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

Unexciting non-Abelian electric fields

Tanmay Vachaspati Physics Department, Arizona State University, Tempe, Arizona 85287, USA.
Abstract

Electric fields in QED are known to discharge due to Schwinger pair production of charged particles. Corresponding electric fields in non-Abelian theory are known to discharge due to the production of gluons. Yet electric flux tubes in QCD ought to be stable to the production of charged gluons as they confine quarks. We resolve this conundrum by finding electric field configurations in pure non-Abelian gauge theory in which the Schwinger process is absent and the electric field is protected against quantum dissipation. We comment on the implications for QCD flux tubes.

I Introduction

Quantum particle production in time-dependent backgrounds continues to be a topic of great interest. The situation arises in the context of gravitational collapse and leads to Hawking radiation Hawking (1975), in cosmology where particle creation occurs due to the expansion of the universe Birrell and Davies (1984), and in Schwinger pair production Schwinger (1951) when the electric field is described in terms of a time-dependent gauge field. The production of particles implies that there is backreaction on the background, and external agencies must maintain the background or else it will dissipate. For example, black holes evaporate and capacitors discharge.

In a recent paper Vachaspati (2022) we discussed time-dependent backgrounds that are “unexciting”, i.e. time-dependent backgrounds in which there is no net production of particles. (This is connected to “shortcuts to adiabaticity (STA)” in quantum mechanical systems reviewed in Guéry-Odelin et al. (2019), and also related to certain gravitational systems discussed in Parikh et al. (2012); Parikh and Samantray (2013).) In most such backgrounds, particles are produced and then later absorbed so that the net particle production vanishes. A subset of unexciting backgrounds are those for which the particle production vanishes at all times. Such backgrounds are of interest because they are protected against quantum dissipation and no external agency is required to maintain the background. Their time-dependence is of a stationary nature. An example is that of a boosted soliton that is coupled to other quantum degrees of freedom: a boosted soliton is time-dependent but does not radiate particles.

Here we are interested in electric field backgrounds in pure non-Abelian gauge theory. Generally we would expect such electric field backgrounds to discharge due to the Schwinger pair production of gluon excitations. However, confinement suggests that electric flux tube configurations should be protected against quantum dissipation. By carefully choosing the time-dependence of the electric field it is possible to suppress particle production in a given excitation mode Kim and Schubert (2011); Vachaspati (2022), yet it is unclear what, if anything, could prevent Schwinger pair production completely. Even an exponentially suppressed pair production rate would eventually cause the electric flux tube to dissipate.

To clarify this motivation further, consider Schwinger pair production in the case of electrodynamics with a uniform electric field of strength EE and when the charge carriers have mass mem_{e} and charge ee. The rate of particle production goes as Schwinger (1951),

n˙e2E2exp(πme2/eE){\dot{n}}\propto e^{2}E^{2}\exp(-\pi m_{e}^{2}/eE) (1)

and can be understood in different ways depending on the choice of gauge.

If one adopts Coulomb gauge, the gauge potential for a uniform electric field along the zz-direction is

Aμ=(Ez,0,0,0)A^{\mu}=(-Ez,0,0,0) (2)

Then the gauge potential diverges asymptotically in the zz direction. As discussed in Ref. Itzykson and Zuber (1980); Kim and Page (2002, 2006) for example, particle production can be viewed as a tunneling process in a potential which is infinitely negative as zz\to\infty (see Fig. 1) so that the produced particles escape to infinity. For an electric field in a finite but large domain, the potential does not diverge but goes to a constant as zz\to\infty and the particles still escape to infinity.

Refer to caption
Figure 1: A sketch for the tunneling picture of Schwinger pair production. The Coulomb curves (small-dashed) show the Coulomb attractive potential between the pair of charges; the long dashed line shows the potential energy of the positive charge due to the external electric field, and the solid curves are the sum of the Coulomb and external electric potentials. Quantum fluctuations deep in the potential well with energy E can tunnel out and escape to infinity and would be interpreted as Schwinger particle pairs.

Alternatively, if one adopts temporal gauge, as we shall do, the gauge potential is

Aμ=(0,0,0,Et)A^{\mu}=(0,0,0,Et) (3)

Now the gauge field is spatially well-behaved but varies with time. Any quantum excitations of charged fields will obtain time-dependent frequencies, just as for a simple harmonic oscillator with a time-varying spring constant. The charged quantum modes in the vacuum will get excited due to this time-dependent background leading to pair production. The rate of particle production can be calculated using the standard machinery of Bogolyubov coefficients (e.g. Bogolyubov (1958); Birrell and Davies (1984)), or in the framework of the “classical-quantum correspondence” where quantum particle production is described in terms of solutions of the classical equations Vachaspati and Zahariade (2018, 2019).

Here we consider a pure non-Abelian SU(2) gauge theory with a background (“color”) electric field in temporal gauge. In addition, the theory contains “gluon” excitations that are massless and charged. A background electric field that is analogous to that in ordinary electrodynamics is known to pair produce gluons Matinyan and Savvidy (1978); Brown and Weisberger (1979); Yildiz and Cox (1980); Ambjorn and Hughes (1982a, b); Nayak and van Nieuwenhuizen (2005); Cooper and Nayak (2006); Cooper et al. (2008); Nair and Yelnikov (2010); Kim and Schubert (2011); Ilderton (2022); Huet et al. (2014); Ragsdale and Singleton (2017); Karabali et al. (2019); Cardona and Vachaspati (2021). One important difference from the original Schwinger calculation is that the gluons are massless and the exponential suppression in (1) is absent. In fact, there are ultraviolet and infrared divergences as discussed in Ref. Cardona and Vachaspati (2021) that are presumably controlled by asymptotic freedom and confinement. However, it appears that no matter how weak the electric field strength is, there is always some particle production and hence the electric field should decay.

If any non-Abelian electric field decays due to the Schwinger process, it would imply that any external electric charge would get shielded by gluons and the resulting long range electric field would vanish. This runs counter to the picture that QCD has electric flux tubes that confine electric charges, and we are led to the question if there can be non-Abelian electric field configurations that are immune to the Schwinger process, i.e. non-Abelian electric fields that are unexciting. Such electric flux tubes would be models for the QCD string responsible for confinement that have been discussed now for nearly half a century Kogut and Susskind (1974, 1975); Takahashi et al. (2002); Bissey et al. (2007).

A guess for an unexciting non-Abelian electric field configuration was suggested in Ref. Vachaspati (2022). One needs the electric field background to be stationary. Already we have mentioned boosted solitons as unexciting backgrounds. An alternative is to have “rotating” backgrounds. In non-Abelian gauge theories, for example when quantizing magnetic monopole backgrounds, it is known that there are rotor degrees of freedom that, when excited, endow a monopole with electric charge and convert it into a dyon. Could such rotor degrees of freedom be relevant for unexciting non-Abelian electric fields?

Approaching the problem from a different point of view, one wishes to construct “stationary” gauge fields that lead to a uniform electric field. Fortunately this problem has been analyzed in detail in Ref. Brown and Weisberger (1979) and it is found that there are two gauge inequivalent classes of gauge fields that lead to the same non-Abelian electric field. One of these ways is analogous to the Abelian gauge potential, while the second one is necessarily due to the non-Abelian nature of the model. We will explain this in more detail in Sec. III but suffice it to say that this second description of the electric field corresponds precisely to the uniform rotation of a rotor degree of freedom with quantized angular momentum (Sec. V). The analysis of Sec. IV shows explicitly that this gauge background is stationary and does not lead to particle production, and consequently is protected against quantum dissipation. Quantum excitations on top of the classical background will settle into some ground state which is very difficult to determine because of the strongly coupled nature of the system but, whatever the state may be, it will be stationary. In Sec. VII we discuss the simpler quantization of the homogeneous modes in the linearized approximation. Even this limited analysis has some novel features. Most of our analysis is done for a uniform electric field as this is simpler but in Sec. VI we remark on strategies to determine the profile of a flux tube. We start our discussion with a motivating illustration of an unexciting electric field in 1+1D in Sec. II and summarize our conclusions in Sec. VIII.

