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

Titolo

Abstract

We answer the question: If a vacuum sector Hamiltonian is regularized by an energy cutoff, how is the one-kink sector Hamiltonian regularized? We find that it is not regularized by an energy cutoff, indeed normal modes of all energies are present in the kink Hamiltonian, but rather the decomposition of the field into normal mode operators yields coefficients which lie on a constrained surface that forces them to become small for energies above the cutoff. This explains the old observation that an energy cutoff of the kink Hamiltonian leads to an incorrect one-loop kink mass. To arrive at our conclusion, we impose that the regularized kink sector Hamiltonian is unitarily equivalent to the regularized vacuum sector Hamiltonian. This condition implies that the two regularized Hamiltonians have the same spectrum and so guarantees that the kink Hamiltonian yields the correct kink mass.

Cut-Off Kinks



Jarah Evslin***[email protected]1,2, Andrew B. Royston[email protected]3 and Baiyang Zhang[email protected]4


1) Institute of Modern Physics, NanChangLu 509, Lanzhou 730000, China

2) University of the Chinese Academy of Sciences, YuQuanLu 19A, Beijing 100049, China

3) Department of Physics, Penn State Fayette, The Eberly Campus, 2201 University Drive, Lemont Furnace, PA 15456, USA

4) Institute of Contemporary Mathematics, School of Mathematics and Statistics, Henan University, Kaifeng, Henan 475004, P. R. China

1 Introduction

The conventional approach [1, 2, 3, 4] to the computation of the mass of a quantum soliton is as follows. One begins with a Hamiltonian which defines a theory and introduces a one-soliton sector Hamiltonian that describes the same physics in terms of the field expanded about the classical soliton solution. Both Hamiltonians are then regularized.111In the canonical transformation approach of [5], one applies the transformation to a perturbative sector Hamiltonian that includes counterterms. The Hamiltonian is not regularized, however, because it is not expressed in terms of a regularized set of degrees of freedom – Fourier modes up to a cut-off momentum, for example. One defines the soliton mass to be the difference in the eigenvalues between a soliton ground state computed using the one-soliton sector Hamiltonian and the vacuum computed using the original vacuum Hamiltonian. This mass depends on the regulators of both Hamiltonians. Then the regulators must both be taken to infinity. Unfortunately the mass found depends on the choice of how these two regulators are matched as they are taken to infinity [6].

Many prescriptions for this matching have been given in the literature, some leading to the right or the wrong answer. In particular, a cut-off regularization of both Hamiltonians with a matching of the cut-off energies leads to the wrong answer [6].

Several proposals for resolving this problem have appeared in the literature. The dependence on the regulator matching condition arises because of the sharp dependence of the energy on the regulator, which in turn is due to the quadratic ultraviolet divergence in the one-loop energies. In Ref. [7] the authors noted that this problem could be removed by calculating not the energy itself, but rather its derivative with respect to an energy scale, as its divergence will be suppressed by one power. In models simple enough for the constant of integration to be fixed using dimensional analysis, this allows the kink mass to become regulator independent and so one arrives at the right answer, even with the wrong regulator matching condition. While this approach is sufficient for the calculation of the mass in many models of interest, it is not applicable to models with multiple scales and it cannot be used to calculate quantities with a steeper dependence on the energy scale.

These shortcomings motivated Ref. [8] to try to design a prescription for an energy cutoff that would reproduce the known one-loop kink mass. This led them to an appealing physical principle, that normal modes sufficiently above the regulator scale should not be affected by the soliton. Unfortunately this principle alone is not enough to fix the mass, as one must also choose how the density of states scales at high energies and the mass obtained depends on this choice.

The origin of the above difficulties is clear. In standard approaches, one transforms an unregularized defining Hamiltonian to the soliton sector and then regularizes, whereas one should apply the transformation directly to the regularized defining Hamiltonian. For example, the original and accepted result for the one-loop correction to the soliton mass in ϕ4\phi^{4} theory [1] was obtained using mode number regularization. Periodic boundary conditions for a box of size LL are applied to the relevant fluctuation operator in each sector, and the same number of modes is kept in determining the contribution to the energy. This procedure is motivated by the lattice, but the full Hamiltonian was not defined on a lattice in [1]. If it had been, and the map between sectors given within this defining theory, then there would be no ambiguity in determining how the perturbative and soliton sector regulators are related. While ϕ4\phi^{4}-theory has certainly been studied on a finite spatial lattice starting with [9], where the authors were interested in the phase structure of the vacuum, a lattice version of the transformation from perturbative to soliton sectors has not been developed.

A satisfactory resolution to the above issue, then, is to provide a finite-dimensional lattice version of the soliton-sector canonical transformation given in [5, 10]. This is the approach we will follow in a forthcoming paper. It is conceptually straightforward but technically nontrivial, and one might ask if the same idea –transforming the regularized Hamiltonian to the soliton sector– can be accomplished with a defining Hamiltonian regularized by an energy (or momentum) cutoff. This paper answers in the affirmative.

Our work here follows in the footsteps of [8] in that we want to a obtain a consistent energy cut-off regularization and renormalization for kinks in 1+1 dimensions. Like them, we recognize that the kink profile and the normal modes will depend on the regulator and in a consistent treatment these dependencies should be considered, at least to be sure that they do not affect the answer to any given question. However unlike them, we make no conjectures. We derive our answer, as follows.

Consider a real scalar field ϕ(x,t)\phi(x,t). Let ϕ(x,t)=f(x)\phi(x,t)=f(x) be the classical kink solution. The quantum theory will be treated in the Schrodinger picture, and so fields will be considered at a fixed tt, and the tt argument will be dropped from the notation. Now it is conventional to expand ϕ(x)\phi(x) about the classical solution as

ϕ(x)=f(x)+η(x).\phi(x)=f(x)+\eta(x). (1.1)

In this case one could rewrite the Hamiltonian in terms of the quantum field η(x)\eta(x) [11].

We will choose a different approach. It is convenient to expand the field not about f(x)f(x) but about a zero value of the field, so that higher moments of the field correspond to interactions. This could be achieved using a passive transformation of the field ϕη=ϕf\phi\rightarrow\eta=\phi-f. Instead, following [1], we will employ an active transformation of the Hamiltonian and momentum functionals which act on the field. In particular we will transform the Hamiltonian as

H[ϕ,π]H[ϕ,π]=H[ϕ+f,π].H[\phi,\pi]\rightarrow H^{\prime}[\phi,\pi]=H[\phi+f,\pi]. (1.2)

This definition of HH^{\prime} is sufficient for classical field theory, but in quantum field theory HH is regularized and we would like to define a regularized HH^{\prime}.

The observation that underlies this paper is that (1.2) is a unitary equivalence. More specifically, we construct a unitary operator 𝒟f\mathcal{D}_{f} which maps the vacuum sector to the kink sector [12, 13]. We define the regularized kink sector Hamiltonian HH^{\prime} by conjugating the defining, regularized vacuum-sector Hamiltonian HH with this operator 𝒟f\mathcal{D}_{f}

H=𝒟fH𝒟f.H^{\prime}=\mathcal{D}_{f}^{\dagger}H\mathcal{D}_{f}. (1.3)

In this sense, the momentum cutoff Λ\Lambda in the kink sector is inherited from the vacuum sector. Since the two Hamiltonians are similar, they will have the same eigenvalues and their eigenvectors, which differ by the operator 𝒟f\mathcal{D}_{f}, represent the same states. The kink mass is the difference between the eigenvalues of the eigenstates corresponding to the vacuum and the kink ground state. These two eigenstates may be identified in either Hamiltonian, as they have the same eigenvalues and states, however in perturbation theory we may only find the vacuum as an eigenvector |Ω|\Omega\rangle of the vacuum Hamiltonian and the kink ground state |K|K\rangle as an eigenvector |0|0\rangle of the kink Hamiltonian. In other words, once we have used perturbation theory to find the kink ground state as an eigenstate |0|0\rangle of HH^{\prime} then the corresponding eigenstate of the defining Hamiltonian HH is

|K=𝒟f|0.|K\rangle=\mathcal{D}_{f}|0\rangle. (1.4)

Note that this approach is not sensitive to the exact choice of 𝒟f\mathcal{D}_{f}. Any 𝒟f\mathcal{D}_{f} which allows one to find the desired eigenvectors of HH^{\prime} is sufficient, since they all have the same eigenvalues and the eigenvectors are easily mapped between the various eigenspaces using 𝒟f\mathcal{D}_{f}.

The regulator Λ\Lambda can then be taken to infinity unambiguously, using the renormalization conditions, to arrive at the renormalized kink mass. This limit is unambiguous as there is only one regulator Λ\Lambda which needs to be taken to infinity, not two. The similarity transformation (1.3) guarantees the correct regulator for the kink Hamiltonian. Previous approaches to an energy cutoff in these models failed222In Ref. [6] it was shown that they yield the wrong one-loop kink mass. because they regularized the kink sector Hamiltonian by imposing an energy cutoff in the spectrum of normal modes, but we will see that HH^{\prime} includes all normal modes.

In this work we only include an ultraviolet (UV) regulator since the computations considered will not require infrared (IR) regularization. In forthcoming work we will present a soliton-sector transformation for a finite-dimensional lattice regularization that uses both.

Infrared (IR) regularization via compactification or the inclusion of antisolitons is common in the literature. However, such IR regularization is necessarily different for the vacuum and soliton sectors, which makes it difficult to treat nonperturbative effects which mix distinct sectors. Mixing may seem unimportant at weak coupling, but at strong coupling it is responsible for the symmetry restoring phase transition in the ϕ4\phi^{4} theory. Also the dual Thirring description for the Sine-Gordon model includes a mixing of two-fermion and zero-fermion states.

We begin in Sec. 2 by defining the regularized theory and the vacuum sector. Next in Sec. 3 we define the similarity transform which takes the defining Hamiltonian to the kink sector Hamiltonian. Finally in Sec. 4 we find the regularized kink sector Hamiltonian and calculate its one-loop ground state and mass. When the regulator is taken to infinity, we reproduce the known mass formula. Our notation is summarized in Table 1.

Operator Description
ϕ(Λ)(x),π(Λ)(x)\phi^{(\Lambda)}(x),\ \pi^{(\Lambda)}(x) The real scalar field and its conjugate momentum
Ap,ApA^{\ddagger}_{p},\ A_{p} Creation and annihilation operators in plane wave basis
Bk,BkB^{\ddagger}_{k},\ B_{k} Creation and annihilation operators in normal mode basis
BBO,BBOB^{\ddagger}_{BO},\ B_{BO} Creation/annihilation operators of odd shape modes
ϕB,πB\phi_{B},\ \pi_{B} Zero mode of ϕ(x)\phi(x) and π(x)\pi(x) in normal mode basis
::a::_{a} Normal ordering with respect to AA operators respectively
Hamiltonian Description
HH The original Hamiltonian
HH^{\prime} HH with ϕ(Λ)(x)\phi^{(\Lambda)}(x) shifted by regularized kink solution f(Λ)(x)f^{(\Lambda)}(x)
HnH^{\prime}_{n} The ϕ(Λ)n\phi^{{(\Lambda)}n} term in HH^{\prime}
Symbol Description
Λ\Lambda Momentum cutoff
f0(x)f_{0}(x) The unregularized classical kink solution
f1(x)f_{1}(x) First correction to the unregularized classical kink solution
f(Λ)(x)f^{(\Lambda)}(x) Regularized classical kink solution
𝒟f\mathcal{D}_{f} Operator that translates ϕ(Λ)(x)\phi^{(\Lambda)}(x) by the regularized classical kink solution
gB(x)g_{B}(x) The kink linearized translation mode
gk(x)g_{k}(x) Normal modes including discrete modes
pp Momentum
kk Normal mode label
ωk,ωp\omega_{k},\ \omega_{p} The frequency corresponding to kk or pp
g~\tilde{g} Inverse Fourier transform of gg
State Description
|K,|Ω|K\rangle,\ |\Omega\rangle Kink and vacuum sector ground states
Table 1: Summary of Notation

2 The Vacuum Sector

2.1 Classical Theory

Let us consider a classical theory of a real scalar field ϕ~(Λ)(x)\tilde{\phi}^{(\Lambda)}(x) and its conjugate π(Λ)(x)\pi^{(\Lambda)}(x) in 1+1 dimensions,

H=𝑑x,=π(Λ)2(x)+xϕ~(Λ)(x)xϕ~(Λ)(x)2+V~[gϕ~(Λ)]g2+δm22ϕ~(Λ)2+γ~H=\int dx\mathcal{H},\hskip 21.68121pt\mathcal{H}=\frac{\pi^{{(\Lambda)}2}(x)+\partial_{x}\tilde{\phi}^{(\Lambda)}(x)\partial_{x}\tilde{\phi}^{(\Lambda)}(x)}{2}+\frac{\tilde{V}[g\tilde{\phi}^{(\Lambda)}]}{g^{2}}+\frac{\delta m^{2}}{2}\tilde{\phi}^{{(\Lambda)}2}+\tilde{\gamma} (2.1)

where we have included counterterms δm2\delta m^{2} and γ~\tilde{\gamma} and a potential which in the case of the ϕ4\phi^{4} double well theory is

V~[gϕ~(Λ)]=14(g2ϕ~(Λ)2g2v2)2.\tilde{V}[g\tilde{\phi}^{(\Lambda)}]=\frac{1}{4}\left(g^{2}\tilde{\phi}^{{(\Lambda)}2}-g^{2}v^{2}\right)^{2}. (2.2)

The potential is always written as a function of gϕg\phi because this combination is dimensionless. The notation (Λ){(\Lambda)} emphasizes that the quantity in question depends on the regulator Λ\Lambda. In this note we will focus on the ϕ4\phi^{4} theory, where renormalization conditions can be chosen such that the coupling is not renormalized and the only required counterterms are those proportional to δm2\delta m^{2} and γ~\tilde{\gamma} [6]. However the strategy can be readily generalized to other potentials by adding counterterms for the additional interactions and corresponding renormalization conditions.

We will be interested in cases in which V[gϕ]V[g\phi] has multiple minima, so that there is a kink solution in the classical theory. To simplify the discussion below, we will shift the field by the location of one of these minima, v-v by defining

ϕ(Λ)(x)=ϕ~(Λ)(x)+v,V[gϕ(Λ)]=V~[g(ϕ(Λ)v)],γ=γ~+δm22v2\phi^{(\Lambda)}(x)=\tilde{\phi}^{(\Lambda)}(x)+v,\hskip 21.68121ptV[g\phi^{(\Lambda)}]=\tilde{V}[g(\phi^{(\Lambda)}-v)],\hskip 21.68121pt\gamma=\tilde{\gamma}+\frac{\delta m^{2}}{2}v^{2} (2.3)

leading to

=π(Λ)2(x)+xϕ(Λ)(x)xϕ(Λ)(x)2+V[gϕ(Λ)]g2+δm22ϕ(Λ)2vδm2ϕ(Λ)(x)+γ.\mathcal{H}=\frac{\pi^{{(\Lambda)}2}(x)+\partial_{x}{\phi}^{(\Lambda)}(x)\partial_{x}{\phi}^{(\Lambda)}(x)}{2}+\frac{{V}[g{\phi}^{(\Lambda)}]}{g^{2}}+\frac{\delta m^{2}}{2}{\phi}^{{(\Lambda)}2}-v\delta m^{2}\phi^{(\Lambda)}(x)+{\gamma}. (2.4)

In the case of the ϕ4\phi^{4} double well one obtains

V[gϕ(Λ)]=g2ϕ(Λ)24(gϕ(Λ)2gv)2.{V}[g{\phi}^{(\Lambda)}]=\frac{g^{2}\phi^{{(\Lambda)}2}}{4}\left(g\phi^{{(\Lambda)}}-2gv\right)^{2}. (2.5)

2.2 Quantum Theory

As the action has the dimensions of \hbar, the operator ϕ(Λ)(x)\phi^{(\Lambda)}(x) has dimensions O(1/2)O(\hbar^{1/2}) while gg has dimensions O(1/2)O(\hbar^{-1/2}). Therefore vv also has dimensions O(1/2)O(\hbar^{1/2}). This means that in the quantum theory, they will appear in the dimensionless combinations g1/2g\hbar^{1/2} and v1/2v\hbar^{-1/2}. Thus in the semiclassical expansions below, each power of vv will be treated as one power of 1/g1/g.

