =\displaystyle= 1+tanh(x2)sinωcosω+ρ2cos4ω.\displaystyle-1+\tanh\left(\frac{x}{2}\right)\,\sin{\omega}\cos{\omega}+\rho^{2}\cos^{4}{\omega}. (6.3)

To justify this representation, we observe in view of the uniqueness theorem for solutions of the initial value problems of ordinary differential equations that if ρ\rho satisfying (6.2) is positive at a point then it will stay positive where it exists. So, we may impose the initial condition

ρ(0)=a>0,ω(0)=b.\rho(0)=a>0,\quad{\omega}(0)=b. (6.4)

Thus, in particular, to produce an even solution of (2.17), we take b=0b=0, and an odd solution, we impose b=π2b=\frac{\pi}{2}. In the subsequent discussion, we fix bb and adjust aa to achieve multiple zeros for the field WW. Accordingly, in view of (6.1), such a goal may be achieved by making the angular variable ω{\omega} circulate through the full circle multiple times, in a sense to be made precise below.

Indeed, we fix b,K0>0b,K_{0}>0. Then the theorem of continuous dependence of solutions on the initial data allows us to get some number a0>0a_{0}>0 such that the solution to (6.2)–(6.4) satisfies

0<ρ(x)12,0xK0,0<\rho(x)\leq\frac{1}{2},\quad 0\leq x\leq K_{0}, (6.5)

when a(0,a0]a\in(0,a_{0}]. Applying (6.5) to (6.3) and using sinωcosω12\sin{\omega}\cos{\omega}\leq\frac{1}{2}, we get

ω(x)14,0xK0.{\omega}^{\prime}(x)\leq-\frac{1}{4},\quad 0\leq x\leq K_{0}. (6.6)

Therefore, for any integer n1n\geq 1, we may set K0=4nπK_{0}=4n\pi (say). Then (6.6) indicates that there are nn points x1,,xnx_{1},\dots,x_{n} in (0,K0](0,K_{0}] such that

ω(xi)=an odd multiple of π2,i=1,,n.{\omega}(x_{i})=\mbox{an odd multiple of }\,\frac{\pi}{2},\quad i=1,\dots,n. (6.7)

As a consequence of this observation and (6.1), we see that if bb is not an odd multiple of π2\frac{\pi}{2}, then x1,,xnx_{1},\dots,x_{n} are the zeros of WW, and if bb is an odd multiple of π2\frac{\pi}{2}, then b,x1,,xnb,x_{1},\dots,x_{n} are the zeros of WW. Hence a construction of a local solution with a prescribed number of zeros is obtained. Note that such a solution may be even, odd, or neither.

For the interior monopole equation (2.23), if we use the polar representatiion (6.1) with the flippled variable x=yx=-y, then we get the transformed governing system

ρ\displaystyle\rho^{\prime} =\displaystyle= coth(y2)ρsin2ωρ3cos3ωsinω,\displaystyle\coth\left(\frac{y}{2}\right)\,\rho\sin^{2}{\omega}-\rho^{3}\cos^{3}{\omega}\sin{\omega}, (6.8)
ω\displaystyle{\omega}^{\prime} =\displaystyle= 1+coth(y2)sinωcosωρ2cos4ω,\displaystyle 1+\coth\left(\frac{y}{2}\right)\,\sin{\omega}\cos{\omega}-\rho^{2}\cos^{4}{\omega}, (6.9)

where 0<y<0<y<\infty and ρ=dρdy\rho^{\prime}=\frac{\mathrm{d}\rho}{\mathrm{d}y}, etc. Although coth(y2)\coth(\frac{y}{2}) is unbounded as y0y\to 0, corresponding to the curvature singularity, we may again restrict our attention to a region away from the singularity in order to get solutions with multiple zeros in WW. For example, it suffices to request coth(y2)<2\coth(\frac{y}{2})<2 for our method here to work. Such a condition translates into

y>ln3.y>\ln 3. (6.10)

In view of (2.18), this restriction gives us the interior region

2M3<r<M,\frac{2M}{3}<r<M, (6.11)

to accommodate oscillatory monopole solutions similar to those in the exterior region.

In order to obtain a global solution with a prescribed number of zeros with desired asymptotic limits at infinity, W(±)=±1W(\pm\infty)=\pm 1, additional work is needed for the choice of the initial data. Here we omit the details since much of the idea along the line is presented in [3] mathematically and numerically. See also [15, 21, 31].

The authors would like to thank an anonymous referee whose comments and constructive suggestions helped improve the presentation of this paper.

Data availability statement: The data that supports the findings of this study are available within the article.