II An illustrative example in 1+1D

An example of an electric field configuration without Schwinger pair production is already known in massless QED in 1+1 dimensions Chu and Vachaspati (2010); Gold et al. (2021) with action,

S=d2x[ψ¯γμ(iμ+eAμ)ψ14FμνFμν]S=\int d^{2}x\left[{\bar{\psi}}\gamma^{\mu}(i\partial_{\mu}+eA_{\mu})\psi-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}\right] (4)

where ψ\psi is a fermion field, AμA_{\mu} is a U(1) gauge field, and FμνF_{\mu\nu} is the field strength.

An unexciting electric field background is given by Chu and Vachaspati (2010),

F01\displaystyle F_{01} =\displaystyle= Q(Θ(x+L/2)Θ(xL/2))\displaystyle Q(\Theta(x+L/2)-\Theta(x-L/2)) (5)
+g(f(x+L/2)f(xL/2))\displaystyle+g(f(x+L/2)-f(x-L/2))

where g=e/πg=e/\sqrt{\pi} and QQ is the external charge on a capacitor with plate separation LL, and

f(x)=Q2gsgn(x)(1eg|x|).f(x)=-\frac{Q}{2g}{\rm sgn}(x)\left(1-e^{-g|x|}\right). (6)

The terms proportional to QQ in (5) give the electric field of the classical capacitor, while the last two terms give the contribution of a quantum condensate of fermions. Quantum effects provide extra sources that screen some of the classical electric field, resulting in a net electric field in which there is no Schwinger pair production.

Similarly in the non-Abelian case discussed below, we consider an electric field configuration that solves the classical equations of motion only in the presence of some sources (see Sec. IV). These sources can be external or be generated internally by quantum effects due to higher order interactions.

III Uniform electric field

As discussed in Brown and Weisberger (1979), a homogeneous non-Abelian electric field can be derived from several gauge inequivalent potentials. Say we want the gauge potentials for an electric field in the third isospin direction and pointing along the zz direction, i.e. EEz3=W0z3E\equiv E^{3}_{z}=W^{3}_{0z} where WμaW^{a}_{\mu} are gauge potentials from which the field strength WμνaW^{a}_{\mu\nu} is derived in the usual way,

Wμνa=μWνaνWμa+ϵabcWμbWνcW^{a}_{\mu\nu}=\partial_{\mu}W^{a}_{\nu}-\partial_{\nu}W^{a}_{\mu}+\epsilon^{abc}W^{b}_{\mu}W^{c}_{\nu} (7)

where we have set the gauge coupling to unity since we will only be considering non-interacting quantum fluctuations on an electric field background.

The first “trivial” way to obtain Ez3E^{3}_{z} is to take,

Wμa=Etδa3μzW^{a}_{\mu}=-E\,t\delta^{a3}\partial_{\mu}z (8)

Then

Eia(tWiaiWta+ϵabcWtbWic)=Eδa3δizE^{a}_{i}\equiv-(\partial_{t}W^{a}_{i}-\partial_{i}W^{a}_{t}+\epsilon^{abc}W^{b}_{t}W^{c}_{i})=E\delta^{a3}\delta_{iz} (9)

A second way to obtain Ez3E^{3}_{z} is to take Brown and Weisberger (1979),

Wμ1=(Ω,0,0,0),Wμ2=(0,0,0,E/Ω),Wμ3=0W^{1}_{\mu}=(\Omega,0,0,0),\ \ W^{2}_{\mu}=(0,0,0,-E/\Omega),\ \ W^{3}_{\mu}=0 (10)

where Ω\Omega is a constant (that will turn out to be quantized in Sec. V.) With (10) we also obtain

Eia=Eδa3δiz.E^{a}_{i}=E\delta^{a3}\delta_{iz}. (11)

Even though the gauge potentials in Eqs. (8) and (10) yield the same field strength, they are gauge inequivalent, as are the gauge fields for different values of Ω\Omega, as can be seen by computing other gauge invariant quantities such as Brown and Weisberger (1979)111We are using the signature (+,,,)(+,-,-,-).,

(DνWμν)a(DσWμσ)a=E2(Ω2E2Ω2)(D_{\nu}W^{\mu\nu})^{a}(D_{\sigma}W^{\ \sigma}_{\mu})^{a}=-E^{2}\left(\Omega^{2}-\frac{E^{2}}{\Omega^{2}}\right) (12)

where

(DνWμν)aνWμνa+ϵabcWνbWμνc.(D_{\nu}W^{\mu\nu})^{a}\equiv\partial_{\nu}W^{\mu\nu a}+\epsilon^{abc}W_{\nu}^{b}W^{\mu\nu c}. (13)

Since the gauge invariant quantity on the left-hand side of (12) depends on Ω\Omega, gauge fields for different Ω\Omega values are gauge inequivalent. However, the energy density of the configuration is independent of Ω\Omega, since the electric field does not depend on Ω\Omega.

In the context of “unexciting” backgrounds discussed in Vachaspati (2022), we would like to work in temporal gauge and set W0a=0W^{a}_{0}=0. A gauge transformation yields,

𝐖μ=U1𝐖μU+iU1μU{\bf W}^{\prime}_{\mu}=U^{-1}{\bf W}_{\mu}U+iU^{-1}\partial_{\mu}U (14)

where now 𝐖μWμaTa=Wμaσa/2{\bf W}_{\mu}\equiv W^{a}_{\mu}T^{a}=W^{a}_{\mu}\sigma^{a}/2, TaT^{a} are the generators of SU(2) normalized to Tr(TaTb)=δab/2{\rm Tr}(T^{a}T^{b})=\delta^{ab}/2 and σa\sigma^{a} are the Pauli spin matrices. The gauge transformation UU that takes us to temporal gauge is given by,

U=eiσ1Ωt/2U=e^{i\sigma^{1}\Omega t/2} (15)

Using the identities,

ein^σθ\displaystyle e^{i\hat{n}\cdot\sigma\theta} =\displaystyle= cosθ+in^σsinθ\displaystyle\cos\theta+i\hat{n}\cdot\sigma\sin\theta
ein^σθσein^σθ\displaystyle e^{i\hat{n}\cdot{\vec{\sigma}}\theta}{\vec{\sigma}}e^{-i\hat{n}\cdot{\vec{\sigma}}\theta} =\displaystyle= σcos(2θ)+n^×σsin(2θ)\displaystyle{\vec{\sigma}}\cos(2\theta)+{\hat{n}}\times{\vec{\sigma}}\sin(2\theta) (16)
+n^n^σ(1cos(2θ))\displaystyle+{\hat{n}}\,{\hat{n}}\cdot{\vec{\sigma}}(1-\cos(2\theta))

we find

Wμ1\displaystyle W^{1}_{\mu}{}^{\prime} =\displaystyle= 0\displaystyle 0
Wμ2\displaystyle W^{2}_{\mu}{}^{\prime} =\displaystyle= EΩcos(Ωt)μz,\displaystyle-\frac{E}{\Omega}\cos(\Omega t)\partial_{\mu}z,
Wμ3\displaystyle W^{3}_{\mu}{}^{\prime} =\displaystyle= EΩsin(Ωt)μz.\displaystyle-\frac{E}{\Omega}\sin(\Omega t)\partial_{\mu}z. (17)

A global SU(2) rotation can be used to bring the gauge fields to a form that we will use

Wμ1\displaystyle W^{1}_{\mu} =\displaystyle= EΩcos(Ωt)μz,\displaystyle-\frac{E}{\Omega}\cos(\Omega t)\partial_{\mu}z,
Wμ2\displaystyle W^{2}_{\mu} =\displaystyle= EΩsin(Ωt)μz,\displaystyle-\frac{E}{\Omega}\sin(\Omega t)\partial_{\mu}z,
Wμ3\displaystyle W^{3}_{\mu} =\displaystyle= 0\displaystyle 0 (18)

and we have dropped the primes for convenience.