We will not directly quantize ϕ(Λ)(x)\phi^{(\Lambda)}(x) and π(Λ)(x)\pi^{(\Lambda)}(x). Instead we will decompose

ϕ(Λ)(x)=ΛΛdp2πϕpeipx,π(Λ)(x)=ΛΛdp2ππpeipx\phi^{(\Lambda)}(x)=\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\phi_{p}e^{-ipx},\hskip 21.68121pt\pi^{(\Lambda)}(x)=\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\pi_{p}e^{-ipx} (2.6)

and we will quantize ϕp\phi_{p} and πp\pi_{p} by imposing

[ϕp,πq]=2πiδ(p+q).[\phi_{p},\pi_{q}]=2\pi i\delta(p+q). (2.7)

This note will be entirely in the Schrodinger picture, and so (2.7) defines Schrodinger operators ϕ(Λ)(x)\phi^{(\Lambda)}(x) and π(Λ)(x)\pi^{(\Lambda)}(x). They are not local fields; indeed they satisfy

[ϕ(Λ)(x),π(Λ)(y)]=iΛΛdp2πeip(xy)=iα(xy),α(x)=sin(Λx)πx.\left[\phi^{(\Lambda)}(x),\pi^{(\Lambda)}(y)\right]=i\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}e^{-ip(x-y)}=i\alpha(x-y),\hskip 21.68121pt\alpha(x)=\frac{\textrm{sin}(\Lambda x)}{\pi x}. (2.8)

The nonlocality of the commutator arises from the fact that the integration in the momentum space is cut off by Λ\Lambda. Note that locality is restored as the cutoff is removed because

Λlimα(x)=δ(x).\stackrel{{\scriptstyle\rm{lim}}}{{{}_{\Lambda\rightarrow\infty}}}\alpha(x)=\delta(x). (2.9)

Interestingly, (2.8) are the same canonical commutation relations found in Pearson’s thesis [14] (see also [15]) for an infinite lattice regularization with lattice spacing 2π/Λ2\pi/\Lambda. In this context, ϕ(Λ),π(Λ)\phi^{(\Lambda)},\pi^{(\Lambda)} are interpolating fields for underlying lattice degrees of freedom that do satisfy canonical commutation relations. Using the interpolating fields to define the regularized Hamiltonian, as we have done here, implies that our theory is equivalent to the infinite lattice theory of Pearson. We do not utilize the lattice language in this paper since it is not advantageous for the computations we perform. We will utilize it, however, in forthcoming work when we construct a finite-dimensional version of the soliton-sector canonical transformation.

Our problem is now well-defined. One can insert (2.6) into (2.4) to obtain the Hamiltonian as a function of the operators ϕp\phi_{p} and πp\pi_{p}. The vacuum and the kink ground state will be eigenstates of this Hamiltonian, and the difference between their eigenvalues is the kink mass. The Schrodinger picture is sufficient for determining the spectra and so we need never introduce time, and the nonlocality of our operators will not impede our narrow task.

To do this concretely, we will choose the renormalization condition

H|Ω=0H|\Omega\rangle=0~{} (2.10)

and will expand

|Ω=i=0|Ωi,|Ωi=O(g2i),γ=i=0γi,γi=O(g2i).|\Omega\rangle=\sum_{i=0}^{\infty}|\Omega\rangle_{i},\hskip 21.68121pt|\Omega\rangle_{i}=O(g^{2i}),\hskip 21.68121pt\gamma=\sum_{i=0}^{\infty}\gamma_{i},\hskip 21.68121pt\gamma_{i}=O(g^{2i}). (2.11)

Let us begin with the case i=0i=0. We are interested in terms in \mathcal{H} with no powers of gg. Let us assume for the moment that δm2\delta m^{2} contains only terms of order at least O(g2)O(g^{2}). We will see shortly that this choice is consistent. Then we are left with

H0\displaystyle H_{0} =\displaystyle= 𝑑x0,\displaystyle\int dx\mathcal{H}_{0}, (2.12)
0\displaystyle\mathcal{H}_{0} =\displaystyle= π(Λ)2(x)+xϕ(Λ)(x)xϕ(Λ)(x)+m2ϕ(Λ)2(x)2+γ0,m=V′′[0].\displaystyle\frac{\pi^{{(\Lambda)}2}(x)+\partial_{x}{\phi}^{(\Lambda)}(x)\partial_{x}{\phi}^{(\Lambda)}(x)+m^{2}\phi^{{(\Lambda)}2}(x)}{2}+{\gamma}_{0},\hskip 21.68121ptm=\sqrt{{V}^{\prime\prime}[0]}.

For the ϕ4\phi^{4} double well, m=2vgm=\sqrt{2}vg. Define the linear combinations

Ap=ϕp2iπp2ωp,Ap=ωpϕp+iπp,ωp=m2+p2A^{\ddagger}_{p}=\frac{\phi_{p}}{2}-i\frac{\pi_{p}}{2\omega_{p}},\hskip 21.68121ptA_{-p}=\omega_{p}\phi_{p}+i\pi_{p},\hskip 21.68121pt\omega_{p}=\sqrt{m^{2}+p^{2}} (2.13)

where the Hermitian conjugate of ApA_{p} is 2ωpAp2\omega_{p}A^{\ddagger}_{p}. The canonical commutation relations (2.7) imply that these each satisfy a Heisenberg algebra

[Ap,Aq]=2πδ(pq).[A_{p},A^{\ddagger}_{q}]=2\pi\delta(p-q). (2.14)

To impose (2.10) at O(g0)O(g^{0}), it suffices to consider |Ω0|\Omega\rangle_{0} which is the free vacuum

Ap|Ω0=0.A_{p}|\Omega\rangle_{0}=0. (2.15)

The definitions (2.13) are easily inverted

ϕp=Ap+Ap2ωp,πp=i(ωpApAp2).\phi_{p}=A^{\ddagger}_{p}+\frac{A_{-p}}{2\omega_{p}},\hskip 21.68121pt\pi_{p}=i\left(\omega_{p}A^{\ddagger}_{p}-\frac{A_{-p}}{2}\right). (2.16)

Substituting this into (2.12) one finds

0=H0|Ω0=𝑑x(γ0+ΛΛdp2πωp2)|Ω0.0=H_{0}|\Omega\rangle_{0}=\int dx\left(\gamma_{0}+\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\frac{\omega_{p}}{2}\right)|\Omega\rangle_{0}. (2.17)

and so the leading order counterterm is

γ0=ΛΛdp2πωp2=14π[ΛωΛ+m2ln(Λ+ωΛm)]\gamma_{0}=-\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\frac{\omega_{p}}{2}=-\frac{1}{4\pi}\left[\Lambda\omega_{\Lambda}+m^{2}{\rm{ln}}\left(\frac{\Lambda+\omega_{\Lambda}}{m}\right)\right] (2.18)

where we have integrated by parts. We have now satisfied the renormalization condition (2.10) at order O(g0)O(g^{0}).

To fix δm2\delta m^{2} we will impose another renormalization condition, that tadpoles vanish. Recall that δm2\delta m^{2} is a series in gg beginning at order O(g2)O(g^{2}). For the computation of these leading order terms, the no-tadpole condition is equivalent to imposing that H|Ω0H|\Omega\rangle_{0} contains no terms of the form A|Ω0A^{\ddagger}|\Omega\rangle_{0} at O(g)O(g). The relevant O(g)O(g) terms in the Hamiltonian are

H1=𝑑x1,1=gV(3)[0]ϕ(Λ)3(x)6vδm2ϕ(Λ)(x)H_{1}=\int dx\mathcal{H}_{1},\hskip 21.68121pt\mathcal{H}_{1}=g\frac{V^{(3)}[0]\phi^{{(\Lambda)}3}(x)}{6}-v\delta m^{2}\phi^{(\Lambda)}(x) (2.19)

where V(n)[0]V^{(n)}[0] is the nnth derivative of VV and we recall that vv is of order O(1/g)O(1/g).

Using the terms with a single AA^{\ddagger}

𝑑xϕ(Λ)(x)|Ω0=A0|Ω0,𝑑xϕ(Λ)3(x)|Ω03ΛΛdp2π12ωpA0|Ω0\int dx\phi^{(\Lambda)}(x)|\Omega\rangle_{0}=A^{\ddagger}_{0}|\Omega\rangle_{0},\hskip 21.68121pt\int dx\phi^{{(\Lambda)}3}(x)|\Omega\rangle_{0}\supset 3\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\frac{1}{2\omega_{p}}A^{\ddagger}_{0}|\Omega\rangle_{0} (2.20)

one obtains

δm2=g2vV(3)[0]ΛΛdp2π12ωp=g4πvV(3)[0]ln(Λ+ωΛm).\delta m^{2}=\frac{g}{2v}V^{(3)}[0]\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\frac{1}{2\omega_{p}}=\frac{g}{4\pi v}V^{(3)}[0]{\rm{ln}}\left(\frac{\Lambda+\omega_{\Lambda}}{m}\right). (2.21)

In the case of the ϕ4\phi^{4} double well this is

δm2=3g22πln(Λ+ωΛm).\delta m^{2}=-\frac{3g^{2}}{2\pi}{\rm{ln}}\left(\frac{\Lambda+\omega_{\Lambda}}{m}\right). (2.22)

2.3 The Relation to Normal Ordering

One may define a normal ordering on the operators AA^{\ddagger} and AA by placing all of the former on the left. We will denote this normal ordering :O:a:O:_{a} for any operator OO. It is easily evaluated on the operators appearing in the Hamiltonian at the orders considered above

π(Λ)2(x)\displaystyle\pi^{{(\Lambda)}2}(x) =\displaystyle= :π(Λ)2(x):a+ΛΛdp2πωp2,\displaystyle:\pi^{{(\Lambda)}2}(x):_{a}+\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\frac{\omega_{p}}{2}, (2.23)
ϕ(Λ)2(x)\displaystyle\phi^{{(\Lambda)}2}(x) =\displaystyle= :ϕ(Λ)2(x):a+ΛΛdp2π12ωp,\displaystyle:\phi^{{(\Lambda)}2}(x):_{a}+\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\frac{1}{2\omega_{p}},
(xϕ(Λ)(x))2\displaystyle\left(\partial_{x}\phi^{{(\Lambda)}}(x)\right)^{2} =\displaystyle= :(xϕ(Λ)(x))2:a+ΛΛdp2πp22ωp,\displaystyle:\left(\partial_{x}\phi^{{(\Lambda)}}(x)\right)^{2}:_{a}+\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\frac{p^{2}}{2\omega_{p}},
ϕ(Λ)3(x)\displaystyle\phi^{{(\Lambda)}3}(x) =\displaystyle= :ϕ(Λ)3(x):a+3ϕ(Λ)(x)ΛΛdp2π12ωp.\displaystyle:\phi^{{(\Lambda)}3}(x):_{a}+3\phi^{(\Lambda)}(x)\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\frac{1}{2\omega_{p}}.

Inserting these into (2.12) and (2.19) one finds

0\displaystyle\mathcal{H}_{0} =\displaystyle= :π(Λ)2(x)+xϕ(Λ)(x)xϕ(Λ)(x)+m2ϕ(Λ)2(x):a2\displaystyle\frac{:\pi^{{(\Lambda)}2}(x)+\partial_{x}{\phi}^{(\Lambda)}(x)\partial_{x}{\phi}^{(\Lambda)}(x)+m^{2}\phi^{{(\Lambda)}2}(x):_{a}}{2}
1\displaystyle\mathcal{H}_{1} =\displaystyle= gV(3)[0]:ϕ(Λ)3(x):a6.\displaystyle g\frac{V^{(3)}[0]:\phi^{{(\Lambda)}3}(x):_{a}}{6}. (2.24)

We see that order by order, properly chosen renormalization conditions are equivalent to a Hamiltonian that is normal ordered from the beginning. More specifically, the required renormalization condition sets to zero all diagrams with a loop involving a single vertex. In the case of two-dimensional scalar field theories, normal ordering is sufficient to remove all divergences (see for example Ref. [16]). Thus it would be possible to take the limit Λ\Lambda\rightarrow\infty now. This would reduce the problem to that solved already in Ref. [17, 18] at one loop and [19] at two loops where it was shown that one arrives at the correct kink mass. However, we will not follow this approach as it will not generalize to theories with fermions and theories in higher dimensions, which motivate this line of research.

In the case of the ϕ4\phi^{4} theory, with the renormalization conditions of Subsec. 2.2, the total Hamiltonian density is then

=:π(Λ)2(x):a+:xϕ(Λ)(x)xϕ(Λ)(x):a2+:V[gϕ(Λ)]:ag2+γ~1+i=2γi.\mathcal{H}=\frac{:\pi^{{(\Lambda)}2}(x):_{a}+:\partial_{x}{\phi}^{(\Lambda)}(x)\partial_{x}{\phi}^{(\Lambda)}(x):_{a}}{2}+\frac{:{V}[g{\phi}^{(\Lambda)}]:_{a}}{g^{2}}+\tilde{\gamma}_{1}+\sum_{i=2}^{\infty}\gamma_{i}. (2.25)

Here we have shifted the subleading counterterm to absorb the contribution from Wick contractions appearing in the normal ordering of the ϕ4\phi^{4} interaction

γ~1=γ13g24(ΛΛdp2π12ωp)2.\tilde{\gamma}_{1}=\gamma_{1}-\frac{3g^{2}}{4}\left(\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\frac{1}{2\omega_{p}}\right)^{2}. (2.26)

In Refs. [20, 21], the calculation of γ~1\tilde{\gamma}_{1} was reviewed and it played an essential role in the elimination of IR divergences in the calculation of the two-loop kink mass. The shifted counterterm and all successive terms are finite at Λ\Lambda\rightarrow\infty, with the first few given in Eq. (47) of Ref. [22]. In the case of a general potential, normal ordering again leads to such infinite shifts and finite remainders.

3 Defining the Kink Sector

3.1 The Similarity Transformation

Let f(x)f(x) be a real-valued function. Define the displacement operator

𝒟f=exp(i𝑑xf(x)π(Λ)(x))\mathcal{D}_{f}={\rm{exp}}\left(-i\int dxf(x)\pi^{(\Lambda)}(x)\right) (3.1)

and

f~(p)=𝑑xf(x)eipx.{\tilde{f}}(p)=\int dxf(x)e^{ipx}. (3.2)

This integral is often divergent. Our prescription for defining f~(p){\tilde{f}}(p) in this case is described in Appendix A.