References

  • [1] A. Actor, Classical solutions of SU(2)SU(2) Yang–Mills theories, Rev. Mod. Phys. 51 (1979) 461–525.
  • [2] R. Bartnik and J. McKinnon, Particlelike solutions of the Einstein–Yang–Mills equations Phys. Rev. Lett. 61 (1988) 141–144.
  • [3] P. Bizoń and M. Kahl, A Yang–Mills field on the extremal Reissner–Nordström black hole, Class. Quantum Grav. 33 (2016) 175013.
  • [4] M. Born and L. Infeld, Foundation of the new field theory, Nature 132 (1933) 1004.
  • [5] M. Born and L. Infeld, Foundation of the new field theory, Proc. Roy. Soc. (London) A 144 (1934) 425–451.
  • [6] L. Cao, S. Chen, and Y. Yang, Domain wall solitons arising in classical gauge field theories, Commun. Math. Phys. 369 (2019) 317–349.
  • [7] S. Chen and Y. Yang, Phase transition solutions in geometrically constrained magnetic domain wall models, J. Math. Phys. 51 (2010) 023504.
  • [8] S. Chen and Y. Yang, Domain wall equations, Hessian of superpotential, and Bogomol’nyi bounds, Nucl. Phys. B 904 (2016) 470–493.
  • [9] P. Curie, Sur la possibilité d’existence de la conductibilité magnétique et du magnétisme libre, Séances de la Société Française de Physique (Paris) (1894) 76–77.
  • [10] P. Dirac, Quantised singularities in the electromagnetic field, Proc. Roy. Soc. (London) A 133 (1931) 60–72.
  • [11] L. C. Evans, Partial Differential Equations, Graduate Studies in Mathematics 19, Amer. Math. Soc., Providence, Rhode Island, 2002.
  • [12] D. Gilbarg and N. Trudinger, Elliptic Partial Differential Equations of Second Order, Springer, Berlin and New York, 1977.
  • [13] P. Goddard and D. I. Olive, Magnetic monopoles in gauge field theories, Rep. Prog. Phys. 41 (1978) 1357–1437.
  • [14] J. Greensite, An Introduction to the Confinement Problem, Lecture Notes in Physics 821, Springer-Verlag, Berlin and New York, 2011.
  • [15] S. P. Hastings, J. B. McLeod, and W. C. Troy, Static spherically symmetric solutions of a Yang–Mills field coupled to a dilaton, Proc. Roy. Soc. (London) A 449 (1995) 479–491.
  • [16] A. Jaffe and C. H. Taubes, Vortices and Monopoles, Birkhäuser, Boston, 1980.
  • [17] B. Julia and A. Zee, Poles with both magnetic and electric charges in non-Abelian gauge theory, Phys. Rev. D 11 (1975) 2227–2232.
  • [18] O. A. Ladyzhenskaya and N. N. Uraltseva, Linear and Quasilinear Elliptic Equations, Academic Press, New York and London, 1968.
  • [19] S. Mandelstam, Vortices and quark confinement in non-Abelian gauge theories, Phys. Lett. B 53 (1975) 476–478.
  • [20] S. Mandelstam, General introduction to confinement, Phys. Rep. C 67 (1980) 109–121.
  • [21] K. McLeod, W. C. Troy, and F. B. Weissler, Radial solutions of Δu+f(u)=0\Delta u+f(u)=0 with prescribed numbers of zeros, J. Diff. Eqs. 83 (1990) 368–378.
  • [22] C. W. Misner, K. S. Thorne, and J. A. Wheeler, Gravitation, Freeman, New York, 1973.
  • [23] Y. Nambu, Strings, monopoles, and gauge fields, Phys. Rev. D 10 (1974) 4262–4268.
  • [24] A. M. Polyakov, Particle spectrum in the quantum field theory, JETP Lett. 20 (1974) 194–195.
  • [25] J. Preskill, Magnetic monopoles, Annu. Rev. Nucl. Part. Sci. 34 (1984) 461–530.
  • [26] R. Rajaraman, Solitons and Instantons, North Holland, Amsterdam, 1982.
  • [27] L. H. Ryder, Quantum Field Theory, 2nd ed., Cambridge U. Press, Cambridge, U. K., 1996.
  • [28] J. Schwinger, A magnetic model of matter, Science 165 (1969) 757–761.
  • [29] M. Shifman and A. Yung, Supersymmetric solitons and how they help us understand non-Abelian gauge theories, Rev. Mod. Phys. 79 (2007) 1139–1196.
  • [30] M. Shifman and A. Yung, Supersymmetric Solitons, Cambridge U. Press, Cambridge, U. K., 2009.
  • [31] J. Smoller, A. Wasserman, S. T. Yau, and J. B. McLeod, Smooth static solutions of the Einstein–Yang/Mills equations, Bull. Amer. Math. Soc. (N.S.) 27 (1992) 239–242; Commun. Math. Phys. 143 (1991) 115–147.
  • [32] G. ’t Hooft, Magnetic monopoles in unified gauge theories, Nucl. Phys. B 79 (1974) 276–284.
  • [33] G. ’t Hooft, On the phase transition towards permanent quark confinement, Nucl. Phys. B 138 (1978) 1–25.
  • [34] G. ’t Hooft, Topology of the gauge condition and new confinement phases in non-Abelian gauge theories, Nucl. Phys. B 190 (1981) 455–478.
  • [35] Yu. S. Tyupkin, V. A. Fateev, and A. S. Shvarts, Particle-like solutions of the equations of gauge theories, Theoret. Math. Phys. 26 (1976) 270–273.
  • [36] T. Vachaspati, Kinks and Domain Walls: An Introduction to Classical and Quantum Solitons, Cambridge U. Press, 2006, Cambridge, U. K.
  • [37] R. M. Wald, General Relativity, U. Chicago Press, Chicago and London, 1984.
  • [38] E. J. Weinberg, Classical Solutions in Quantum Field Theory: Solitons and Instantons in High Energy Physics, Cambridge U. Press, Cambridge, U. K., 2012.
  • [39] Y. Yang, Solitons in Field Theory and Nonlinear Analysis, Springer-Verlag, Berlin and New York, 2001.
  • [40] X. Zhang and Y. Yang, Solutions to the minimization problem arising in a dark monopole model in gauge field theory, Nucl. Phys. B 951 (2020) 114851.