Ref. Brown and Weisberger (1979) wrote the gauge field in the form of (10) to show that the same electric field can be obtained with a one parameter (Ω\Omega) family of gauge inequivalent gauge fields. We are only interested in a fixed value of Ω\Omega and it is more convenient to write the non-vanishing components as

Wμ1=ϵcos(Ωt)μz,Wμ2=ϵsin(Ωt)μz.W^{1}_{\mu}=-\epsilon\cos(\Omega t)\partial_{\mu}z,\ W^{2}_{\mu}=-\epsilon\sin(\Omega t)\partial_{\mu}z. (19)

Then the electric field is given by

E=Ωϵ.E=\Omega\epsilon. (20)

Equivalently, we can work with Wμ±Wμ1±iWμ2W^{\pm}_{\mu}\equiv W^{1}_{\mu}\pm iW^{2}_{\mu},

Wμ±=ϵe±iΩtμz,Wμ3=0.W^{\pm}_{\mu}=-\epsilon e^{\pm i\Omega t}\partial_{\mu}z,\ \ W^{3}_{\mu}=0. (21)

Defining

Wμν±=Wμν1±iWμν2W_{\mu\nu}^{\pm}=W_{\mu\nu}^{1}\pm iW_{\mu\nu}^{2} (22)

we find

Wμν±=μWν±νWμ±±i(Wμ3Wν±Wν3Wμ±)W_{\mu\nu}^{\pm}=\partial_{\mu}W_{\nu}^{\pm}-\partial_{\nu}W_{\mu}^{\pm}\pm i(W_{\mu}^{3}W_{\nu}^{\pm}-W_{\nu}^{3}W_{\mu}^{\pm}) (23)

and

Wμν3=μWν3νWμ3+i2(Wμ+WνWν+Wμ).W_{\mu\nu}^{3}=\partial_{\mu}W_{\nu}^{3}-\partial_{\nu}W_{\mu}^{3}+\frac{i}{2}(W_{\mu}^{+}W_{\nu}^{-}-W_{\nu}^{+}W_{\mu}^{-}). (24)

For the background in (18) we get,

Wμν±=iE(μtνzνtμz)e±iΩt,Wμν3=0W_{\mu\nu}^{\pm}=\mp iE(\partial_{\mu}t\,\partial_{\nu}z-\partial_{\nu}t\,\partial_{\mu}z)e^{\pm i\Omega t},\ \ W_{\mu\nu}^{3}=0 (25)

The Lagrangian density for the non-Abelian gauge sector is

g=14[Wμν+Wμν+Wμν3Wμν3]{\cal L}_{g}=-\frac{1}{4}\left[W_{\mu\nu}^{+}W^{\mu\nu-}+W_{\mu\nu}^{3}W^{\mu\nu 3}\right] (26)

The full model will necessarily include a Lagrangian density for external sources and their couplings to the gauge fields as we will discuss in Sec. IV.

The important lesson of this section is that in non-Abelian gauge theories we have several gauge inequivalent choices for the gauge potential corresponding to an electric field background. The gauge potential of interest to us is time-dependent only because it is rotating in gauge field space as in (21).

IV Expansion around a flux tube background

We now consider the electric field configuration in (18) as a background and denote it by AμA_{\mu}. We assume it is produced by some unspecified external sources, and we wish to determine if it leads to Schwinger pair creation. We write

Wμ±=Aμ±+e±iΩtQμ±,Wμ3=Aμ3+Qμ3W_{\mu}^{\pm}=A_{\mu}^{\pm}+e^{\pm i\Omega t}Q_{\mu}^{\pm},\ \ W_{\mu}^{3}=A_{\mu}^{3}+Q_{\mu}^{3} (27)

where now AμaA_{\mu}^{a} includes an unspecified radial profile function, f(r)f(r), (r=x2+y2r=\sqrt{x^{2}+y^{2}}),

Aμ±=ϵe±iΩtf(r)μz,Aμ3=0,A^{\pm}_{\mu}=-\epsilon e^{\pm i\Omega t}f(r)\,\partial_{\mu}z,\ \ A^{3}_{\mu}=0, (28)

and QμaQ_{\mu}^{a} are quantum excitations around the background. More explicitly,

Wμ±=e±iΩt(ϵf(r)μz+Qμ±),Wμ3=Qμ3W_{\mu}^{\pm}=e^{\pm i\Omega t}\left(-\epsilon f(r)\,\partial_{\mu}z+Q_{\mu}^{\pm}\right),\ \ W_{\mu}^{3}=Q_{\mu}^{3} (29)

Then

Wμν±\displaystyle W_{\mu\nu}^{\pm} =\displaystyle= e±iΩt[𝒜μν±+(𝒟μQν±𝒟νQμ±)\displaystyle e^{\pm i\Omega t}\bigl{[}{\mathscr{A}}_{\mu\nu}^{\pm}+({\mathscr{D}}_{\mu}Q_{\nu}^{\pm}-{\mathscr{D}}_{\nu}Q_{\mu}^{\pm}) (30)
±i(Qμ3Qν±Qν3Qμ±)]\displaystyle\hskip 56.9055pt\pm i(Q_{\mu}^{3}Q_{\nu}^{\pm}-Q_{\nu}^{3}Q_{\mu}^{\pm})\bigr{]}

where

𝒜μν±\displaystyle{\mathscr{A}}_{\mu\nu}^{\pm} \displaystyle\equiv ±iEf(r)(μzνtνzμt)\displaystyle\pm iEf(r)(\partial_{\mu}z\partial_{\nu}t-\partial_{\nu}z\partial_{\mu}t) (31)
+ϵf(r)(μzνrνzμr),\displaystyle\hskip 42.67912pt+\epsilon f^{\prime}(r)(\partial_{\mu}z\partial_{\nu}r-\partial_{\nu}z\partial_{\mu}r),
𝒟μQν±μQν±±iΩμtQν±±iϵf(r)μzQν3.{\mathscr{D}}_{\mu}Q_{\nu}^{\pm}\equiv\partial_{\mu}Q_{\nu}^{\pm}\pm i\Omega\partial_{\mu}t\,Q_{\nu}^{\pm}\pm i\epsilon f(r)\partial_{\mu}z\,Q_{\nu}^{3}. (32)

We also have

Wμν3=𝒟μQν3𝒟νQμ3+i2(Qμ+QνQν+Qμ)W_{\mu\nu}^{3}={\mathscr{D}}_{\mu}Q_{\nu}^{3}-{\mathscr{D}}_{\nu}Q_{\mu}^{3}+\frac{i}{2}(Q_{\mu}^{+}Q_{\nu}^{-}-Q_{\nu}^{+}Q_{\mu}^{-}) (33)

with

𝒟μQν3\displaystyle{\mathscr{D}}_{\mu}Q_{\nu}^{3} \displaystyle\equiv μQν3+iϵ2f(r)μzQν+iϵ2f(r)μzQν\displaystyle\partial_{\mu}Q_{\nu}^{3}+i\frac{\epsilon}{2}f(r)\partial_{\mu}z\,Q_{\nu}^{+}-i\frac{\epsilon}{2}f(r)\partial_{\mu}z\,Q_{\nu}^{-} (34)
=\displaystyle= μQν3ϵf(r)μzQν2\displaystyle\partial_{\mu}Q_{\nu}^{3}-\epsilon f(r)\partial_{\mu}zQ_{\nu}^{2}