Exponentiating the commutators

[𝑑xf(x)π(Λ)(x),Aq]=idp2πωpf~(p)[Ap,Aq]=iωqf~(q)\left[\int dxf(x)\pi^{(\Lambda)}(x),A_{q}\right]=i\int\frac{dp}{2\pi}\omega_{p}{\tilde{f}}(-p)\left[A^{\ddagger}_{p},A_{q}\right]=-i\omega_{q}{\tilde{f}}(-q) (3.3)

and

[𝑑xf(x)π(Λ)(x),Aq]=idp2π12f~(p)[Ap,Aq]=i2f~(q)\left[\int dxf(x)\pi^{(\Lambda)}(x),A^{\ddagger}_{q}\right]=-i\int\frac{dp}{2\pi}\frac{1}{2}{\tilde{f}}(-p)\left[A_{-p},A^{\ddagger}_{q}\right]=-\frac{i}{2}{\tilde{f}}(q) (3.4)

one finds

[𝒟f,Ap]=ωpf~(p)𝒟f,[𝒟f,Ap]=f~(p)2𝒟f.\left[\mathcal{D}_{f},A_{p}\right]=-\omega_{p}{\tilde{f}}(-p)\mathcal{D}_{f},\hskip 21.68121pt\left[\mathcal{D}_{f},A^{\ddagger}_{p}\right]=-\frac{{\tilde{f}}(p)}{2}\mathcal{D}_{f}. (3.5)

Therefore

[𝒟f,ϕ(Λ)(x)]=ΛΛdp2π([𝒟f,Ap]+[𝒟f,Ap]2ωp)eipx=f(Λ)(x)𝒟f[\mathcal{D}_{f},\phi^{(\Lambda)}(x)]=\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\left([\mathcal{D}_{f},A^{\ddagger}_{p}]+\frac{[\mathcal{D}_{f},A_{-p}]}{2\omega_{p}}\right)e^{-ipx}=-f^{(\Lambda)}(x)\mathcal{D}_{f} (3.6)

where

f(Λ)(x)=ΛΛdp2πf~(p)eipxf^{(\Lambda)}(x)=\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}{\tilde{f}}(p)e^{-ipx} (3.7)

satisfies f(Λ)(x)=f(x)f^{(\Lambda)}(x)=f(x) if f~(q)=0{\tilde{f}}(q)=0 for all qq with |q|>Λ|q|>\Lambda. Note that 𝒟f\mathcal{D}_{f} commutes with π(Λ)\pi^{(\Lambda)} as [πp,πq]=0[\pi_{p},\pi_{q}]=0.

Finally we are ready to introduce the similarity transformation at the heart of our construction. The kink sector Hamiltonian HH^{\prime} is defined by

𝒟fH[ϕ(Λ)(x),π(Λ)(x)]𝒟f=H[ϕ(Λ)(x),π(Λ)(x)]=H[ϕ(Λ)(x)+f(Λ)(x),π(Λ)(x)]\mathcal{D}_{f}^{\dagger}H[\phi^{(\Lambda)}(x),\pi^{(\Lambda)}(x)]\mathcal{D}_{f}=H^{\prime}[\phi^{(\Lambda)}(x),\pi^{(\Lambda)}(x)]=H[\phi^{(\Lambda)}(x)+f^{(\Lambda)}(x),\pi^{(\Lambda)}(x)] (3.8)

for a suitable choice of f(x)f(x), which will be made momentarily. In the rest of this note we will be concerned with HH^{\prime}. As it is similar to HH, it has the same spectrum. In particular if

H|K=EK|KH|K\rangle=E_{K}|K\rangle (3.9)

then, using (1.4),

H|0=EK|0H^{\prime}|0\rangle=E_{K}|0\rangle (3.10)

and so HH^{\prime} may be used to compute the energy of any state, and in particular the energy EKE_{K} of the kink ground state.

3.2 The Kink-Sector Tadpole

Let us consider f(Λ)(x)f^{(\Lambda)}(x) to be of order O(1/g)O(1/g). Then the kink sector Hamiltonian HH^{\prime} can be expanded order by order in the coupling gg

H=i=0Hi,Hi=𝑑xiH^{\prime}=\sum_{i=0}^{\infty}H^{\prime}_{i},\hskip 21.68121ptH^{\prime}_{i}=\int dx\mathcal{H}^{\prime}_{i} (3.11)

where each HiH^{\prime}_{i} and i\mathcal{H}^{\prime}_{i} is of order O(gi2)O(g^{i-2}). At leading order one finds the classical energy corresponding to the field configuration ϕ(Λ)(x)=f(Λ)(x)\phi^{(\Lambda)}(x)=f^{(\Lambda)}(x)

0=(xf(Λ)(x))22+V[gf(Λ)(x)]g2.\mathcal{H}^{\prime}_{0}=\frac{\left(\partial_{x}f^{(\Lambda)}(x)\right)^{2}}{2}+\frac{V[gf^{{(\Lambda)}}(x)]}{g^{2}}. (3.12)

The kink-sector tadpole appears at the next order:

H1=𝑑xϕ(Λ)(x)[x2f(Λ)(x)+V[gf(Λ)(x)]/g].H^{\prime}_{1}=\int dx\phi^{(\Lambda)}(x)\left[-\partial^{2}_{x}f^{{(\Lambda)}}(x)+V^{\prime}[gf^{(\Lambda)}(x)]/g\right]. (3.13)

We will define f(Λ)f^{(\Lambda)} by imposing that the expression in brackets vanishes at momenta below the cutoff333Note that, although the higher modes of f(Λ)f^{{(\Lambda)}} vanish, those of the VV^{\prime} term do not vanish. However they do not contribute to the tadpole in Eq. (3.13) as the higher modes of ϕ(Λ)\phi^{(\Lambda)} vanish.

𝑑x[x2f(Λ)(x)+V[gf(Λ)(x)]/g]eipx=0,|p|Λ\int dx\left[-\partial^{2}_{x}f^{{(\Lambda)}}(x)+V^{\prime}[gf^{(\Lambda)}(x)]/g\right]e^{ipx}=0,\hskip 21.68121pt|p|\leq\Lambda (3.14)

which will automatically eliminate the kink sector tadpole as ϕΛ(x)\phi^{\Lambda}(x) satisfies the opposite condition

𝑑xϕ(Λ)(x)eipx=0,|p|>Λ.\int dx\phi^{(\Lambda)}(x)e^{ipx}=0,\hskip 21.68121pt|p|>\Lambda. (3.15)

In the case Λ\Lambda\rightarrow\infty this condition corresponds to the classical equations of motion for a time-independent configuration in the unregularized theory. Thus f()(x)f^{(\infty)}(x) will be equal to the classical kink solution f0(x)f_{0}(x) and 0\mathcal{H}^{\prime}_{0} will reduce to its classical mass density.

The definition (3.14) can be solved as an expansion in large Λ\Lambda about the classical solution f0(x)f_{0}(x). To illustrate this procedure, in Appendix A we find the leading correction in the case of the ϕ4\phi^{4} double well and show that it is exponentially small in Λ/m\Lambda/m.

4 The Kink Sector

4.1 The Setup

In this section we will restrict our attention to the ϕ4\phi^{4} double well theory. As noted above, an analogous treatment of models with other interactions generically requires additional counterterms multiplying various powers of ϕ(Λ)\phi^{(\Lambda)}. Inserting (2.25) into (3.8) one finds that the gg-independent terms are

H2=𝑑x:π(Λ)2(x):a+:xϕ(Λ)(x)xϕ(Λ)(x):a+V′′[gf(Λ)(x)]:ϕ(Λ)2(x):a2.H^{\prime}_{2}=\int dx\frac{:\pi^{{(\Lambda)}2}(x):_{a}+:\partial_{x}{\phi}^{(\Lambda)}(x)\partial_{x}{\phi}^{(\Lambda)}(x):_{a}+V^{\prime\prime}[gf^{(\Lambda)}(x)]:\phi^{{(\Lambda)}2}(x):_{a}}{2}. (4.1)

These are the only terms that contribute at one loop, as they are suppressed by a factor of g2g^{2} with respect to the leading terms in H0H^{\prime}_{0} and we have chosen f(Λ)(x)f^{(\Lambda)}(x) so that H1H^{\prime}_{1} vanishes. The term gf(Λ)gf^{(\Lambda)} is gg-independent and becomes gf0gf_{0} as Λ\Lambda\to\infty. However, we argued in Appendix A that the corrections f(Λ)f0f^{(\Lambda)}-f_{0} are exponentially small in m/Λm/\Lambda. Since perturbative computations only produce power-law divergences in Λ\Lambda, such exponentially small corrections do not contribute to any perturbative quantities. Therefore in the following we will replace V′′[gf(Λ)(x)]V^{\prime\prime}[gf^{(\Lambda)}(x)] with V′′[gf0(x)]V^{\prime\prime}[gf_{0}(x)].

For completeness we note

H3\displaystyle H^{\prime}_{3} =\displaystyle= 𝑑xgV(3)[gf(Λ)(x)]:ϕ(Λ)2(x):a6\displaystyle\int dx\frac{gV^{(3)}[gf^{(\Lambda)}(x)]:\phi^{{(\Lambda)}2}(x):_{a}}{6} (4.2)
H4\displaystyle H^{\prime}_{4} =\displaystyle= 𝑑x[g2V(4)[gf(Λ)(x)]:ϕ(Λ)2(x):a24+γ~1].\displaystyle\int dx\left[\frac{g^{2}V^{(4)}[gf^{(\Lambda)}(x)]:\phi^{{(\Lambda)}2}(x):_{a}}{24}+\tilde{\gamma}_{1}\right].

In the rest of this note we will use (4.1) to study the one-loop kink spectrum. This equation describes a theory which is free, as the Hamiltonian is quadratic in the field. However the operators AA and AA^{\ddagger} do not diagonalize the Hamiltonian as a result of the xx-dependent mass term. Our strategy will thus be to diagonalize HH using a Bogoliubov transformation from the basis of operators that create plane waves to a basis of operators that create normal modes of the regularized kink.

First let us find these normal modes. Inserting the Ansatz444This constant frequency Ansatz is inconsistent with the restriction that the higher Fourier modes of ϕ(Λ)\phi^{(\Lambda)} vanish. However, when we construct the quantum field ϕ(Λ)\phi^{(\Lambda)} below, we will consider only those linear combinations which satisfy the restriction. The key observation will be that these combinations do not have constant frequency, and so our regularized kink Hamiltonian is not obtained by cutting off frequencies above any sharp threshold.

ϕ(Λ)(x,t)=eiωktgk(x)\phi^{(\Lambda)}(x,t)=e^{i\omega_{k}t}g_{k}(x) (4.3)

into the classical equations of motion for (3.8) one finds

V′′[gf0(x)]gk(x)=(ωk2+x2)gk(x).V^{\prime\prime}[gf_{0}(x)]g_{k}(x)=(\omega_{k}^{2}+\partial_{x}^{2})g_{k}(x). (4.4)

The index kk will include a continuous spectrum with energy

ωk=m2+k2\omega_{k}=\sqrt{m^{2}+k^{2}} (4.5)

as well as discrete modes. In the case of the ϕ4\phi^{4} modes there are two discrete solutions, the zero mode gB(x)g_{B}(x) corresponding to the translation symmetry with ωB=0\omega_{B}=0 and also an odd shape gBO(x)g_{BO}(x) with ωBO=m3/2\omega_{BO}=m\sqrt{3}/2.

As VV is real we may choose

gk(x)=gk(x)g^{*}_{k}(x)=g_{k}(-x) (4.6)

and so gB(x)g_{B}(x) is real, gBO(x)g_{BO}(x) is imaginary and in the case of continuum modes

gk(x)=gk(x).g^{*}_{k}(x)=g_{-k}(x). (4.7)

We normalize the solutions such that

𝑑x|gB(x)|2=𝑑x|gBO(x)|2=1\int dx|g_{B}(x)|^{2}=\int dx|g_{BO}(x)|^{2}=1 (4.8)

and in the case of continuum modes

𝑑xgk1(x)gk2(x)=2πδ(k1+k2).\int dxg_{k_{1}}(x)g_{k_{2}}(x)=2\pi\delta(k_{1}+k_{2}). (4.9)

These satisfy the completeness relation

gB(x)gB(y)+gBO(x)gBO(y)+dk2πgk(x)gk(y)=δ(xy).g_{B}(x)g^{*}_{B}(y)+g_{BO}(x)g^{*}_{BO}(y)+\int\frac{dk}{2\pi}g_{k}(x)g_{-k}(y)=\delta(x-y). (4.10)

From here on we will adopt the shorthand that dk2π\int\frac{dk}{2\pi} implicitly includes a sum over the discrete modes gBg_{B} and gBOg_{BO}. We also introduce the inverse Fourier transform

g~k(p)=𝑑xgk(x)eipx.\tilde{g}_{k}(p)=\int dxg_{k}(x)e^{ipx}. (4.11)

The completeness relation (4.10) implies that the functions gk(x)g_{k}(x) are a basis of all functions so we may decompose

ϕ(Λ)(x)=ΛΛdp2πϕpeipx=dk2πϕk(Λ)gk(x),π(Λ)(x)=ΛΛdp2ππpeipx=dk2ππk(Λ)gk(x).\phi^{(\Lambda)}(x)=\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\phi_{p}e^{-ipx}=\int\frac{dk}{2\pi}\phi^{(\Lambda)}_{k}g_{k}(x),\hskip 21.68121pt\pi^{(\Lambda)}(x)=\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\pi_{p}e^{-ipx}=\int\frac{dk}{2\pi}\pi^{(\Lambda)}_{k}g_{k}(x). (4.12)

Recall that the kk integral implicitly sums over bound states, and so for example we have also introduced ϕB\phi_{B} and πB\pi_{B}. Note that the kk integrals are never cut off.

As the pp integral is bounded, the field satisfies the constraint

𝑑xϕ(Λ)(x)eiqx=𝑑xπ(Λ)(x)eiqx=0,|q|>Λ.\int dx\phi^{(\Lambda)}(x)e^{iqx}=\int dx\pi^{(\Lambda)}(x)e^{iqx}=0,\hskip 21.68121pt|q|>\Lambda. (4.13)

The kk integral is not bounded, indeed including the discrete sums it runs over a complete basis. Therefore (4.13) implies a constraint on the coefficients

0=dk2πϕk(Λ)g~k(q)=dk2ππk(Λ)g~k(q),|q|>Λ.0=\int\frac{dk}{2\pi}\phi^{(\Lambda)}_{k}\tilde{g}_{k}(q)=\int\frac{dk}{2\pi}\pi^{(\Lambda)}_{k}\tilde{g}_{k}(q),\hskip 21.68121pt|q|>\Lambda. (4.14)

Here we see the difference between our approach and the traditional mode matching [1] or energy cut-off [6] regularization schemes. In the traditional approach, one limits the kk integration to include either the same number of modes as in the pp integration or else an integral out to the same energy. Here instead we integrate over all kk but with a constraint on the coefficients ϕk(Λ)\phi^{(\Lambda)}_{k} which leads these coefficients to be small555The ϕk(Λ)\phi^{(\Lambda)}_{k} are operators. They are small at large |k||k| in the sense that they are superpositions of ϕp\phi_{p} with small coefficients. at k>Λk>\Lambda but nonzero. We claim that this is the correct way to regularize the kink Hamiltonian with a UV energy cutoff because in this approach the regularized kink Hamiltonian is related by a similarity transformation to the vacuum Hamiltonian, which defines the theory. In particular they have the same spectrum, and thus the kink mass, which is the difference between two eigenvalues, can be calculated using one eigenvalue from each Hamiltonian.