The Lagrangian density for QμaQ^{a}_{\mu} is evaluated from (26),

g\displaystyle{\cal L}_{g} =\displaystyle= 14|𝒜μν++(𝒟μQν+𝒟νQμ+)\displaystyle-\frac{1}{4}\biggl{|}{\mathscr{A}}_{\mu\nu}^{+}+({\mathscr{D}}_{\mu}Q_{\nu}^{+}-{\mathscr{D}}_{\nu}Q_{\mu}^{+}) (35)
+i(Qμ3Qν+Qν3Qμ+)|2\displaystyle\hskip 71.13188pt+i(Q_{\mu}^{3}Q_{\nu}^{+}-Q_{\nu}^{3}Q_{\mu}^{+})\biggr{|}^{2}
14|𝒟μQν3𝒟νQμ3+i2(Qμ+QνQν+Qμ)|2\displaystyle\hskip-14.22636pt-\frac{1}{4}\biggl{|}{\mathscr{D}}_{\mu}Q_{\nu}^{3}-{\mathscr{D}}_{\nu}Q_{\mu}^{3}+\frac{i}{2}(Q_{\mu}^{+}Q_{\nu}^{-}-Q_{\nu}^{+}Q_{\mu}^{-})\biggr{|}^{2}

The Lagrangian density (35) describes the interaction of QμaQ_{\mu}^{a} with the background AμA_{\mu}. Even at this stage it is clear that there can be no particle production: the background dependent terms in (35), for example 𝒜μν±{\mathscr{A}}_{\mu\nu}^{\pm} and in 𝒟μQνa{\mathscr{D}}_{\mu}Q_{\nu}^{a}, are all independent of time and there is no zz-dependence either. For Schwinger pair production, the gauge field background should either have non-trivial time-dependence or it should have non-trivial spatial dependence, as discussed in Sec. I. The interactions of QμaQ_{\mu}^{a} with the background will lead to a non-trivial ground state wavefunctional but without any time-dependent excitations. Hence the electric-magnetic field background in (31) is “unexciting”.

We will now examine the system more explicitly by expanding the Lagrangian density in powers of QQ.

IV.1 First order Lagrangian density

The Lagrangian density up to linear order terms in QQ is,

g(1)\displaystyle{\cal L}_{g}^{(1)} =\displaystyle= 14𝒜μν+(𝒟μQν+𝒟νQμ+)+c.c.\displaystyle-\frac{1}{4}\mathscr{A}_{\mu\nu}^{+}\bigl{(}{\mathscr{D}}^{\mu}Q^{\nu+}-{\mathscr{D}}^{\nu}Q^{\mu+}\bigr{)}^{*}+{\rm c.c.} (36)
\displaystyle\to +12(𝒟μ𝒜μν+)Qν+c.c.\displaystyle+\frac{1}{2}({\mathscr{D}}^{\mu}\mathscr{A}_{\mu\nu}^{+})Q^{\nu-}+{\rm c.c.}

where in the second line we have dropped total derivative terms. The linear order variation doesn’t vanish; neither do they for the illustrative example in (5) since there are external sources for the electric field and quantum effects induce a condensate of sources. Here we simply supplement this linear order Lagrangian density with a source term that couples external currents, jμaj_{\mu}^{a}, to WμaW_{\mu}^{a},

s=jμaWμa=jμaAμa+e±iΩtjμQμ±+jμ3Qμ3{\cal L}_{s}=j_{\mu}^{a}W^{\mu a}=j_{\mu}^{a}A^{\mu a}+e^{\pm i\Omega t}j_{\mu}^{\mp}Q^{\mu\pm}+j_{\mu}^{3}Q^{\mu 3} (37)

The external current includes sources that are necessary to generate the background net flux of electric field. The currents can also contain effective quantum contributions that arise due to higher order interactions as these backreact at the linear level (see Sec. II). Here we simply assume the existence of such a current without any dynamical explanation.

Requiring that the linear order terms vanish up to total derivatives gives,

jμ±\displaystyle j_{\mu}^{\pm} =\displaystyle= ϵ2e±iΩt(f′′+fr+Ω2f)μz,\displaystyle-\frac{\epsilon}{2}e^{\pm i\Omega t}\left(f^{\prime\prime}+\frac{f^{\prime}}{r}+\Omega^{2}f\right)\partial_{\mu}z,
jμ3\displaystyle j_{\mu}^{3} =\displaystyle= Ωϵ2f2μt.\displaystyle-\Omega\epsilon^{2}f^{2}\,\partial_{\mu}t. (38)

These currents do not include the external sources located at z=±z=\pm\infty. To see this it is most transparent to consider the Abelian case of a homogeneous electric field with field strength tensor

Fμν=E(μtνzνtμz).F_{\mu\nu}=E(\partial_{\mu}t\partial_{\nu}z-\partial_{\nu}t\partial_{\mu}z). (39)

Insertion into Maxwell’s equations gives

jμ=νFνμ=0j_{\mu}=-\partial^{\nu}F_{\nu\mu}=0 (40)

and the charged capacitor plates at z=±z=\pm\infty are not included in jμj_{\mu}.

Since the currents in (38) do not include the asymptotic sources, they must arise entirely as a quantum condensate similar to the second line in (5). To show that such a condensate arises due to the strong interactions is difficult but some progress may be made in the semiclassical approximation. We start with the equation of motion for the gauge fields,

DνWνμa=0D_{\nu}W^{\nu\mu a}=0 (41)

where the covariant derivative is defined in (13). We then substitute

Wμa=Aμa+qμaW_{\mu}^{a}=A_{\mu}^{a}+q_{\mu}^{a} (42)

in (41) where AμaA^{\mu a} is classical while qμaq^{\mu a} is quantum. The operator equation (41) is replaced by its ground state expectation value. If we assume that expectation values of odd powers of qμaq^{\mu a} vanish, some algebra yields,

Dν(A)Aνμa=jμaD_{\nu}^{(A)}A^{\nu\mu a}=-j^{\mu a} (43)

where Dν(A)AνμaD_{\nu}^{(A)}A^{\nu\mu a} is as in (13) with WW replaced by AA and

jμa\displaystyle j^{\mu a} \displaystyle\equiv ϵabcνqνbqμcqνbμqνc+2qνbνqμcR\displaystyle\epsilon^{abc}\langle\partial_{\nu}q^{\nu b}q^{\mu c}-q^{\nu b}\partial^{\mu}q_{\nu}^{c}+2q^{\nu b}\partial_{\nu}q^{\mu c}\rangle_{R} (44)
+Aνbqνaqμb2qνbqμaR+AνaqνbqμbR\displaystyle+A_{\nu}^{b}\langle q^{\nu a}q^{\mu b}-2q^{\nu b}q^{\mu a}\rangle_{R}+A_{\nu}^{a}\langle q^{\nu b}q^{\mu b}\rangle_{R}
+AμbqνbqνaRAμaqνbqνbR\displaystyle+A^{\mu b}\langle q^{b}_{\nu}q^{\nu a}\rangle_{R}-A^{\mu a}\langle q_{\nu}^{b}q^{\nu b}\rangle_{R}

The symbol R\langle\cdot\rangle_{R} refers to the renormalized vacuum expectation value. One possible renormalization scheme would be to subtract out the expectation value in the trivial vacuum with zero electric field,

𝒪R=𝒪𝒪ϵ=0\langle{\cal O}\rangle_{R}=\langle{\cal O}\rangle-\langle{\cal O}\rangle_{\epsilon=0} (45)

for any quantum operator 𝒪{\cal O}. To illustrate how the current might result in a charge condensate, consider the μ=0\mu=0, a=3a=3 component of (44). Since we are working in temporal gauge and qμaq_{\mu}^{a} is defined in terms of QμaQ_{\mu}^{a} using (27), the expression simplifies, and we get

j0(3)\displaystyle j^{(3)}_{0} =\displaystyle= Ω(Qi(1))2+(Qi(2))2R\displaystyle\Omega\langle(Q_{i}^{(1)})^{2}+(Q_{i}^{(2)})^{2}\rangle_{R} (46)
+Qi(1)Q˙i(2)Qi(2)Q˙i(1)R\displaystyle+\langle Q_{i}^{(1)}{\dot{Q}}_{i}^{(2)}-Q_{i}^{(2)}{\dot{Q}}_{i}^{(1)}\rangle_{R}

Assuming the ground state of QμaQ_{\mu}^{a} has zero “angular momentum” (also see Sec. VII), we get a charge density,

j0(3)=Ω(Qi(1))2+(Qi(2))2R.\displaystyle j^{(3)}_{0}=\Omega\langle(Q_{i}^{(1)})^{2}+(Q_{i}^{(2)})^{2}\rangle_{R}. (47)

This can be compared to (38) and for a homogeneous electric field (f=1f=1) we have,

ϵ2=(Qi(1))2+(Qi(2))2R.\epsilon^{2}=-\langle(Q_{i}^{(1)})^{2}+(Q_{i}^{(2)})^{2}\rangle_{R}. (48)

Note that the right-hand side can be positive due to the definition in (45).

In summary, the background need not satisfy the classical equations of motion; instead it is more realistic to ask if the background satisfies the semiclassical equations of motion. By examining the semiclassical equations, we discover that quantum effects can indeed provide appropriate sources for the electric field background under consideration.

IV.2 Second order Lagrangian density

The quadratic order Lagrangian density is,

g(2)\displaystyle{\cal L}_{g}^{(2)} =\displaystyle= 14|𝒟μQν+𝒟νQμ+|214|𝒟μQν3𝒟νQμ3|2\displaystyle-\frac{1}{4}\bigl{|}{\mathscr{D}}_{\mu}Q_{\nu}^{+}-{\mathscr{D}}_{\nu}Q_{\mu}^{+}\bigr{|}^{2}-\frac{1}{4}\bigl{|}{\mathscr{D}}_{\mu}Q_{\nu}^{3}-{\mathscr{D}}_{\nu}Q_{\mu}^{3}\bigr{|}^{2} (49)
+i4[𝒜μν+(Qμ3QνQν3Qμ)c.c.].\displaystyle+\frac{i}{4}\left[\mathscr{A}_{\mu\nu}^{+}(Q^{\mu 3}Q^{\nu-}-Q^{\nu 3}Q^{\mu-})-c.c.\right].

or, explicitly,

g(2)=\displaystyle\hskip-21.33955pt{\cal L}_{g}^{(2)}= (50)
12(Q˙i(1)ΩQi(2))2+12(Q˙i(2)+ΩQi(1))2+12(Q˙i(3))2\displaystyle\frac{1}{2}({\dot{Q}}^{(1)}_{i}-\Omega Q^{(2)}_{i})^{2}+\frac{1}{2}({\dot{Q}}^{(2)}_{i}+\Omega Q^{(1)}_{i})^{2}+\frac{1}{2}({\dot{Q}}^{(3)}_{i})^{2}
\displaystyle- 14(iQj(1)jQi(1))2\displaystyle\frac{1}{4}(\partial_{i}Q^{(1)}_{j}-\partial_{j}Q^{(1)}_{i})^{2}
\displaystyle- 14(iQj(2)jQi(2)+ϵf(z^iQj(3)z^jQi(3)))2\displaystyle\frac{1}{4}(\partial_{i}Q^{(2)}_{j}-\partial_{j}Q^{(2)}_{i}+\epsilon f({\hat{z}}_{i}Q^{(3)}_{j}-{\hat{z}}_{j}Q^{(3)}_{i}))^{2}
\displaystyle- 14(iQj(3)jQi(3)ϵf(z^iQj(2)z^jQi(2)))2\displaystyle\frac{1}{4}(\partial_{i}Q^{(3)}_{j}-\partial_{j}Q^{(3)}_{i}-\epsilon f({\hat{z}}_{i}Q^{(2)}_{j}-{\hat{z}}_{j}Q^{(2)}_{i}))^{2}
+\displaystyle+ ϵf(z^ir^jz^jr^i)Qi(2)Qj(3)\displaystyle\epsilon f^{\prime}({\hat{z}}_{i}{\hat{r}}_{j}-{\hat{z}}_{j}{\hat{r}}_{i})Q^{(2)}_{i}Q^{(3)}_{j}

where Qi±=Qi(1)±iQi(2)Q^{\pm}_{i}=Q^{(1)}_{i}\pm iQ^{(2)}_{i}, z^i{\hat{z}}_{i}, r^i{\hat{r}}_{i} are unit vectors in the zz- and rr- directions, and the contraction of spatial indices is with the Kronecker delta, e.g. (Q˙i(3))2=Q˙i(3)Q˙i(3)({\dot{Q}}^{(3)}_{i})^{2}={\dot{Q}}^{(3)}_{i}{\dot{Q}}^{(3)}_{i}.

Denoting the momentum conjugate to QiaQ^{a}_{i} by PiaP^{a}_{i}, the Hamiltonian density corresponding to g(2){\cal L}_{g}^{(2)} is,

g(2)=12(Pi(a))2+14(iQj(1)jQi(1))2\displaystyle\hskip-21.33955pt{\cal H}_{g}^{(2)}=\frac{1}{2}(P^{(a)}_{i})^{2}+\frac{1}{4}(\partial_{i}Q^{(1)}_{j}-\partial_{j}Q^{(1)}_{i})^{2} (51)
+\displaystyle+ 14(iQj(2)jQi(2)+ϵf(z^iQj(3)z^jQi(3)))2\displaystyle\frac{1}{4}(\partial_{i}Q^{(2)}_{j}-\partial_{j}Q^{(2)}_{i}+\epsilon f({\hat{z}}_{i}Q^{(3)}_{j}-{\hat{z}}_{j}Q^{(3)}_{i}))^{2}
+\displaystyle+ 14(iQj(3)jQi(3)ϵf(z^iQj(2)z^jQi(2)))2\displaystyle\frac{1}{4}(\partial_{i}Q^{(3)}_{j}-\partial_{j}Q^{(3)}_{i}-\epsilon f({\hat{z}}_{i}Q^{(2)}_{j}-{\hat{z}}_{j}Q^{(2)}_{i}))^{2}
\displaystyle- ϵf(z^ir^jz^jr^i)Qi(2)Qj(3)+𝒥\displaystyle\epsilon f^{\prime}({\hat{z}}_{i}{\hat{r}}_{j}-{\hat{z}}_{j}{\hat{r}}_{i})Q^{(2)}_{i}Q^{(3)}_{j}+{\cal J}

where,

𝒥=Ω(Qi(2)Pi(1)Qi(1)Pi(2)){\cal J}=\Omega(Q^{(2)}_{i}P^{(1)}_{i}-Q^{(1)}_{i}P^{(2)}_{i}) (52)

is an “angular momentum” term.

To check that the background is unexciting, we simply note that the Hamiltonian density does not have any time-dependence. If we decompose the field into eigenmodes, the amplitude of each eigenmode is a quadratic variable that behaves like a simple harmonic oscillator variable with possibly an angular momentum contribution to the Hamiltonian (see Sec. VII).

One potential complication is if the diagonalization of the second order Hamiltonian density, g(2){\cal H}_{g}^{(2)}, leads to modes that have imaginary frequencies, i.e. are simple harmonic oscillators with inverted potentials. An analysis of the spectrum of frequencies was carried out in Ref. Bazak and Mrowczynski (2022) for the special choice Ω=E\Omega=\sqrt{E} and the authors found some unstable modes. These unstable modes imply that the quantum ground state of the QiaQ^{a}_{i} will be something other than the simple harmonic oscillator ground states (at least for this choice of Ω\Omega). The instability will be tamed by the higher order interactions. Since the higher order Lagrangian density is also time-independent (see Subsection. IV.3 below), there can be no particle production and the electric field background is protected against quantum dissipation. However the danger is that the expectation value of the quantum excitations might not vanish and then the separation between the classical background, AμaA_{\mu}^{a}, and the quantum excitations becomes unclear. The idea behind postulating AμaA_{\mu}^{a} as a background is that the quantum excitations are small and their expectation values vanish. For this to happen, the background should not have any instabilities. The stability analysis deserves to be investigated for the entire range of parameters.