The decompositions lead to the Bogoliubov transformations

ϕp=dk2πϕk(Λ)g~k(p),πp=dk2ππk(Λ)g~k(p).\phi_{p}=\int\frac{dk}{2\pi}\phi^{(\Lambda)}_{k}\tilde{g}_{k}(p),\hskip 21.68121pt\pi_{p}=\int\frac{dk}{2\pi}\pi^{(\Lambda)}_{k}\tilde{g}_{k}(p). (4.15)

Integrating (4.12) over xx with weight666If kk is a discrete mode, gk=(1)Pgkg_{-k}=(-1)^{P}g_{k} where PP is the parity of the discrete mode kk. gk(x)g_{-k}(x) one finds the inverse transformations

ϕk(Λ)=ΛΛdp2πϕpg~k(p),πk(Λ)=ΛΛdp2ππpg~k(p).\phi^{(\Lambda)}_{k}=\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\phi_{p}\tilde{g}_{-k}(-p),\hskip 21.68121pt\pi^{(\Lambda)}_{k}=\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\pi_{p}\tilde{g}_{-k}(-p). (4.16)

4.2 Finding the One-Loop Hamiltonian

H2H^{\prime}_{2} consists of three terms

H2=A+B+C.H^{\prime}_{2}=A+B+C. (4.17)

First consider the potential term

A\displaystyle A =\displaystyle= 𝑑xV′′[gf(Λ)(x)]:ϕ(Λ)2(x):a2\displaystyle\int dx\frac{V^{\prime\prime}[gf^{(\Lambda)}(x)]:\phi^{{(\Lambda)}2}(x):_{a}}{2}
=\displaystyle= 𝑑x𝑑yV′′[gf(Λ)(x)]δ(xy):ϕ(Λ)(x)ϕ(Λ)(y):a2\displaystyle\int dx\int dy\frac{V^{\prime\prime}[gf^{(\Lambda)}(x)]\delta(x-y):\phi^{{(\Lambda)}}(x)\phi^{(\Lambda)}(y):_{a}}{2}
=\displaystyle= 𝑑x𝑑ydk2πV′′[gf(Λ)(x)]gk(x)gk(y):ϕ(Λ)(x)ϕ(Λ)(y):a2\displaystyle\int dx\int dy\int\frac{dk}{2\pi}\frac{V^{\prime\prime}[gf^{(\Lambda)}(x)]g_{k}(x)g^{*}_{k}(y):\phi^{{(\Lambda)}}(x)\phi^{(\Lambda)}(y):_{a}}{2}
=\displaystyle= 𝑑x𝑑ydk2π[(ωk2+x2)gk(x)]gk(y):ϕ(Λ)(x)ϕ(Λ)(y):a2.\displaystyle\int dx\int dy\int\frac{dk}{2\pi}\frac{\left[(\omega_{k}^{2}+\partial_{x}^{2})g_{k}(x)\right]g^{*}_{k}(y):\phi^{{(\Lambda)}}(x)\phi^{(\Lambda)}(y):_{a}}{2}.

The derivative term cancels

B\displaystyle B =\displaystyle= 𝑑x:ϕ(Λ)(x)x2ϕ(Λ)(x):a2\displaystyle-\int dx\frac{:{\phi}^{(\Lambda)}(x)\partial^{2}_{x}{\phi}^{(\Lambda)}(x):_{a}}{2}
=\displaystyle= 𝑑x𝑑ydk2πgk(x)gk(y):ϕ(Λ)(y)x2ϕ(Λ)(x):a2\displaystyle-\int dx\int dy\int\frac{dk}{2\pi}\frac{g_{k}(x)g^{*}_{k}(y):{\phi}^{(\Lambda)}(y)\partial^{2}_{x}{\phi}^{(\Lambda)}(x):_{a}}{2}

after integrating xx by parts twice, leaving

A+B\displaystyle A+B =\displaystyle= 𝑑x𝑑ydk2πωk2gk(x)gk(y):ϕ(Λ)(x)ϕ(Λ)(y):a2\displaystyle\int dx\int dy\int\frac{dk}{2\pi}\frac{\omega_{k}^{2}g_{k}(x)g^{*}_{k}(y):\phi^{{(\Lambda)}}(x)\phi^{(\Lambda)}(y):_{a}}{2}
=\displaystyle= dk2πΛΛdp12πΛΛdp22πωk2g~k(p1)g~k(p2):ϕp1ϕp2:a2=D+E\displaystyle\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\frac{\omega_{k}^{2}\tilde{g}_{k}(-p_{1})\tilde{g}^{*}_{k}(p_{2}):\phi_{p_{1}}\phi_{p_{2}}:_{a}}{2}=D+E

where

D\displaystyle D =\displaystyle= dk2πΛΛdp12πΛΛdp22πωk2g~k(p1)g~k(p2)ϕp1ϕp22\displaystyle\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\frac{\omega_{k}^{2}\tilde{g}_{k}(-p_{1})\tilde{g}^{*}_{k}(p_{2})\phi_{p_{1}}\phi_{p_{2}}}{2}
=\displaystyle= dk2πωk2ϕk(Λ)ϕk(Λ)2.\displaystyle\int\frac{dk}{2\pi}\omega_{k}^{2}\frac{\phi^{(\Lambda)}_{k}\phi^{(\Lambda)}_{-k}}{2}.

Here in the case of a discrete mode kk with parity PP, ϕk=(1)Pϕk\phi_{-k}=(-1)^{P}\phi_{k}. The contraction term is

E\displaystyle E =\displaystyle= dk2πΛΛdp12πΛΛdp22πωk2g~k(p1)g~k(p2)22πδ(p1+p2)2ωp1\displaystyle-\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\frac{\omega_{k}^{2}\tilde{g}_{k}(-p_{1})\tilde{g}^{*}_{k}(p_{2})}{2}\frac{2\pi\delta(p_{1}+p_{2})}{2\omega_{p_{1}}}
=\displaystyle= dk2πΛΛdp2πg~k(p)g~k(p)ωk24ωp.\displaystyle-\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\tilde{g}_{k}(p)\tilde{g}^{*}_{k}(p)\frac{\omega_{k}^{2}}{4\omega_{p}}.

The last term in H2H^{\prime}_{2} is

C\displaystyle C =\displaystyle= 𝑑x:π(Λ)2(x):a2=𝑑x𝑑yδ(xy):π(Λ)(x)π(Λ)(y):a2\displaystyle\int dx\frac{:\pi^{{(\Lambda)}2}(x):_{a}}{2}=\int dx\int dy\delta(x-y)\frac{:\pi^{{(\Lambda)}}(x)\pi^{(\Lambda)}(y):_{a}}{2}
=\displaystyle= 𝑑x𝑑ydk2πgk(x)gk(y):π(Λ)(x)π(Λ)(y):a2\displaystyle\int dx\int dy\int\frac{dk}{2\pi}\frac{g_{k}(x)g^{*}_{k}(y):\pi^{{(\Lambda)}}(x)\pi^{(\Lambda)}(y):_{a}}{2}
=\displaystyle= 𝑑x𝑑ydk2πΛΛdp12πΛΛdp22πgk(x)gk(y)eip1xip2y:πp1πp2:a2\displaystyle\int dx\int dy\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\frac{g_{k}(x)g^{*}_{k}(y)e^{-ip_{1}x-ip_{2}y}:\pi_{p_{1}}\pi_{p_{2}}:_{a}}{2}
=\displaystyle= dk2πΛΛdp12πΛΛdp22πg~k(p1)g~k(p2):πp1πp2:a2=F+G\displaystyle\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\frac{\tilde{g}_{k}(-p_{1})\tilde{g}^{*}_{k}(p_{2}):\pi_{p_{1}}\pi_{p_{2}}:_{a}}{2}=F+G

where

F=dk2πΛΛdp12πΛΛdp22πg~k(p1)g~k(p2)πp1πp22=dk2ππk(Λ)πk(Λ)2F=\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\frac{\tilde{g}_{k}(-p_{1})\tilde{g}^{*}_{k}(p_{2})\pi_{p_{1}}\pi_{p_{2}}}{2}=\int\frac{dk}{2\pi}\frac{\pi^{(\Lambda)}_{k}\pi^{(\Lambda)}_{-k}}{2} (4.24)

and

G\displaystyle G =\displaystyle= dk2πΛΛdp12πΛΛdp22πg~k(p1)g~k(p2)ωp22πδ(p1+p2)4\displaystyle-\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\frac{\tilde{g}_{k}(-p_{1})\tilde{g}^{*}_{k}(p_{2})\omega_{p_{2}}2\pi\delta(p_{1}+p_{2})}{4} (4.25)
=\displaystyle= dk2πΛΛdp2πg~k(p)g~k(p)ωp4.\displaystyle-\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\tilde{g}_{k}(p)\tilde{g}^{*}_{k}(p)\frac{\omega_{p}}{4}.

In all, we see that H2H^{\prime}_{2} is the sum of a scalar E+GE+G plus an operator

D+F=12dk2π[πk(Λ)πk(Λ)+ωk2ϕk(Λ)ϕk(Λ)].D+F=\frac{1}{2}\int\frac{dk}{2\pi}\left[\pi^{(\Lambda)}_{k}\pi^{(\Lambda)}_{-k}+\omega_{k}^{2}\phi^{(\Lambda)}_{k}\phi^{(\Lambda)}_{-k}\right]. (4.26)

This looks like an infinite sum of quantum harmonic oscillators, however ϕk(Λ)\phi^{(\Lambda)}_{k} and πk(Λ)\pi^{(\Lambda)}_{k} are not quite canonical variables in the regulated theory as

iβk1k2=[ϕk1(Λ),πk2(Λ)]\displaystyle i\beta_{k_{1}k_{2}}=[\phi^{(\Lambda)}_{k_{1}},\pi^{(\Lambda)}_{k_{2}}] =\displaystyle= ΛΛdp12πg~k1(p1)ΛΛdp22πg~k2(p2)[ϕp1,πp2]\displaystyle\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\tilde{g}_{-{k_{1}}}(-p_{1})\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\tilde{g}_{-k_{2}}(-p_{2})[\phi_{p_{1}},\pi_{p_{2}}]
=\displaystyle= iΛΛdp2πg~k1(p)g~k2(p)=i𝑑x𝑑yΛΛdp2πgk1(x)gk2(y)eip(xy)\displaystyle i\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\tilde{g}_{-{k_{1}}}(p)\tilde{g}_{-k_{2}}(-p)=i\int dx\int dy\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}g_{-{k_{1}}}(x)g_{-k_{2}}(y)e^{ip(x-y)}
=\displaystyle= i𝑑x𝑑ygk1(x)gk2(y)α(xy).\displaystyle i\int dx\int dyg_{-{k_{1}}}(x)g_{-k_{2}}(y)\alpha(x-y).

Nonetheless let us try to solve it as if it were a sum of harmonic oscillators, by defining

Bk=ϕk(Λ)2iπk(Λ)2ωk,Bk=ωkϕk(Λ)+iπk(Λ)\displaystyle B_{k}^{\ddagger}=\frac{\phi^{(\Lambda)}_{k}}{2}-i\frac{\pi^{(\Lambda)}_{k}}{2\omega_{k}},\hskip 21.68121ptB_{-k}=\omega_{k}\phi^{(\Lambda)}_{k}+i\pi^{(\Lambda)}_{k} (4.28)

so that

ωkBkBk=ωk2ϕk(Λ)ϕk(Λ)+πk(Λ)πk(Λ)2ωk2ΛΛdp2πg~k(p)g~k(p).\omega_{k}B_{k}^{\ddagger}B_{k}=\frac{\omega^{2}_{k}\phi^{(\Lambda)}_{k}\phi^{(\Lambda)}_{-k}+\pi^{(\Lambda)}_{k}\pi^{(\Lambda)}_{-k}}{2}-\frac{\omega_{k}}{2}\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\tilde{g}_{{-k}}(p)\tilde{g}_{k}(-p). (4.29)

Then we can rewrite the operator part of H2H^{\prime}_{2} as

D+F=I+J,I=dk2πωkBkBk,J=dk2πΛΛdp2πg~k(p)g~k(p)ωk2D+F=I+J,\hskip 21.68121ptI=\int\frac{dk}{2\pi}\omega_{k}B_{k}^{\ddagger}B_{k},\hskip 21.68121ptJ=\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\tilde{g}^{*}_{{k}}(-p)\tilde{g}_{k}(-p)\frac{\omega_{k}}{2} (4.30)

The operators BB and BB^{\ddagger} operators automatically solve analogous constraints to (4.14)

0=dk2πBkg~k(q)=dk2πBkg~k(q),|q|>Λ.0=\int\frac{dk}{2\pi}B^{\ddagger}_{k}\tilde{g}_{k}(q)=\int\frac{dk}{2\pi}B_{k}\tilde{g}_{k}(q),\hskip 21.68121pt|q|>\Lambda. (4.31)

Summarizing, we may write H2H^{\prime}_{2} as the sum of a scalar E+G+JE+G+J plus a term II which is of the quantum harmonic oscillator form

H2=dk2πΛΛdp2πg~k(p)g~k(p)(ωpωk)24ωp+dk2πωkBkBk.H^{\prime}_{2}=-\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\tilde{g}_{k}(p)\tilde{g}^{*}_{k}(p)\frac{\left(\omega_{p}-\omega_{k}\right)^{2}}{4\omega_{p}}+\int\frac{dk}{2\pi}\omega_{k}B_{k}^{\ddagger}B_{k}. (4.32)

We have argued that at one loop this Hamiltonian completely characterizes the kink sector. In Appendix B we argue that the naive IR divergence in the cc-number term at k=pk=p vanishes.

If we now set the regulator to infinity, then substituting (2.9) into (4.2) one finds the standard canonical commutation relations

Λlim[ϕk1(Λ),πk2(Λ)]=i𝑑xgk1(x)gk2(x)=iδ(k1+k2).\stackrel{{\scriptstyle\rm{lim}}}{{{}_{\Lambda\rightarrow\infty}}}[\phi^{(\Lambda)}_{k_{1}},\pi^{(\Lambda)}_{k_{2}}]=i\int dxg_{-k_{1}}(x)g_{-k_{2}}(x)=i\delta(k_{1}+k_{2}). (4.33)

We can then easily read off the exact spectrum of the regularized Hamiltonian at one loop. The leading order kink ground state |00|0\rangle_{0} is the state annihilated by all BkB_{k} and it has mass given by the first term in (4.32), which, when Λ\Lambda\rightarrow\infty indeed agrees with the formula in Refs. [11, 17, 18].

Using

[Bk1,Bk2]\displaystyle[B_{k_{1}},B_{k_{2}}] =\displaystyle= (ωk2ωk1)βk1k2,[Bk1,Bk2]=ωk2ωk14ωk1ωk2βk1k2\displaystyle(\omega_{k_{2}}-\omega_{k_{1}})\beta_{-k_{1}-k_{2}},\hskip 21.68121pt[B^{\ddagger}_{k_{1}},B^{\ddagger}_{k_{2}}]=\frac{\omega_{k_{2}}-\omega_{k_{1}}}{4\omega_{k_{1}}\omega_{k_{2}}}\beta_{k_{1}k_{2}}
[Bk1,Bk2]\displaystyle{[B_{k_{1}},B^{\ddagger}_{k_{2}}]} =\displaystyle= ωk1+ωk22ωk2βk1k2\displaystyle\frac{\omega_{k_{1}}+\omega_{k_{2}}}{2\omega_{k_{2}}}\beta_{-k_{1}k_{2}} (4.34)

one finds when Λ\Lambda\rightarrow\infty, that the excited states can be obtained by acting with BkB^{\ddagger}_{k}, each of which increases the energy by

ΔE=ωk.\Delta E=\omega_{k}. (4.35)

Our compact notation, in which integrals over kk included sums over discrete states, hid the role of the zero mode. In the case of the zero mode, ωB=0\omega_{B}=0 and so the definition of BBB^{\ddagger}_{B} is singular. However in that case the oscillator (4.26) has no ϕ2\phi^{2} term and so its contribution to H2H^{\prime}_{2} is simply the nonrelativistic center of mass kinetic energy πB2/2\pi^{2}_{B}/2. Thus the spectrum therefore also includes various center of mass momenta for the kink. This, together wth the shape and the continuum quantum harmonic oscillator spectrum, yield the known one-loop spectrum of Ref. [1].

4.3 One-Loop Kink Ground State of the Regularized Hamiltonian

The nondiagonal nature of (4.34) means that the normal mode basis does not diagonalize H2H^{\prime}_{2} when Λ\Lambda is finite. Nevertheless, we will show that the ground state is still annihilated by all BkB_{k} and hence, even at finite Λ\Lambda, the one-loop kink mass is given by the first term in (4.32).

Let us go back a few steps to (4.2) and (4.24)

D+F=dk2πΛΛdp12πΛΛdp22πg~k(p1)g~k(p2)2(πp1πp2+ωk2ϕp1ϕp2).D+F=\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\frac{\tilde{g}_{k}(-p_{1})\tilde{g}^{*}_{k}(p_{2})}{2}\left(\pi_{p_{1}}\pi_{p_{2}}+\omega_{k}^{2}\phi_{p_{1}}\phi_{p_{2}}\right). (4.36)

Define the eigenstates of the operators ϕp\phi_{p} as

ϕp|ψ=ψp|ψ\phi_{p}|\psi\rangle=\psi_{p}|\psi\rangle (4.37)

where ψ\psi is a real-valued function on the interval [Λ,Λ][-\Lambda,\Lambda]. Let us write the arbitrary state |Ψ|\Psi\rangle in the Schrodinger representation

|Ψ=𝒟ψΨ[ψ]|ψ|\Psi\rangle=\int\mathcal{D}\psi\Psi[\psi]|\psi\rangle (4.38)

where the integral is over all functions ψ\psi and Ψ[ψ]\Psi[\psi] is the Schrodinger wave functional. As the eigenstates |ψ|\psi\rangle are a complete basis, the state |Ψ|\Psi\rangle is arbitrary. It follows that

πp|ψ=2πi𝒟ψδΨ[ψ]δψp|ψ.\pi_{p}|\psi\rangle=-2\pi i\int\mathcal{D}\psi\frac{\delta\Psi[\psi]}{\delta\psi_{-p}}|\psi\rangle~{}. (4.39)

Inserting the Ansatz

Ψ[ψ]=exp(12ΛΛdp12πΛΛdp22πAp1p2ψp1ψp2)\Psi[\psi]=\hbox{\rm exp}\left(-\frac{1}{2}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}A_{p_{1}p_{2}}\psi_{p_{1}}\psi_{p_{2}}\right) (4.40)

into (4.36) one finds an expression of the form

(D+F)|Ψ=(+ΛΛdq12πΛΛdq22πBq1q2ψq1ψq2)|Ψ(D+F)|\Psi\rangle=\left(\mathcal{E}+\int^{\Lambda}_{-\Lambda}\frac{dq_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dq_{2}}{2\pi}B_{q_{1}q_{2}}\psi_{q_{1}}\psi_{q_{2}}\right)|\Psi\rangle (4.41)

where

Bq1q2\displaystyle B_{q_{1}q_{2}} =\displaystyle= 12ΛΛdp12πΛΛdp22πdk2πg~k(p1)g~k(p2)[Ap1,q1Ap2,q2+ωk2δ(p1q1)δ(p2q2)]\displaystyle\frac{1}{2}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\int\frac{dk}{2\pi}\tilde{g}_{k}(p_{1})\tilde{g}_{-k}(p_{2})\left[-A_{p_{1},-q_{1}}A_{p_{2},-q_{2}}+\omega_{k}^{2}\delta(p_{1}-q_{1})\delta(p_{2}-q_{2})\right]
=\displaystyle= 12ΛΛdp12πΛΛdp22πdk2πg~k(p1)g~k(p2)Ap1,q1Ap2,q2+12dk2πg~k(q1)g~k(q2)ωk2\displaystyle-\frac{1}{2}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\int\frac{dk}{2\pi}\tilde{g}_{k}(p_{1})\tilde{g}_{-k}(p_{2})A_{p_{1},-q_{1}}A_{p_{2},-q_{2}}+\frac{1}{2}\int\frac{dk}{2\pi}\tilde{g}_{k}(q_{1})\tilde{g}_{-k}(q_{2})\omega_{k}^{2}
\displaystyle\mathcal{E} =\displaystyle= 12ΛΛdp12πΛΛdp22πdk2πg~k(p1)g~k(p2)Ap1,p2.\displaystyle\frac{1}{2}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\int\frac{dk}{2\pi}\tilde{g}_{k}(p_{1})\tilde{g}_{-k}(p_{2})A_{-p_{1},-p_{2}}~{}. (4.42)

Our strategy will be to find the Ap1p2A_{p_{1}p_{2}} such that Bq1q2=0B_{q_{1}q_{2}}=0 and use it to find \mathcal{E}. A sufficient condition for BB to vanish is

ΛΛdp2πg~k(p)Ap,q=ωkg~k(q)\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\tilde{g}_{k}(p)A_{p,-q}=\omega_{k}\tilde{g}_{k}(q) (4.43)

for all kk and qq. Multiplying by g~k(r)\tilde{g}_{-k}(-r) and integrating over kk, using completeness of gg, one finds

dk2πg~k(r)ΛΛdp2πg~k(p)Ap,q=Ar,q=dk2πg~k(r)ωkg~k(q).\int\frac{dk}{2\pi}\tilde{g}_{-k}(-r)\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\tilde{g}_{k}(p)A_{p,-q}=A_{r,-q}=\int\frac{dk}{2\pi}\tilde{g}_{-k}(-r)\omega_{k}\tilde{g}_{k}(q). (4.44)

We claim that this ApqA_{pq}, inserted into the Ansatz (4.40), is the kink ground state, or more precisely 𝒟f|K\mathcal{D}_{f}^{\dagger}|K\rangle. To see this, note that it is annihilated by all BkB_{k}:

Bk|Ψ\displaystyle B_{-k}|\Psi\rangle =\displaystyle= ΛΛdp2πg~k(p)(ωkϕp+iπp)|Ψ\displaystyle\int_{-\Lambda}^{\Lambda}\frac{dp}{2\pi}\tilde{g}_{-k}(-p)(\omega_{k}\phi_{p}+i\pi_{p})|\Psi\rangle (4.45)
=\displaystyle= ΛΛdp2πg~k(p)(ωkψp2πΛΛdq2πAp,qψq)|Ψ\displaystyle\int_{-\Lambda}^{\Lambda}\frac{dp}{2\pi}\tilde{g}_{-k}(-p)\left(\omega_{k}\psi_{p}-2\pi\int_{-\Lambda}^{\Lambda}\frac{dq}{2\pi}A_{-p,q}\psi_{q}\right)|\Psi\rangle (4.46)
=\displaystyle= 0,\displaystyle 0, (4.47)

where the last step follows from (4.43). Hence |Ψ|\Psi\rangle is annihilated by the last term of (4.32) and since this term is a positive operator |Ψ|\Psi\rangle must be the lowest energy state. As a check, note that the corresponding energy \mathcal{E} is

\displaystyle\mathcal{E} =\displaystyle= 12ΛΛdp12πΛΛdp22πd2k(2π)2g~k1(p1)g~k1(p2)g~k2(p1)ωk2g~k2(p2)\displaystyle-\frac{1}{2}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\int\frac{d^{2}k}{(2\pi)^{2}}\tilde{g}_{k_{1}}(p_{1})\tilde{g}_{-k_{1}}(p_{2})\tilde{g}_{-k_{2}}(p_{1})\omega_{k_{2}}\tilde{g}_{k_{2}}(p_{2}) (4.48)
=\displaystyle= 12ΛΛdp12πΛΛdp22πdk22π2πδ(p1+p2)g~k2(p1)ωk2g~k2(p2)\displaystyle\frac{1}{2}\int^{\Lambda}_{-\Lambda}\frac{dp_{1}}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp_{2}}{2\pi}\int\frac{dk_{2}}{2\pi}2\pi\delta(p_{1}+p_{2})\tilde{g}_{-k_{2}}(p_{1})\omega_{k_{2}}\tilde{g}_{k_{2}}(p_{2})
=\displaystyle= 12ΛΛdp2πdk2πωkg~k(p)g~k(p),\displaystyle\frac{1}{2}\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\int\frac{dk}{2\pi}\omega_{k}\tilde{g}_{-k}(-p)\tilde{g}_{k}(p),

which matches the JJ from Eq. (4.30).

Therefore, even in the regularized theory, JJ is the energy contribution from D+FD+F, or more precisely the eigenvalue of the operator D+FD+F acting on the kink ground state. Including the scalar terms in H2H^{\prime}_{2}, one finds that the total ground state energy of the regularized kink is

E+G+J=dk2πΛΛdp2πg~k(p)g~k(p)(ωpωk)24ωp.E+G+J=-\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\tilde{g}_{k}(p)\tilde{g}^{*}_{k}(p)\frac{\left(\omega_{p}-\omega_{k}\right)^{2}}{4\omega_{p}}. (4.49)

This is the ground state energy for the unregularized kink found in Ref. [11] using mode matching, but now with the pp integral cut off. Therefore the limit Λ\Lambda\rightarrow\infty agrees with the known result.

5 Remarks

In this paper we have described a quantum model which is regularized by a cutoff from the beginning. It exhibits a quantum kink and we found, at one-loop, its ground state and mass. The results were almost trivial generalizations of the corresponding results in the unregularized theory, in which one merely restricts the domain of integration of the momentum pp to the interval [Λ,Λ][-\Lambda,\Lambda]. This is in part because at one-loop, the theory is free, although it is diagonal in the normal mode basis kk and not the momentum pp basis. However this was also suggested by the fact that two-dimensional scalar models can be rendered finite by normal ordering without ever regularizing, and so all quantities may be computed without recourse to regularization [2, 17]. Thus one may already suspect that no sign of regularization may remain when the regulator is taken to infinity, as it could have been avoided from the beginning.

In more interesting models, with fermions or more dimensions, normal ordering is not sufficient to remove all divergences. Thus it is possible that the limit in which the regulator is taken to infinity will leave some nonzero residue, indeed in the case of supersymmetric kinks one may expect to arrive at the contribution from the one-loop anomaly [24].

In the supersymmetric case one may hope that sufficient supersymmetry will allow a nonperturbative approach. However we saw here that the kink solution itself has corrections of order eΛ/me^{-\Lambda/m}. In perturbation theory, we expect this to be always multiplied by finite powers of Λ\Lambda and so such contributions will vanish in the Λ\Lambda\rightarrow\infty limit, but in the nonperturbative regime, which is the relevant regime for applications to paradigms [25, 26, 27] of QCD confinement, it is possible that these corrections will be physically relevant if not dominant.

Appendix A Finding the Shift: The ϕ4\phi^{4} Double Well

In this appendix we evaluate the leading correction to f(Λ)(x)f^{(\Lambda)}(x) in the case of the ϕ4\phi^{4} double well. We show that it is exponentially small in Λ/m\Lambda/m but nonzero. Here we consider the ϕ4\phi^{4} double well without shifting ϕ~(Λ)\tilde{\phi}^{(\Lambda)} by vv as in Eq. (2.3). Recall that, in momentum space, this corresponds a shift of the transformed field by 2πvδ(p)2\pi v\delta(p) and in particular it only affects the momentum space field at p=0p=0.

In this model the definition (3.14) is

V~[gf]=\displaystyle\tilde{V}[gf]= 14(g2f2m2/2)2\displaystyle~{}\frac{1}{4}\left(g^{2}f^{2}-m^{2}/2\right)^{2}
V~[gf]/g=\displaystyle\tilde{V}^{\prime}[gf]/g= g2f3m2f/2\displaystyle~{}g^{2}f^{3}-m^{2}f/2
0=\displaystyle 0= 𝑑x[x2f(Λ)(x)m2f(Λ)(x)/2+g2f(Λ)3(x)]eipx,|p|Λ.\displaystyle~{}\int dx\left[-\partial^{2}_{x}f^{{(\Lambda)}}(x)-m^{2}f^{(\Lambda)}(x)/2+g^{2}f^{{(\Lambda)}3}(x)\right]e^{ipx},\qquad|p|\leq\Lambda. (A.1)

The basic approach is to insert the finite-Λ\Lambda Fourier transform

f(Λ)(x)=\displaystyle f^{(\Lambda)}(x)= ΛΛdq2πf~(q)eiqx,\displaystyle~{}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}\frac{dq}{2\pi}\tilde{f}(q)e^{-iqx}~{}, (A.2)

into (A.1). Our Ansatz for f~\tilde{f} is the series expansion

f~(q)=n=0f~n(q),\tilde{f}(q)=\sum_{n=0}\tilde{f}_{n}(q)~{}, (A.3)

where f~0(q)\tilde{f}_{0}(q) is just the Fourier transform of the classical solution f(x)f(x) and f~n1(q)\tilde{f}_{n\geq 1}(q) is bounded, for all qq, by a polynomial in Λ\Lambda times eπnΛ/me^{-\pi n\Lambda/m}. We will then solve for the f~n(q)\tilde{f}_{n}(q) perturbatively, and along the way will see that our Ansatz is consistent. This will imply the main result of this appendix, that f(x)f(x) can be expanded in a power series with the leading Λ\Lambda dependence suppressed by order O(eπΛ/m)O(e^{-\pi\Lambda/m}).

There are, however, two technical points that require explanation before proceeding. First, we must define the \mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int symbol in (A.2). Ordinary integration, for f(x)f(x) a classical kink solution, would be ill-defined at q=0q=0. We are free to define the Fourier transform f~\tilde{f} as we like, so long as we are able to use it to demonstrate the main result written above. Therefore, we choose \mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int to be a principal value integral, defined by

ΛΛ:=limϵ0+(Λϵ+ϵΛ).\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}:=\lim_{\epsilon\to 0_{+}}\left(\int_{-\Lambda}^{-\epsilon}+\int_{\epsilon}^{\Lambda}\right)~{}. (A.4)

This type of integral can be used to obtain the standard kink profile f()(x)=m2gtanh(mx/2)f^{(\infty)}(x)=\frac{m}{2g}\textrm{tanh}(mx/2) from its Fourier transform f~0(q)=2πigcsch(πq/m)\tilde{f}_{0}(q)=\frac{\sqrt{2}\pi i}{g}\textrm{csch}(\pi q/m):

m2gtanh(mx/2)=dq2π(2πigcsch(πq/m))eiqx,\frac{m}{2g}\textrm{tanh}(mx/2)=\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\infty}^{\infty}\frac{dq}{2\pi}\left(\frac{\sqrt{2}\pi i}{g}\textrm{csch}(\pi q/m)\right)e^{-iqx}~{}, (A.5)

The integral on the right would be undefined without the principle value prescription. More generally, as a distribution the Fourier transform [f()]\mathcal{F}[f^{(\infty)}] acts on any test function b(q)b(q) via integration against f~0(q)\tilde{f}_{0}(q) with the principle value prescription. Indeed, the proper setting for the kink profile and its Fourier transform is to view both as tempered distributions. The kink profile, f()f^{(\infty)}, is a nonsingular distribution, meaning that it is defined globally by a smooth function. The Fourier transform f~0=[f()]\tilde{f}_{0}=\mathcal{F}[f^{(\infty)}] is singular, meaning that it can only be represented locally by a smooth function (the csch above); its definition as a distribution contains more information – namely the principal value prescription. Although we do not a priori know the explicit finite-Λ\Lambda analogs, f(Λ)f^{(\Lambda)} and f~\tilde{f}, we we expect they have the same large-|x||x| asymptotics and the same singularity structure at q=0q=0 as the leading order configurations f(),f~0f^{(\infty)},\tilde{f}_{0}, respectively, hence the appearance of the principal value integral in (A.2).