To evaluate the quantum state of the QiaQ^{a}_{i} in the full interacting theory is beyond the scope of the present work, but we discuss the simpler problem of the quantum state for homogeneous modes in the quadratic Hamiltonian in Sec. VII.

IV.3 Higher order Lagrangian density

For completeness we also give the cubic and quartic order terms in the Lagrangian density,

g(3+4)\displaystyle{\cal L}_{g}^{(3+4)} =\displaystyle= i4[(𝒟iQj+𝒟jQi+)(Qi3QjQj3Qi)\displaystyle\frac{i}{4}\biggl{[}({\mathscr{D}}_{i}Q_{j}^{+}-{\mathscr{D}}_{j}Q_{i}^{+})(Q_{i}^{3}Q_{j}^{-}-Q_{j}^{3}Q_{i}^{-}) (53)
+12(𝒟iQj3𝒟jQi3)(Qi+QjQj+Qi)c.c.]\displaystyle\hskip-28.45274pt+\frac{1}{2}({\mathscr{D}}_{i}Q_{j}^{3}-{\mathscr{D}}_{j}Q_{i}^{3})(Q_{i}^{+}Q_{j}^{-}-Q_{j}^{+}Q_{i}^{-})-c.c.\biggr{]}
14|Qi3Qj+Qj3Qi+|2116|Qi+QjQj+Qi|2\displaystyle\hskip-49.79231pt-\frac{1}{4}\bigl{|}Q_{i}^{3}Q_{j}^{+}-Q_{j}^{3}Q_{i}^{+}\bigr{|}^{2}-\frac{1}{16}\bigl{|}Q_{i}^{+}Q_{j}^{-}-Q_{j}^{+}Q_{i}^{-}\bigr{|}^{2}

To summarize Sec. IV, the electric field background (28) is unexciting. There will be quantum fluctuations around this background and they will be in some stationary quantum ground state but particle production will be absent. For this reason, the electric field background is stable to quantum dissipation.

V Symmetry and quantization

As discussed in the introduction, a static soliton can be boosted to give a time-dependent background but this will not lead to particle production. This is because the static soliton has translational symmetry and a boosted soliton provides time dependence to the background but the excitation frequencies are still time-independent. In other words, the boosted soliton is a “stationary” background. Similarly, there should be a symmetry of the non-Abelian gauge theory, and the electric field background we have been discussing should be due to a time-dependence in the symmetry variables. The situation is very similar to the symmetries of monopoles and the rotor degree of freedom that leads to dyons, discussed for example in Refs. Julia and Zee (1975); Callan (1982) and extensively in Section 2.7 of Preskill (1987).

Consider the global gauge transformation

Wμ±Wμ=±e±iθWμ±,Wμ3Wμ=3Wμ3W_{\mu}^{\pm}\to W^{\prime}_{\mu}{}^{\pm}=e^{\pm i\theta}W_{\mu}^{\pm},\ \ W_{\mu}^{3}\to W^{\prime}_{\mu}{}^{3}=W_{\mu}^{3} (54)

where θ\theta is a constant parameter. The boosted soliton background corresponds in this case to promoting the rotor degree of freedom: θθ(t)\theta\to\theta(t).

Next let

Wμ±=e±iθvμ±,Wμ3=vμ3W_{\mu}^{\pm}=e^{\pm i\theta}v_{\mu}^{\pm},\ \ W_{\mu}^{3}=v_{\mu}^{3} (55)

where v0a=0v_{0}^{a}=0 in temporal gauge and we also assume that the vμav_{\mu}^{a} background is static (just like the static soliton). The angular variable θ\theta is assumed to only depend on time. Substituting (55) in the Lagrangian for WμaW_{\mu}^{a} (i.e. (26) integrated over space) gives the Lagrangian for θ\theta,

Lθ=12Iθ˙2L_{\theta}=\frac{1}{2}I{\dot{\theta}}^{2} (56)

where the “moment of inertia” is

I=d3x|vi+|2.I=\int d^{3}x|v_{i}^{+}|^{2}. (57)

Solving for the quantum dynamics of θ\theta is straightforward and leads to the quantization of angular momentum and correspondingly

Ω=θ˙=lI,l=0,±1,±2,\Omega={\dot{\theta}}=\frac{l}{I},\ \ l=0,\pm 1,\pm 2,\ldots (58)

Thus the parameter Ω\Omega is quantized. The electric field strength is also quantized as

E=Ωϵ=ϵIl.E=\Omega\epsilon=\frac{\epsilon}{I}l. (59)

The quantum number ll has the interpretation of the number of charge quanta. To see this we evaluate the total charge, qaq^{a}, by integrating the μ=0\mu=0 component of (38) over a volume VV for the homogeneous case (f=1f=1). This gives

qa=VΩϵ2δa3=lδa3q^{a}=-V\Omega\epsilon^{2}\delta^{a3}=-l\delta^{a3} (60)

in units where the gauge coupling constant g=1g=1 and we have used (58) and I=ϵ2VI=\epsilon^{2}V.

VI Flux tube profile

The quantum state of the fluctuations QiaQ^{a}_{i} will depend on the profile f(r)f(r). If we know the quantum state, we can calculate the expectation value of the Hamiltonian, H\langle H\rangle, for the choice of f(r)f(r). If H\langle H\rangle is minimized for some f(r)f(r), subject to a suitable constraint on the background electric field, it would specify the profile of the electric field flux tube and its tension. We can define the background electric flux as

Φ±d2x𝒜0z±=iEe±iΩtd2xf(r),\Phi^{\pm}\equiv\int d^{2}x\,{\mathscr{A}}_{0z}^{\pm}=\mp iEe^{\pm i\Omega t}\int d^{2}x\,f(r), (61)

with 𝒜μν±{\mathscr{A}}_{\mu\nu}^{\pm} given in (31) and Φ3=0\Phi^{3}=0. The flux itself is not gauge invariant but, having chosen a fixed form of the background gauge fields, one can restrict the class of functions in this background gauge. This amounts to exploring profile functions for which the integral on the right-hand side of (61) is unity,

2π𝑑rrf(r)=1.2\pi\int dr\,rf(r)=1. (62)

Unfortunately, there is no simple way to find the quantum state of QiaQ^{a}_{i} for given f(r)f(r), especially as the fluctuations are strongly interacting. The only hope seems to be to determine f(r)f(r) numerically, using lattice techniques.

VII Quantum state of the homogeneous modes

We now consider the Hamiltonian density in (51) for homogeneous excitations on a uniform electric field background. Then iQja=0\partial_{i}Q^{a}_{j}=0 and f=1f=1, and the Hamiltonian is related to the Hamiltonian density by suitable factors of the spatial volume VV,

Hg(2)=12V(Pi(a))2+ϵ2V4(z^iQj(3)z^jQi(3))2\displaystyle\hskip-21.33955ptH_{g}^{(2)}=\frac{1}{2V}(P^{(a)}_{i})^{2}+\frac{\epsilon^{2}V}{4}({\hat{z}}_{i}Q^{(3)}_{j}-{\hat{z}}_{j}Q^{(3)}_{i})^{2}
+ϵ2V4(z^iQj(2)z^jQi(2))2+Ω(Qi(2)Pi(1)Qi(1)Pi(2))\displaystyle\hskip-24.18501pt+\frac{\epsilon^{2}V}{4}({\hat{z}}_{i}Q^{(2)}_{j}-{\hat{z}}_{j}Q^{(2)}_{i})^{2}+\Omega(Q^{(2)}_{i}P^{(1)}_{i}-Q^{(1)}_{i}P^{(2)}_{i}) (63)