A second technical point is the following. The standard result for smooth functions that the Fourier transform of the pointwise product is the convolution of Fourier transforms,

[fg]=12π[f][g],\mathcal{F}[f\cdot g]=\frac{1}{2\pi}\mathcal{F}[f]\star\mathcal{F}[g]~{}, (A.6)

also holds for tempered distributions,777It holds for distributions when the product is defined. In general this involves a condition on the wave front sets of the distributions [28]. In the case of f()f^{(\infty)}, which is defined globally by integration against a smooth function, the wave front set is trivial. The product of this distribution with any other temprered distribution is defined and the convolution theorem holds. where the factor of 2π2\pi is due to our Fourier transform conventions. Here, if f,gf,g are ordinary functions then (fg)(x)=f(x)g(x)(f\cdot g)(x)=f(x)g(x) denotes the pointwise product and (fg)(x)=𝑑yf(y)g(xy)(f\star g)(x)=\int dyf(y)g(x-y) denotes the convolution. The convolution of two distributions is defined through the convolution of a distribution with a smooth test function as follows. If ff is a distribution defined locally by f(x)f(x) and bb a smooth function with compact support, then the convolution fbf\star b is a smooth function with value (fb)(x)=𝑑yf(y)b(xy)(f\star b)(x)=\int dyf(y)b(x-y). The convolution of two distributions f,g,f,g, is then the unique distribution such that (fg)b=f(gb)(f\star g)\star b=f\star(g\star b). This definition and the corresponding convolution theorem can be extended to

[f1fn]=1(2π)n1f~1f~n,\mathcal{F}[f_{1}\cdot\ldots\cdot f_{n}]=\frac{1}{(2\pi)^{n-1}}\tilde{f}_{1}\star\ldots\star\tilde{f}_{n}~{}, (A.7)

where f~j=[fj]\tilde{f}_{j}=\mathcal{F}[f_{j}] and the distribution on the right is defined through sequential action on a test function.

An instructive example is to show that (A.6) holds for the kink profile f=g=f()f=g=f^{(\infty)} and its Fourier transform. Using tanh2=1sech2\textrm{tanh}^{2}=1-\textrm{sech}^{2}, one finds that [f()f()]\mathcal{F}[f^{(\infty)}\cdot f^{(\infty)}] has

[f()f()](q)=πm2g2δ(q)2πqg2sinh(πq/m).\mathcal{F}[f^{(\infty)}\cdot f^{(\infty)}](q)=\frac{\pi m^{2}}{g^{2}}\delta(q)-\frac{2\pi q}{g^{2}\textrm{sinh}(\pi q/m)}~{}. (A.8)

Below, we will recover this result in the Λ\Lambda\to\infty limit of a finite-Λ\Lambda computation of the convolution, f~0f~0\tilde{f}_{0}\star\tilde{f}_{0}.

We can think of f(Λ)f^{(\Lambda)} in (A.2) as the ordinary inverse Fourier transform of f~\tilde{f} if we define f~\tilde{f} to vanish outside of [Λ,Λ][-\Lambda,\Lambda]. For example, if f~0\tilde{f}_{0}, initially defined on all of \mathbb{R}, is set to zero outside [Λ,Λ][-\Lambda,\Lambda], the resulting distribution is still a tempered distribution (now with compact support). Henceforth we will write f~(Λ)\tilde{f}^{(\Lambda)} for distributions with support on [Λ,Λ][-\Lambda,\Lambda]. In particular, for any test function bb we have

ΛΛdq2πf~0(q)b(q)=dq2πf~0(Λ)(q)b(q).\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}\frac{dq}{2\pi}\tilde{f}_{0}(q)b(q)=\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\infty}^{\infty}\frac{dq}{2\pi}\tilde{f}_{0}^{(\Lambda)}(q)b(q)~{}. (A.9)

The convolution theorem continues to hold for such distributions, so that for powers of f(Λ)f^{(\Lambda)} we have

[(f(Λ))n]=\displaystyle\mathcal{F}[(f^{(\Lambda)})^{n}]= 1(2π)n1f~(Λ)f~(Λ).\displaystyle~{}\frac{1}{(2\pi)^{n-1}}\tilde{f}^{(\Lambda)}\star\ldots\star\tilde{f}^{(\Lambda)}~{}. (A.10)

Note this means that the support of the nn-\star convolution must be [nΛ,nΛ][-n\Lambda,n\Lambda] such that

nΛnΛ(f~(Λ)f~(Λ))(q)b(pq)=ΛΛ𝑑q1f~(q1)ΛΛ𝑑q2f~(q2)ΛΛ𝑑qnf~(qn)b(pΣiqi).\displaystyle\int_{-n\Lambda}^{n\Lambda}\left(\tilde{f}^{(\Lambda)}\star\ldots\star\tilde{f}^{(\Lambda)}\right)(q)b(p-q)=\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}dq_{1}\tilde{f}(q_{1})\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}dq_{2}\tilde{f}(q_{2})\cdots\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}dq_{n}\tilde{f}(q_{n})b(p-\Sigma_{i}q_{i})~{}. (A.11)

This reflects the fact that the pointwise product of nn f(Λ)f^{(\Lambda)}’s on the left of (A.10) contains momentum modes up to |q|=nΛ|q|=n\Lambda.

With these preliminaries out of the way we can proceed with the perturbative solution of (A.1). Inserting (A.2) into (A.1) gives

0=\displaystyle 0= (p2m22)f~(Λ)(p)+g2(2π)2(f~(Λ)f~(Λ)f~(Λ))(p),|p|Λ.\displaystyle~{}\left(p^{2}-\frac{m^{2}}{2}\right)\tilde{f}^{(\Lambda)}(p)+\frac{g^{2}}{(2\pi)^{2}}\left(\tilde{f}^{(\Lambda)}\star\tilde{f}^{(\Lambda)}\star\tilde{f}^{(\Lambda)}\right)(p)~{},\qquad|p|\leq\Lambda~{}. (A.12)

This is an integral equation for the local function f~(Λ)(p)\tilde{f}^{(\Lambda)}(p) on [Λ,Λ][-\Lambda,\Lambda]. We expect the distribution f~(Λ)\tilde{f}^{(\Lambda)} to be singular at p=0p=0 and defined via the principal value prescription as discussed above. Therefore, we restrict attention in (A.12) to p0p\neq 0.

Inserting the expansion (A.3) into (A.12), and assuming that f~0\tilde{f}_{0} solves the equation when Λ\Lambda\to\infty (which we will check), we find the following linear equation for the first correction, f~1\tilde{f}_{1}:

((p2m22)f~1(Λ)+3g2(2π)2(f~0(Λ)f~0(Λ))f~1(Λ))(p)=F1(p),|p|<Λ,\left(\left(p^{2}-\tfrac{m^{2}}{2}\right)\cdot\tilde{f}_{1}^{(\Lambda)}+\frac{3g^{2}}{(2\pi)^{2}}\left(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\right)\star\tilde{f}_{1}^{(\Lambda)}\right)(p)=F_{1}(p)~{},\qquad|p|<\Lambda~{}, (A.13)

where F1(p)F_{1}(p) is the first finite-Λ\Lambda correction from the non-linear term evaluated on the leading-order solution:

F1(p):=g2(2π)2(f~0(Λ)f~0(Λ)f~0(Λ))(p)|first subleading behavior in m/Λ.F_{1}(p):=-\frac{g^{2}}{(2\pi)^{2}}\left(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\right)(p)\bigg{|}_{\textrm{first subleading behavior in $m/\Lambda$}}~{}. (A.14)

Hence the main tasks now are to compute the finite-Λ\Lambda convolutions of f~0(Λ)\tilde{f}_{0}^{(\Lambda)} with itself appearing in (A.13), (A.14), and to invert the linear operator in (A.13). In the following we describe the approach and present the results but suppress all intermediate steps.888A more detailed presentation of this calculation may appear elsewhere.

The first convolution we need is 12π(f~0(Λ)f~0(Λ))\frac{1}{2\pi}(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}), given by

12π2Λ2Λ𝑑q(f~0(Λ)f~0(Λ))(q)b(pq)=\displaystyle\frac{1}{2\pi}\int_{-2\Lambda}^{2\Lambda}dq\left(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\right)(q)b(p-q)= (A.15)
=πg2ΛΛ𝑑q1csch(πq1/m)ΛΛ𝑑q2csch(πq2/m)b(pq1q2),\displaystyle\qquad\qquad=-\frac{\pi}{g^{2}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}dq_{1}\textrm{csch}(\pi q_{1}/m)\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}dq_{2}\textrm{csch}(\pi q_{2}/m)b(p-q_{1}-q_{2})~{}, (A.16)

for any smooth test function bb. We introduce infinitesimal parameters ϵ1,2\epsilon_{1,2} associated with the principal value integrals over q1,2q_{1,2} respectively, and we can assume ϵ2<ϵ1\epsilon_{2}<\epsilon_{1} since the ϵ20\epsilon_{2}\to 0 limit is to be taken first. The basic idea is to change integration variables in the double integral from (q1,q2)(q_{1},q_{2}) to (q,q2)(q,q_{2}), where q=q1+q2q=q_{1}+q_{2}, and carry out the q2q_{2} integral to extract (f~0(Λ)f~0(Λ))(q)(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)})(q). This, however, is only valid away from q=0q=0, since it involves exchanging the ϵ20\epsilon_{2}\to 0 limit with the integral over q1q_{1}. If q=0q=0 then the csch(πq1/m)\textrm{csch}(\pi q_{1}/m) factor is not smooth at the q2q_{2} pole, and one must take ϵ20\epsilon_{2}\to 0 limit first, before evaluating the q1q_{1} integral. Therefore we divide the integration into two pieces: one over a region Rϵ={(q1,q2)[Λ,Λ]2:|q1+q2|<ϵ}R_{\epsilon}=\{(q_{1},q_{2})\in[-\Lambda,\Lambda]^{2}:|q_{1}+q_{2}|<\epsilon\} and the other over [Λ,Λ]2Rϵ[-\Lambda,\Lambda]^{2}\setminus R_{\epsilon}. We take ϵ>ϵ1+ϵ2\epsilon>\epsilon_{1}+\epsilon_{2} and send ϵ0\epsilon\to 0 at the very end. See Figure 1.

Refer to caption
Figure 1: Integration region for the right-hand side of (A.15). For the integral over the blue region we change variables from (q1,q2)(q_{1},q_{2}) to (q,q2)(q,q_{2}) where q=q1+q2q=q_{1}+q_{2}. The limits of integration on q2q_{2} then become a function of q,Λ,ϵ1,2q,\Lambda,\epsilon_{1,2}. For example, when ϵ<q<Λϵ1\epsilon<q<\Lambda-\epsilon_{1}, the integral over q2q_{2} covers [qΛ,ϵ2][ϵ2,qϵ1][q+ϵ1,Λ][q-\Lambda,-\epsilon_{2}]\cup[\epsilon_{2},q-\epsilon_{1}]\cup[q+\epsilon_{1},\Lambda]. The test function does not depend on q2q_{2} and can be pulled out of the q2q_{2} integral. For all cases ϵ<|q|<2Λ\epsilon<|q|<2\Lambda, the result of the q2q_{2} integration, after taking the limits ϵ20\epsilon_{2}\to 0 followed by ϵ10\epsilon_{1}\to 0, is captured by the second term on the right-hand side of (A.17). For the integral over the pink region we instead expand the test function around q=0q=0 and compute leading term in the resulting integrand. That computation requires first integrating over q2q_{2} for fixed q1q_{1} (where the limits of integration depend on q1,ϵ,ϵ2q_{1},\epsilon,\epsilon_{2}), then taking the ϵ20\epsilon_{2}\to 0 limit, then integrating over q1q_{1}, and finally taking ϵ0\epsilon\to 0.

The integral over RϵR_{\epsilon} gives a result for the right-hand side of (A.15) that is proportional to b(p)b(p) in the limit ϵ0\epsilon\to 0, and hence this corresponds to a delta function contribution to (f~0(Λ)f~0(Λ))(q)(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)})(q). The coefficient can be evaluated and turns out to be Λ\Lambda-independent, and this term matches the δ\delta-function term in (A.8). Meanwhile the integral over [Λ,Λ]2Rϵ[-\Lambda,\Lambda]^{2}\setminus R_{\epsilon} gives a Λ\Lambda-dependent contribution. In total we find

12π(f~0(Λ)f~0(Λ))(q)=\displaystyle\frac{1}{2\pi}\left(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\right)(q)= πm2g2δ(q)2mg2sinh(π|q|/m)ln[sinh(πΛ/m)sinh(π|Λ|q||/m)].\displaystyle~{}\frac{\pi m^{2}}{g^{2}}\delta(q)-\frac{2m}{g^{2}\textrm{sinh}(\pi|q|/m)}\ln\left[\frac{\textrm{sinh}(\pi\Lambda/m)}{\textrm{sinh}(\pi|\Lambda-|q||/m)}\right]~{}. (A.17)

This agrees with the right-hand side of (A.8) in the Λ\Lambda\to\infty limit for fixed qq, as required by the convolution theorem. It has integrable logarithmic singularities at |q|=Λ|q|=\Lambda.

For the double convolution we use

3Λ3Λ𝑑p(f~0(Λ)f~0(Λ)f~0(Λ))(p)b(kp)=\displaystyle\int_{-3\Lambda}^{3\Lambda}dp\left(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\right)(p)b(k-p)= ΛΛ𝑑q1f~0(q1)ΛΛf~0(q2)ΛΛ𝑑q3f~0(q3)b(kΣiqi)\displaystyle~{}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}dq_{1}\tilde{f}_{0}(q_{1})\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}\tilde{f}_{0}(q_{2})\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}dq_{3}\tilde{f}_{0}(q_{3})b(k-\Sigma_{i}q_{i}) (A.18)
=\displaystyle= ΛΛ𝑑q1f~0(q1)2Λ2Λ𝑑q(f~0(Λ)f~0(Λ))(q)b(kq1q).\displaystyle~{}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}dq_{1}\tilde{f}_{0}(q_{1})\int_{-2\Lambda}^{2\Lambda}dq(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)})(q)b(k-q_{1}-q)~{}. (A.19)

We apply (A.15) for (f~0f~0)(q)(\tilde{f}_{0}\star\tilde{f}_{0})(q) and change variables in the remaining double integral from (q1,q)(q_{1},q) to (p,q)(p,q) where p=q1+qp=q_{1}+q. Integrating over qq will then allow us to extract the double convolution (f~0(Λ)f~0(Λ)f~0(Λ))(p)(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)})(p). Note there is no principal value prescription required for the distribution (f~0(Λ)f~0(Λ))(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}) as it has no poles. Thus for all p0p\neq 0 it is possible to obtain a local representation of the double convolution by carrying out the qq integral only. Furthermore, although the support of the double convolution is [3Λ,3Λ][-3\Lambda,3\Lambda], we only require it in the range [Λ,Λ][-\Lambda,\Lambda] for the solution to (A.12). For any pp in this range, the integration over qq covers [pΛ,pϵ1][p+ϵ1,p+Λ][p-\Lambda,p-\epsilon_{1}]\cup[p+\epsilon_{1},p+\Lambda]. Here the outer limits are due to |q1|Λ|q_{1}|\leq\Lambda and the inner limits are due to the principal value prescription on the q1q_{1} integration. Hence, we have that

1(2π)2(f~0(Λ)f~0(Λ)f~(Λ))(p)=\displaystyle\frac{1}{(2\pi)^{2}}\left(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}^{(\Lambda)}\right)(p)= 1(2π)2limϵ10(pΛpϵ1+p+ϵ1p+Λ)dqf~0(pq)(f~0Λf~0)(q)\displaystyle~{}\frac{1}{(2\pi)^{2}}\lim_{\epsilon_{1}\to 0}\left(\int_{p-\Lambda}^{p-\epsilon_{1}}+\int_{p+\epsilon_{1}}^{p+\Lambda}\right)dq\tilde{f}_{0}(p-q)(\tilde{f}_{0}\star_{\Lambda}\tilde{f}_{0})(q)
=\displaystyle= m22g2f~0(p)1g2F(p),|p|<Λ,\displaystyle~{}\frac{m^{2}}{2g^{2}}\tilde{f}_{0}(p)-\frac{1}{g^{2}}F(p)~{},\qquad|p|<\Lambda~{}, (A.20)

where the first term arises from the Dirac delta term in (A.15) and we’ve defined

F(p):=\displaystyle F(p):= (2m)2πi(2π)glimϵ10(pΛpϵ1+p+ϵ1p+Λ)dqcsch(π(pq)/m)csch(π|q|/m)×\displaystyle~{}\frac{(2m)\sqrt{2}\pi i}{(2\pi)g}\lim_{\epsilon_{1}\to 0}\left(\int_{p-\Lambda}^{p-\epsilon_{1}}+\int_{p+\epsilon_{1}}^{p+\Lambda}\right)dq~{}\textrm{csch}(\pi(p-q)/m)\textrm{csch}(\pi|q|/m)\times (A.21)
×ln[sinh(πΛ/m)sinh(π|Λ|q||/m)],\displaystyle~{}\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\times\ln\left[\frac{\textrm{sinh}(\pi\Lambda/m)}{\textrm{sinh}(\pi|\Lambda-|q||/m)}\right]~{}, (A.22)

as giving the contribution from the remaining piece of (A.15). We’ve pulled out an explicit factor of 1/g2-1/g^{2} from the definition of FF so that F1(p)F_{1}(p) in (A.14) is precisely the first subleading term (with respect to the exponential behavior) in the large Λ\Lambda expansion of F(p)F(p).

By changing variables qqq\to-q, one sees that the second integral in (A.21) is the negative of the first with ppp\to-p. Thus F(p)F(p) is an odd function of pp. The integral can be evaluated in closed form in terms of logarithms and dilogarithms. We record the full result here since it is important in demonstrating our claim that the perturbative expansion is an expansion in e2πΛ/me^{-2\pi\Lambda/m}. Setting ε=eπΛ/m\varepsilon=e^{-\pi\Lambda/m} and z=eπp/mz=e^{\pi p/m} for shorthand, we find

F(p)=\displaystyle F(p)= i2m2πgsinh(πp/m){π2p2m2+πpmln[1ε2z21ε2z2]+3πΛmln[(1ε2z2)(1ε2z2)(1ε2)2]+\displaystyle~{}\frac{i\sqrt{2}m^{2}}{\pi g~{}\textrm{sinh}(\pi p/m)}\bigg{\{}\frac{\pi^{2}p^{2}}{m^{2}}+\frac{\pi p}{m}\ln\left[\frac{1-\varepsilon^{2}z^{-2}}{1-\varepsilon^{2}z^{2}}\right]+\frac{3\pi\Lambda}{m}\ln\left[\frac{(1-\varepsilon^{2}z^{2})(1-\varepsilon^{2}z^{-2})}{(1-\varepsilon^{2})^{2}}\right]+ (A.23)
+ln[(zz1)2(1ε2)2]ln[(1ε2)2(1ε2z2)(1ε2z2)]+\displaystyle~{}+\ln\left[\frac{(z-z^{-1})^{2}}{(1-\varepsilon^{2})^{2}}\right]\ln\left[\frac{(1-\varepsilon^{2})^{2}}{(1-\varepsilon^{2}z^{2})(1-\varepsilon^{2}z^{-2})}\right]+ (A.24)
+ln[|z21|1ε2z2]ln[1ε21ε2z2]+ln[|z21|1ε2z2]ln[1ε21ε2z2]+\displaystyle~{}+\ln\left[\frac{|z^{2}-1|}{1-\varepsilon^{2}z^{2}}\right]\ln\left[\frac{1-\varepsilon^{2}}{1-\varepsilon^{2}z^{2}}\right]+\ln\left[\frac{|z^{-2}-1|}{1-\varepsilon^{2}z^{-2}}\right]\ln\left[\frac{1-\varepsilon^{2}}{1-\varepsilon^{2}z^{-2}}\right]+ (A.25)
12li2(ε2z2)12li2(ε2z2)+li2(ε2)li2(1z21ε2)li2(1z21ε2)+\displaystyle~{}-\frac{1}{2}{\operatorname{li}_{2}}(\varepsilon^{2}z^{2})-\frac{1}{2}{\operatorname{li}_{2}}(\varepsilon^{2}z^{-2})+{\operatorname{li}_{2}}(\varepsilon^{2})-{\operatorname{li}_{2}}\left(\frac{1-z^{2}}{1-\varepsilon^{2}}\right)-{\operatorname{li}_{2}}\left(\frac{1-z^{-2}}{1-\varepsilon^{2}}\right)+ (A.26)
+li2(1z21ε2z2)+li2(1z21ε2z2)+li2(ε2(1z2)(1ε2z2))+li2(ε2(1z2)(1ε2z2))},\displaystyle~{}+{\operatorname{li}_{2}}\left(\frac{1-z^{2}}{1-\varepsilon^{2}z^{2}}\right)+{\operatorname{li}_{2}}\left(\frac{1-z^{-2}}{1-\varepsilon^{2}z^{-2}}\right)+{\operatorname{li}_{2}}\left(\frac{\varepsilon^{2}(1-z^{2})}{(1-\varepsilon^{2}z^{2})}\right)+{\operatorname{li}_{2}}\left(\frac{\varepsilon^{2}(1-z^{-2})}{(1-\varepsilon^{2}z^{-2})}\right)\bigg{\}}~{}, (A.27)

and we have checked this result against numerical integration. Noting that li2(x)=x+O(x2){\operatorname{li}_{2}}(x)=x+O(x^{2}) for small xx, it is easy to see that F(p)F(p) has a power series expansion in ε2\varepsilon^{2}, where every term after the very first one in the curly brackets above begins at O(ε2)O(\varepsilon^{2}):

F(p)=p2f~0(p)+n=1Fn(p),F(p)=p^{2}\tilde{f}_{0}(p)+\sum_{n=1}^{\infty}F_{n}(p)~{}, (A.28)

where, for fixed pp, Fn=O(e2πnΛ/m)F_{n}=O(e^{-2\pi n\Lambda/m}) times a linear function in Λ\Lambda. Inserting this result back into the double convolution (Appendix A) gives

(f~0(Λ)f~0(Λ)f~0(Λ))(p)=1g2(m22p2)f~0(p)1g2n=1Fn(p).\left(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\right)(p)=\frac{1}{g^{2}}\left(\frac{m^{2}}{2}-p^{2}\right)\tilde{f}_{0}(p)-\frac{1}{g^{2}}\sum_{n=1}^{\infty}F_{n}(p)~{}. (A.29)

Therefore f~0(p)\tilde{f}_{0}(p) is indeed the leading order solution to the integral equation, (A.12). Furthermore, we find that the source term in the equation for the first correction, (A.13), is

F1(p)=6i2m2πg{2πΛm+1ln[4sinh2(πp/m)]}sinh(πp/m)e2πΛ/m.F_{1}(p)=-\frac{6i\sqrt{2}m^{2}}{\pi g}\left\{\frac{2\pi\Lambda}{m}+1-\ln[4\textrm{sinh}^{2}(\pi p/m)]\right\}\textrm{sinh}(\pi p/m)e^{-2\pi\Lambda/m}~{}. (A.30)

Notice that, while for fixed pp the source term (A.30) is O(Λe2πΛ/m)O(\Lambda\cdot e^{-2\pi\Lambda/m}), it is bounded by a quantity of O(ΛeπΛ/m)O(\Lambda\cdot e^{-\pi\Lambda/m}) for all p[Λ,Λ]p\in[-\Lambda,\Lambda]. In particular F1F_{1} vanishes at p=0p=0 as do all of the FnF_{n}.

Now consider the integral operator on the left-hand side of the linearized equation (A.13), which we denote Δ~(Λ)(pq)\tilde{\Delta}^{(\Lambda)}(p-q):

ΛΛ𝑑qΔ~(Λ)(pq)f~1(Λ)(q)=\displaystyle\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\Lambda}^{\Lambda}dq\tilde{\Delta}^{(\Lambda)}(p-q)\tilde{f}_{1}^{(\Lambda)}(q)= ((p2m22)f~1(Λ)+3g2(2π)2(f~0(Λ)f~0(Λ))f~1(Λ))(p).\displaystyle~{}\left(\left(p^{2}-\tfrac{m^{2}}{2}\right)\cdot\tilde{f}_{1}^{(\Lambda)}+\frac{3g^{2}}{(2\pi)^{2}}\left(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\right)\star\tilde{f}_{1}^{(\Lambda)}\right)(p)~{}. (A.31)

The double convolution can be written in the same form as (A.18), but with f~0(Λ)(q1)\tilde{f}_{0}^{(\Lambda)}(q_{1}) replaced with f~1(Λ)(q1)\tilde{f}_{1}^{(\Lambda)}(q_{1}). The principal value integration from [Λ,Λ][-\Lambda,\Lambda] arises from changing variables in the analog of (Appendix A) from qq to q=pqq^{\prime}=p-q. Thus

Δ~(Λ)(pq)=\displaystyle\quad\tilde{\Delta}^{(\Lambda)}(p-q)= (p2m22)δ(pq)+3g2(2π)2(f~0(Λ)f~0(Λ))(pq)\displaystyle~{}\left(p^{2}-\tfrac{m^{2}}{2}\right)\delta(p-q)+\frac{3g^{2}}{(2\pi)^{2}}\left(\tilde{f}_{0}^{(\Lambda)}\star\tilde{f}_{0}^{(\Lambda)}\right)(p-q) (A.32)
=\displaystyle= (p2+m2)δ(pq)3mπsinh(π|pq|m)ln[sinh(πΛm)sinh(π|Λ|pq||m)].\displaystyle~{}(p^{2}+m^{2})\delta(p-q)-\frac{3m}{\pi\textrm{sinh}(\frac{\pi|p-q|}{m})}\ln\left[\frac{\textrm{sinh}(\frac{\pi\Lambda}{m})}{\textrm{sinh}(\frac{\pi|\Lambda-|p-q||}{m})}\right]~{}.\qquad (A.33)

We would like to find an inverse for Δ~(Λ)\tilde{\Delta}^{(\Lambda)}, whose kernel is a Green function that we denote G~(Λ)(q,k)\tilde{G}^{(\Lambda)}(q,k), for q,k[Λ,Λ]q,k\in[-\Lambda,\Lambda]. In order to extract the leading behavior of f~1(Λ)\tilde{f}_{1}^{(\Lambda)}, however, it is sufficient to find an approximate inverse, whose kernel we denote G~(q,k)\tilde{G}(q,k), such that

ΛΛ𝑑q(ΛΛ𝑑qΔ~(Λ)(pq)G~(q,q))Fn(q)=Fn(p)(1+O((m/Λ)2)),\int_{-\Lambda}^{\Lambda}dq\left(\int_{-\Lambda}^{\Lambda}dq^{\prime}\tilde{\Delta}^{(\Lambda)}(p-q^{\prime})\tilde{G}(q^{\prime},q)\right)F_{n}(q)=F_{n}(p)(1+O((m/\Lambda)^{2}))~{}, (A.34)

where we recall that FnF_{n} is bounded, odd, and O(Λe2πnΛ/m)O(\Lambda\cdot e^{-2\pi n\Lambda/m}) for fixed pp with |p|<Λ|p|<\Lambda.

A standard Neumann series for the inverse, based on the form of (A.32) as a diagonal operator plus “correction,” where the correction is the second term, may not provide a sufficiently good approximation demonstrate this, since the correction is O(1)O(1) in a neighborhood of the diagonal. A natural ansatz for a better first approximation to G~(Λ)\tilde{G}^{(\Lambda)} is based on its Λ\Lambda\to\infty counterpart, which we describe next.

We claim that the approximate Green function can be taken to be the Green function of the standard Λ\Lambda\to\infty fluctuation operator, Δ~()\tilde{\Delta}^{(\infty)}, which we denote by G~()(p,q)\tilde{G}^{(\infty)}(p,q), restricted to p,q[Λ,Λ]p,q\in[-\Lambda,\Lambda]. We will discuss the error in this approximation after presenting Δ~()\tilde{\Delta}^{(\infty)} and its Green function. We have

Δ~()(pq)=\displaystyle\quad\tilde{\Delta}^{(\infty)}(p-q)= (p2+m2)δ(pq)3(pq)sinh(π(pq)/m),\displaystyle~{}(p^{2}+m^{2})\delta(p-q)-\frac{3(p-q)}{\textrm{sinh}(\pi(p-q)/m)}~{}, (A.35)

which can be obtained from the Fourier transform of the position space fluctuation operator around the kink:

Δ~()(pq)=\displaystyle\tilde{\Delta}^{(\infty)}(p-q)= 𝑑x𝑑yeipxiqyΔ()(xy),with\displaystyle~{}\int dxdye^{ipx-iqy}\Delta^{(\infty)}(x-y)~{},\qquad\textrm{with} (A.36)
Δ()(xy)=\displaystyle\Delta^{(\infty)}(x-y)= δ(xy)(x2m22+3g2f()(x)2),\displaystyle~{}\delta(x-y)\left(-\partial_{x}^{2}-\frac{m^{2}}{2}+3g^{2}f^{(\infty)}(x)^{2}\right)~{}, (A.37)

for which the eigenmodes are the gk(x)g_{k}(x) in (4.4). Δ()\Delta^{(\infty)} does not have an inverse due to the zero-mode gBg_{B}. It does have an inverse on the orthogonal complement of gBg_{B}, whose integral kernel is the Green function G()(x,y)G^{(\infty)}(x,y) satisfying

𝑑yΔ(xy)G(y,z)=δ(xz)gB(x)gB(z).\int dy\Delta(x-y)G(y,z)=\delta(x-z)-g_{B}(x)g_{B}(z)~{}. (A.38)

Imposing orthogonality to gBg_{B} and exponential fall-off at large |xy||x-y|, the solution to (A.38) is

G()(x,y)=\displaystyle G^{(\infty)}(x,y)= e2mx<+e2mx>+8(emx<+emx>)6m(x>x<)832mcosh2(mx>/2)cosh2(mx</2),\displaystyle~{}\frac{e^{2mx_{<}}+e^{-2mx_{>}}+8(e^{mx_{<}}+e^{-mx_{>}})-6m(x_{>}-x_{<})-8}{32m~{}\textrm{cosh}^{2}(mx_{>}/2)\textrm{cosh}^{2}(mx_{<}/2)}~{}, (A.39)

where x>=max(x,y)x_{>}=\max(x,y) and x<=min(x,y)x_{<}=\min(x,y).