The Hamiltonian then can be written as a sum of six sub-Hamiltonians,

H1\displaystyle H_{1} =\displaystyle= 12V(Px(1))2+12V(Px(2))2+ϵ2V2(Qx(2))2\displaystyle\frac{1}{2V}(P^{(1)}_{x})^{2}+\frac{1}{2V}(P^{(2)}_{x})^{2}+\frac{\epsilon^{2}V}{2}(Q^{(2)}_{x})^{2} (64)
+Ω(Qx(2)Px(1)Qx(1)Px(2))\displaystyle\hskip 28.45274pt+\Omega(Q^{(2)}_{x}P^{(1)}_{x}-Q^{(1)}_{x}P^{(2)}_{x})
H2\displaystyle H_{2} =\displaystyle= 12V(Px(3))2+ϵ2V2(Qx(3))2\displaystyle\frac{1}{2V}(P^{(3)}_{x})^{2}+\frac{\epsilon^{2}V}{2}(Q^{(3)}_{x})^{2} (65)
H3\displaystyle H_{3} =\displaystyle= 12V(Py(1))2+12V(Py(2))2+ϵ2V2(Qy(2))2\displaystyle\frac{1}{2V}(P^{(1)}_{y})^{2}+\frac{1}{2V}(P^{(2)}_{y})^{2}+\frac{\epsilon^{2}V}{2}(Q^{(2)}_{y})^{2} (66)
+Ω(Qy(2)Py(1)Qy(1)Py(2))\displaystyle\hskip 28.45274pt+\Omega(Q^{(2)}_{y}P^{(1)}_{y}-Q^{(1)}_{y}P^{(2)}_{y})
H4\displaystyle H_{4} =\displaystyle= 12V(Py(3))2+ϵ2V2(Qy(3))2\displaystyle\frac{1}{2V}(P^{(3)}_{y})^{2}+\frac{\epsilon^{2}V}{2}(Q^{(3)}_{y})^{2} (67)
H5\displaystyle H_{5} =\displaystyle= 12V(Pz(1))2+12V(Pz(2))2\displaystyle\frac{1}{2V}(P^{(1)}_{z})^{2}+\frac{1}{2V}(P^{(2)}_{z})^{2} (68)
+Ω(Qz(2)Pz(1)Qz(1)Pz(2))\displaystyle\hskip 28.45274pt+\Omega(Q^{(2)}_{z}P^{(1)}_{z}-Q^{(1)}_{z}P^{(2)}_{z})
H6\displaystyle H_{6} =\displaystyle= 12V(Pz(3))2\displaystyle\frac{1}{2V}(P^{(3)}_{z})^{2} (69)

The Hamiltonians H2H_{2} and H4H_{4} correspond to simple harmonic oscillators, while H6H_{6} is that of a free particle. Their eigenstates are well-known; the simple harmonic oscillators states are all bound, while the free particle only has continuum states. H5H_{5} is the Hamiltonian for a free particle in two dimensions but with an additional angular momentum term, while H1H_{1} and H3H_{3} are identical in structure and can also be thought of as a particle in two dimensions with an angular momentum term plus an anisotropic potential. Thus we only need to consider H1H_{1} and H5H_{5}.

Let us consider H5H_{5} first and write it in a less cumbersome way,

H512πx2+12πy2+Ω(yπxxπy)H_{5}\to\frac{1}{2}\pi_{x}^{2}+\frac{1}{2}\pi_{y}^{2}+\Omega(y\pi_{x}-x\pi_{y}) (70)

where πx\pi_{x} and πy\pi_{y} are momenta conjugate to xx and yy, and we have also chosen units so that V=1V=1. One can check that

[yπxxπy,H5]=0[y\pi_{x}-x\pi_{y},H_{5}]=0 (71)

and there are simultaneous eigenstates of the angular momentum operator, yπxxπyy\pi_{x}-x\pi_{y}, and H5H_{5}. Defining polar coordinates in the usual way

x=rcosθ,y=rsinθx=r\,\cos\theta,\ \ y=r\,\sin\theta (72)

the energy eigenstates (up to an overall normalization factor) are given in terms of Bessel functions,

ψ(r,θ,t)=CeitJl(2(+Ωm)r)eimθ,\psi(r,\theta,t)=Ce^{-i{\cal E}t}J_{l}\left(\sqrt{2({\cal E}+\Omega m)}\,r\right)e^{im\theta}, (73)

with m=0,±1,±2,m=0,\pm 1,\pm 2,\ldots, CC is a normalization constant, and the energy, {\cal E}, is constrained by +Ωm0{\cal E}+\Omega m\geq 0.

An interesting point is that {\cal E} can be negative. In fact, for Ω>0\Omega>0 and mm\to\infty, {\cal E} can be arbitrarily negative. We expect that when the quartic interactions are taken into account, the energy may get bounded from below. For example, if the interactions effectively provide a mass to the homogeneous mode, we could consider a rotationally symmetric simple harmonic oscillator in 2D with an additional angular momentum term

He.g.=12πx2+12πy2+μ22(x2+y2)+Ω(yπxxπy),H_{\rm e.g.}=\frac{1}{2}\pi_{x}^{2}+\frac{1}{2}\pi_{y}^{2}+\frac{\mu^{2}}{2}(x^{2}+y^{2})+\Omega(y\pi_{x}-x\pi_{y}), (74)

As the system has rotational symmetry, the angular momentum is still conserved. The Schrodinger problem can be solved in polar coordinates by elementary means to obtain the energy eigenvalues222We find it more convenient to not define the principal quantum number to be N=n+|m|N=n+|m| as is conventionally done, in which case m=0,±1,,±Nm=0,\pm 1,\ldots,\pm N.,

Enm\displaystyle E_{nm} =\displaystyle= (n+|m|+1)μΩm,\displaystyle(n+|m|+1)\mu-\Omega m, (75)
n=0,1,2,;m=0,±1,±2,\displaystyle n=0,1,2,\ldots;\ \ m=0,\pm 1,\pm 2,\ldots

Consider the case Ω>0\Omega>0. Then, for fixed nn and m+m\to+\infty, we have Enm(μΩ)mE_{nm}\to(\mu-\Omega)m, and the energy is bounded from below only if μΩ\mu\geq\Omega. Considering that Ω\Omega can have either sign, the condition for the energy to be bounded from below is,

μ|Ω|\mu\geq|\Omega| (76)

and then the ground state is for n=m=0n=m=0 with energy μ\mu.

In the notation of (70), H1H_{1} can be written as

H112πx2+12πy2+Ω(yπxxπy)+ϵ22y2H_{1}\to\frac{1}{2}\pi_{x}^{2}+\frac{1}{2}\pi_{y}^{2}+\Omega(y\pi_{x}-x\pi_{y})+\frac{\epsilon^{2}}{2}y^{2} (77)

The last term breaks rotational symmetry in the xyxy-plane and the angular dependence will not be given by angular momentum eigenstates as in (73). The wavefunction will now be peaked on the xx-axis; it will be bounded in the yy-direction while remaining unbounded in the xx-direction.

In principle, we can also treat the case of inhomogeneous excitations since the Hamiltonian is quadratic. However, the diagonalization will be technically difficult and the angular momentum term may resist complete diagonalization as in the case of (77).

In this section we have discussed quantum aspects of the homogeneous quantum excitations. The energy spectrum for some of these homogeneous excitations is unbounded from below. If higher order interaction terms effectively give masses to these excitation modes, the spectrum gets bounded provided the mass is greater than |Ω||\Omega| (see (76)).

VIII Conclusions

We have found that an electric field in pure non-Abelian SU(2) gauge theory can be stable against quantum dissipation to Schwinger pair production. (Similar constructions can be embedded in theories with larger gauge groups.) This is because the gauge fields underlying the electric field can be chosen as a stationary background in which a rotor degree of freedom is rotating with fixed, quantized, angular momentum in internal space. The quantum state of excitations on this stationary background is also stationary and there is no particle production and no dissipation. This resolves a conundrum that one might intuit from the case of Abelian electric fields, where the electric field dissipates due to pair production, even if at an exponentially suppressed rate. If the same conclusion applied to non-Abelian electric fields, QCD flux tubes would be susceptible to decay due to Schwinger pair production of gluons. Thus we conjecture that QCD flux tubes should be described by gauge fields as given in Eq. (28). If we start with a non-Abelian electric field that is in the Abelian configuration of (8), a guess is that it would evolve into the unexciting non-Abelian configuration of (28).

The existence of stable non-Abelian electric fields opens up a number of related questions. One issue we faced is that we had to postulate classical external charges to source the background electric field. At present it is not known whether such sources need to be external or if they can arise due to the strong interactions as discussed in Sec. IV.1. The key open question is to determine properties of the ground state of quantum excitations around the background electric field. In Sec. IV.2 we have already pointed out the danger posed by unstable modes as then the separation between a classical background and quantum fluctuations is not clear. Assuming that there is a range of parameters where there are no unstable modes, there will be a well-defined ground state which will depend on the profile of the electric field flux tube. Perhaps there are lattice techniques that can determine the optimum flux tube profile for which the overall energy is minimized. The quantum state can also settle the question whether the sources necessary for the stationary background can arise from the internal dynamics of quantum excitations. These are fascinating but difficult questions that we hope to examine in the future.

Another system of interest is that of dyons that carry non-Abelian electric charge, as these can also arise as stationary solutions of non-Abelian Yang-Mills-Higgs theories. We expect such dyons to be stable to Schwinger particle production of non-Abelian gauge bosons but a detailed analysis might yield surprises especially for large Ω\Omega, as in the discussion in Sec. VII.

Acknowledgements.
I am grateful to Gia Dvali, Sang Pyo Kim, Parameswaran Nair, Igor Shovkovy, Frank Wilczek and George Zahariade for helpful comments. T.V. was supported by the U.S. Department of Energy, Office of High Energy Physics, under Award No. DE-SC0019470.

References

  • Hawking (1975) S. W. Hawking, Commun. Math. Phys. 43, 199 (1975), [Erratum: Commun.Math.Phys. 46, 206 (1976)].
  • Birrell and Davies (1984) N. D. Birrell and P. C. W. Davies, Quantum Fields in Curved Space, Cambridge Monographs on Mathematical Physics (Cambridge Univ. Press, Cambridge, UK, 1984), ISBN 978-0-521-27858-4, 978-0-521-27858-4.
  • Schwinger (1951) J. S. Schwinger, Phys. Rev. 82, 664 (1951).
  • Vachaspati (2022) T. Vachaspati, Phys. Rev. D 105, 056008 (2022), eprint 2201.02196.
  • Guéry-Odelin et al. (2019) D. Guéry-Odelin, A. Ruschhaupt, A. Kiely, E. Torrontegui, S. Martínez-Garaot, and J. G. Muga, Rev. Mod. Phys. 91, 045001 (2019), URL https://link.aps.org/doi/10.1103/RevModPhys.91.045001.
  • Parikh et al. (2012) M. Parikh, P. Samantray, and E. Verlinde, Phys. Rev. D 86, 024005 (2012), eprint 1112.3433.
  • Parikh and Samantray (2013) M. Parikh and P. Samantray, Phys. Rev. D 87, 125037 (2013), eprint 1212.4487.
  • Kim and Schubert (2011) S. P. Kim and C. Schubert, Phys. Rev. D 84, 125028 (2011), eprint 1110.0900.
  • Itzykson and Zuber (1980) C. Itzykson and J. B. Zuber, Quantum Field Theory, International Series In Pure and Applied Physics (McGraw-Hill, New York, 1980), ISBN 978-0-486-44568-7.
  • Kim and Page (2002) S. P. Kim and D. N. Page, Phys. Rev. D 65, 105002 (2002), eprint hep-th/0005078.
  • Kim and Page (2006) S. P. Kim and D. N. Page, Phys. Rev. D 73, 065020 (2006), eprint hep-th/0301132.
  • Bogolyubov (1958) N. N. Bogolyubov, Sov. Phys. JETP 7, 51 (1958).
  • Vachaspati and Zahariade (2018) T. Vachaspati and G. Zahariade, Phys. Rev. D 98, 065002 (2018), eprint 1806.05196.
  • Vachaspati and Zahariade (2019) T. Vachaspati and G. Zahariade, JCAP 09, 015 (2019), eprint 1807.10282.
  • Matinyan and Savvidy (1978) S. G. Matinyan and G. K. Savvidy, Nucl. Phys. B 134, 539 (1978).
  • Brown and Weisberger (1979) L. S. Brown and W. I. Weisberger, Nucl. Phys. B 157, 285 (1979), [Erratum: Nucl.Phys.B 172, 544 (1980)].
  • Yildiz and Cox (1980) A. Yildiz and P. H. Cox, Phys. Rev. D 21, 1095 (1980).
  • Ambjorn and Hughes (1982a) J. Ambjorn and R. J. Hughes, Nucl. Phys. B 197, 113 (1982a).
  • Ambjorn and Hughes (1982b) J. Ambjorn and R. J. Hughes, Phys. Lett. B 113, 305 (1982b).
  • Nayak and van Nieuwenhuizen (2005) G. C. Nayak and P. van Nieuwenhuizen, Phys. Rev. D 71, 125001 (2005), eprint hep-ph/0504070.
  • Cooper and Nayak (2006) F. Cooper and G. C. Nayak, Phys. Rev. D 73, 065005 (2006), eprint hep-ph/0511053.
  • Cooper et al. (2008) F. Cooper, J. F. Dawson, and B. Mihaila, in Conference on Nonequilibrium Phenomena in Cosmology and Particle Physics (2008), eprint 0806.1249.
  • Nair and Yelnikov (2010) V. P. Nair and A. Yelnikov, Phys. Rev. D 82, 125005 (2010), eprint 1005.2582.
  • Ilderton (2022) A. Ilderton, Phys. Rev. D 105, 016021 (2022), eprint 2108.13885.
  • Huet et al. (2014) A. Huet, S. P. Kim, and C. Schubert, Phys. Rev. D 90, 125033 (2014), eprint 1411.3074.
  • Ragsdale and Singleton (2017) M. Ragsdale and D. Singleton, J. Phys. Conf. Ser. 883, 012014 (2017), eprint 1708.09753.
  • Karabali et al. (2019) D. Karabali, S. Kurkcuoglu, and V. P. Nair, Phys. Rev. D 100, 065006 (2019), eprint 1905.12391.
  • Cardona and Vachaspati (2021) C. Cardona and T. Vachaspati, Phys. Rev. D 104, 045009 (2021), eprint 2105.08782.
  • Kogut and Susskind (1974) J. B. Kogut and L. Susskind, Phys. Rev. D 9, 3501 (1974).
  • Kogut and Susskind (1975) J. B. Kogut and L. Susskind, Phys. Rev. D 11, 395 (1975).
  • Takahashi et al. (2002) T. T. Takahashi, H. Suganuma, Y. Nemoto, and H. Matsufuru, Phys. Rev. D 65, 114509 (2002), eprint hep-lat/0204011.
  • Bissey et al. (2007) F. Bissey, F.-G. Cao, A. R. Kitson, A. I. Signal, D. B. Leinweber, B. G. Lasscock, and A. G. Williams, Phys. Rev. D 76, 114512 (2007), eprint hep-lat/0606016.
  • Chu and Vachaspati (2010) Y.-Z. Chu and T. Vachaspati, Phys. Rev. D 81, 085020 (2010), eprint 1001.2559.
  • Gold et al. (2021) G. Gold, D. A. Mcgady, S. P. Patil, and V. Vardanyan, JHEP 10, 072 (2021), eprint 2012.15824.
  • Bazak and Mrowczynski (2022) S. Bazak and S. Mrowczynski, Phys. Rev. D 105, 034023 (2022), eprint 2111.11396.
  • Julia and Zee (1975) B. Julia and A. Zee, Phys. Rev. D 11, 2227 (1975).
  • Callan (1982) C. G. Callan, Jr., Phys. Rev. D 26, 2058 (1982).
  • Preskill (1987) J. Preskill, in Les Houches School of Theoretical Physics: Architecture of Fundamental Interactions at Short Distances (1987), pp. 235–338.