Then G~()(p,q)\tilde{G}^{(\infty)}(p,q) is the Fourier transform satisfying

𝑑qΔ~(pq)G~(q,q)=δ(pq)g~B(p)g~B(q),\int_{-\infty}^{\infty}dq^{\prime}\tilde{\Delta}(p-q^{\prime})\tilde{G}(q^{\prime},q)=\delta(p-q)-\tilde{g}_{B}(p)\tilde{g}_{B}(-q)~{}, (A.40)

and we can compute it explicitly:

G~()(p,q)=\displaystyle\tilde{G}^{(\infty)}(p,q)= 2π(q2+m2)δ(pq)+G~reg(p,q),with\displaystyle~{}\frac{2\pi}{(q^{2}+m^{2})}\delta(p-q)+\tilde{G}_{\rm reg}(p,q)~{},\qquad\textrm{with} (A.41)
G~reg(p,q)=\displaystyle\tilde{G}_{\rm reg}(p,q)= πsinh(π(pq)/m){12p(q2+m2)212q(p2+m2)213pm2(q2+m2)+13qm2(p2+m2)+\displaystyle~{}\frac{\pi}{\textrm{sinh}(\pi(p-q)/m)}\bigg{\{}\frac{12p}{(q^{2}+m^{2})^{2}}-\frac{12q}{(p^{2}+m^{2})^{2}}-\frac{13p}{m^{2}(q^{2}+m^{2})}+\frac{13q}{m^{2}(p^{2}+m^{2})}+ (A.42)
+6m4(pq)+11pqm5(coth(πp/m)coth(πq/m))+\displaystyle~{}\qquad\qquad\qquad\qquad+\frac{6}{m^{4}}(p-q)+\frac{11pq}{m^{5}}\left(\coth(\pi p/m)-\coth(\pi q/m)\right)+ (A.43)
6pqm5Im(ψ(1)(1+ipm)ψ(1)(1+iqm))}.\displaystyle~{}\qquad\qquad\qquad\qquad-\frac{6pq}{m^{5}}~{}{\rm Im}\left(\psi^{(1)}(-1+\tfrac{ip}{m})-\psi^{(1)}(-1+\tfrac{iq}{m})\right)\bigg{\}}~{}. (A.44)

where ψ(1)(z)=ddzψ(z)\psi^{(1)}(z)=\frac{d}{dz}\psi(z) with ψ(z)\psi(z) the digamma function. Note that G~(p,q)=G~(q,p)\tilde{G}(p,q)=\tilde{G}(q,p).

The regular piece G~reg\tilde{G}_{\rm reg} is smooth and bounded on all of 2\mathbb{R}^{2}. The pole from the csch pre-factor along the diagonal is canceled by a zero from the quantity in curly brackets. For fixed pp, the large |q||q| behavior of the quantity in curly brackets is linear and therefore Greg(p,q)=O(|q|eπ|q|)G_{\rm reg}(p,q)=O(|q|e^{-\pi|q|}) as |q||q|\to\infty. Furthermore, along the diagonal, one can show that Greg(p,p)G_{\rm reg}(p,p) falls off like |p|4|p|^{-4} for large pp.

We use these facts to argue that the restriction of G~()(p,q)\tilde{G}^{(\infty)}(p,q) to p,q[Λ,Λ]p,q\in[-\Lambda,\Lambda] can be taken as the approximate inverse G~\tilde{G} in (A.34):

G~(p,q)=G~()(p,q)|[Λ,Λ]2.\tilde{G}(p,q)=\tilde{G}^{(\infty)}(p,q)\bigg{|}_{[-\Lambda,\Lambda]^{2}}~{}. (A.45)

There are two sources of error in this approximation. First there is the error in replacing the inverse of Δ~(Λ)\tilde{\Delta}^{(\Lambda)} by the inverse of Δ~()|[Λ,Λ]2\tilde{\Delta}^{(\infty)}|_{[-\Lambda,\Lambda]^{2}}. Then there is the error in approximating the inverse of the latter operator, which we will think of as a restriction of Δ~()\tilde{\Delta}^{(\infty)} to an upper-left block, with the upper-left block of its inverse.999There is no issue regarding the fact that G~()\tilde{G}^{(\infty)} is only an inverse to Δ~()\tilde{\Delta}^{(\infty)} on the orthogonal complement to the zero-mode g~B\tilde{g}_{B}. The reason is that we can restrict interest to G~\tilde{G} acting on functions Fn(q)F_{n}(q) that are odd on q[Λ,Λ]q\in[-\Lambda,\Lambda] and so are orthogonal to g~B\tilde{g}_{\rm B}, which is an even function of qq. We discuss each in turn.

To simplify notation, let T=Δ~(Λ)T=\tilde{\Delta}^{(\Lambda)} be the integral operator we have, and let T0=Δ~()|[Λ,Λ]2T_{0}=\tilde{\Delta}^{(\infty)}|_{[-\Lambda,\Lambda]^{2}}. Let the difference be δT\delta T with kernel

δT(p,q)=3(pq)sinh(π(pq)/m)3mπsinh(π|pq|m)ln[sinh(πΛm)sinh(π|Λ|pq||m)],\delta T(p,q)=\frac{3(p-q)}{\textrm{sinh}(\pi(p-q)/m)}-\frac{3m}{\pi\textrm{sinh}(\frac{\pi|p-q|}{m})}\ln\left[\frac{\textrm{sinh}(\frac{\pi\Lambda}{m})}{\textrm{sinh}(\frac{\pi|\Lambda-|p-q||}{m})}\right]~{}, (A.46)

which vanishes along the diagonal and is exponentially small in m/Λm/\Lambda in a neighborhood of the diagonal. Again, we note the logarithmic singularity at |pq|=Λ|p-q|=\Lambda is integrable and has a coefficient that is exponentially small in m/Λm/\Lambda. Thus we expect the corrections in approximating T1T^{-1} by T01T_{0}^{-1} from the Neemann series,

T1=\displaystyle T^{-1}= (T0+δT)1=T01T01δTT01+T01δTT01δTT01+\displaystyle~{}(T_{0}+\delta T)^{-1}=T_{0}^{-1}-T_{0}^{-1}\delta TT_{0}^{-1}+T_{0}^{-1}\delta TT_{0}^{-1}\delta TT_{0}^{-1}-+\cdots (A.47)
\displaystyle\approx T01,\displaystyle~{}T_{0}^{-1}~{}, (A.48)

to be exponentially small in m/Λm/\Lambda.

Now, T0T_{0} is the restriction of 𝒯=Δ~()\mathcal{T}=\tilde{\Delta}^{(\infty)} to functions with support on [Λ,Λ][-\Lambda,\Lambda], while we define G~\tilde{G} as the analogous restriction of G~()\tilde{G}^{(\infty)}, and what we know is that 𝒯G~()=𝟏\mathcal{T}\tilde{G}^{(\infty)}=\mathbf{1}. A finite-dimensional matrix analog of our problem is that we have that

𝒯=(T0T12T21T22),G~()=(G~G~12G~21G~22)\mathcal{T}=\left(\begin{array}[]{c c}T_{0}&T_{12}\\ T_{21}&T_{22}\end{array}\right)~{},\qquad\tilde{G}^{(\infty)}=\left(\begin{array}[]{c c}\tilde{G}&\tilde{G}_{12}\\ \tilde{G}_{21}&\tilde{G}_{22}\end{array}\right)~{} (A.49)

are inverses of each other and we would like an expression for T01T_{0}^{-1}. Assuming G22G_{22} is invertible, then the finite-dimensional formula from block inversion is

T01=G~G~12G~221G~21.T_{0}^{-1}=\tilde{G}-\tilde{G}_{12}\tilde{G}_{22}^{-1}\tilde{G}_{21}~{}. (A.50)

We assume the same formula holds in our setting, and we expect G~22\tilde{G}_{22} to be invertible since it has the form of a diagonal matrix plus a small correction. Then the above remarks on the properties of G~()\tilde{G}^{(\infty)} imply that G~12(p,q)\tilde{G}_{12}(p,q), with |p|<Λ|p|<\Lambda and |q|>Λ|q|>\Lambda, is generally exponentially suppressed, except in neighborhoods of the corners where |p|Λ|q||p|\sim\Lambda\sim|q|. In these neighborhoods we still have that G~12(p,q)\tilde{G}_{12}(p,q) is bounded by a quantity of O(|q|4)O(|q|^{-4}). Analogous remarks apply to G~21=(G~12)T\tilde{G}_{21}=(\tilde{G}_{12})^{T}. Therefore we expect that G~12G~221G~21\tilde{G}_{12}\tilde{G}_{22}^{-1}\tilde{G}_{21} is suppressed101010In more detail, when p,qp,q are away from the diagonal G~12G~221G~21\tilde{G}_{12}\tilde{G}_{22}^{-1}\tilde{G}_{21} has double the exponential suppression in the distance from the diagonal as G~\tilde{G}. Along the diagonal, assuming the dominant behavior of (G~22)1(p,q)(\tilde{G}_{22})^{-1}(p,q) is p2δ(pq)\sim p^{2}\delta(p-q), the contribution of G~12G~221G~21\tilde{G}_{12}\tilde{G}_{22}^{-1}\tilde{G}_{21} is estimated by an integral of the form 0𝑑x(x+Λm)6ex\int_{0}^{\infty}dx(x+\frac{\Lambda}{m})^{-6}e^{-x}, which is O(1/Λ6)O(1/\Lambda^{6}). by at least O(m2/Λ2)O(m^{2}/\Lambda^{2}) relative to G~\tilde{G} for any p,q[Λ,Λ]p,q\in[-\Lambda,\Lambda]. Therefore combining (A.47) and (A.50) we write

G~(Λ)(p,q)T1(p,q)=G~(p,q)(1+O(m2/Λ2)),\tilde{G}^{(\Lambda)}(p,q)\equiv T^{-1}(p,q)=\tilde{G}(p,q)\left(1+O(m^{2}/\Lambda^{2})\right)~{}, (A.51)

which gives (A.34).

It follows that the solution to (A.13) for the first correction is

f~1(p)=ΛΛ𝑑qG~(p,q)F1(q)×(1+O(m2/Λ2)).\displaystyle\tilde{f}_{1}(p)=\int_{-\Lambda}^{\Lambda}dq\tilde{G}(p,q)F_{1}(q)\times(1+O(m^{2}/\Lambda^{2}))~{}. (A.52)

Note that for fixed pp the exponential damping in |pq||p-q| from the csch pre-factor of G~reg\tilde{G}_{\rm reg} balances the exponential growth from the sinh pre-factor of F1F_{1}. Since the integrand then grows linearly in qq at large qq and the explicit leading Λ\Lambda dependence of F1F_{1} is O(Λe2πΛ/m)O(\Lambda\cdot e^{-2\pi\Lambda/m}), one sees that the leading behavior of f~1\tilde{f}_{1} will be O(Λ3e2πΛ/m)O(\Lambda^{3}\cdot e^{-2\pi\Lambda/m}). Since the Λ\Lambda dependence of the integrand of (A.52) is simple, the leading Λ\Lambda behavior of the integral can be computed by applying the fundamental theorem of calculus. The result is

f~1(Λ)(p,Λ)=\displaystyle\tilde{f}_{1}^{(\Lambda)}(p,\Lambda)= (A.53)
=4πi2gsinh(πp/m){(6p4m2p2+5m4)(p2+m2)2+3p2mIm[ψ(1)(1+ipm)]}Λ3m3e2πΛ/m+\displaystyle=-\frac{4\pi i\sqrt{2}}{g}\textrm{sinh}(\pi p/m)\left\{\frac{(6p^{4}-m^{2}p^{2}+5m^{4})}{(p^{2}+m^{2})^{2}}+\frac{3p}{2m}\operatorname{Im}\left[\psi^{(1)}(-1+i\tfrac{p}{m})\right]\right\}\frac{\Lambda^{3}}{m^{3}}e^{-2\pi\Lambda/m}+ (A.54)
+O(Λ2m2e2πΛ/m).\displaystyle\qquad\quad+O\left(\tfrac{\Lambda^{2}}{m^{2}}e^{-2\pi\Lambda/m}\right)~{}. (A.55)

Thus for fixed pp we have f~1(Λ)=O(Λ3e2πΛ/m)\tilde{f}_{1}^{(\Lambda)}=O(\Lambda^{3}\cdot e^{-2\pi\Lambda/m}), and for all p[Λ,Λ]p\in[-\Lambda,\Lambda] we have that f~1(Λ)\tilde{f}_{1}^{(\Lambda)} is bounded by a quantity of O(Λ3eπΛ/m)O(\Lambda^{3}\cdot e^{-\pi\Lambda/m}).

Since f~1(Λ)\tilde{f}_{1}^{(\Lambda)} has the same exponentially suppressed form as the source F1F_{1}, it follows from the above analysis that the linearized equation for the nthn^{\rm th} correction, f~n(Λ)\tilde{f}_{n}^{(\Lambda)}, will be of the form Δ~(Λ)f~n(Λ)=Fn\tilde{\Delta}^{(\Lambda)}\cdot\tilde{f}_{n}^{(\Lambda)}=F_{n}^{\prime}, where FnF_{n}^{\prime} has the same suppression as FnF_{n}. Hence f~n(Λ)\tilde{f}_{n}^{(\Lambda)} will also have the same exponential suppression as FnF_{n}, and will generally involve a polynomial in Λ/m\Lambda/m whose degree grows with nn.

Appendix B IR Divergences

At two or three loops, depending on the theory in general the energy of the kink ground state is IR divergent. More specifically, the Sine-Gordon kink has a divergence at three loops, the ϕ4\phi^{4} kink at two, and if the potential expanded around a minimum corresponding to each end of the kink begins its polynomial expansion at ϕn\phi^{n}, the first IR divergence will occur at n1n-1 loops. These IR divergences do not affect the kink mass as they also appear, at one less loop, in the vacuum energy and the kink mass is the difference between the two energies.

Here at one loop we found the energy

Q1=dk2πΛΛdp2πg~k(p)g~k(p)(ωpωk)24ωp.Q_{1}=-\int\frac{dk}{2\pi}\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\tilde{g}_{k}(p)\tilde{g}^{*}_{k}(p)\frac{\left(\omega_{p}-\omega_{k}\right)^{2}}{4\omega_{p}}. (B.1)

In the cases of the Sine-Gordon and ϕ4\phi^{4} modes, g~k(p)\tilde{g}_{k}(p) contains a δ(kp)\delta(k-p) term. One of these δ\delta functions may be eliminated by the integration over kk, but the other potentially leaves a divergence already at one loop. In coordinate space, it is easily seen that the origin of this divergence is a constant energy density, and so it is an IR divergence.

Such IR divergences have been noted since the first papers [1] on kink mass calculations, and are usually regularized by compactifying the space using periodic boundary conditions. This is quite a high price to pay111111While the price is high, the cut-off compactified theory has a finite number of modes and so can be treated nonperturbatively using Monte Carlo [29, 30, 31] and variational [32] techniques., as the kink itself does not satisfy periodic boundary conditions. One may instead impose boundary conditions that are satisfied by the kink, but then they will not be satisfied by the vacuum.

We will now argue that, at least in the present case, the constant energy density is exactly zero and so this messy issue may be avoided. First, let us return to position space

Q1=ΛΛdp2πdk2π𝑑x𝑑ygk(x)gk(y)eip(yx)(ωpωk)24ωp.Q_{1}=-\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\int\frac{dk}{2\pi}\int dx\int dy{g}_{-k}(x){g}_{k}(y)e^{ip(y-x)}\frac{\left(\omega_{p}-\omega_{k}\right)^{2}}{4\omega_{p}}. (B.2)

The potential divergence now arises from the plane wave terms

gk(x)|k|ωkeikx.g_{k}(x)\supset\frac{|k|}{\omega_{k}}e^{-ikx}. (B.3)

Thus the potentially divergent term in Q1Q_{1} is

Q^1=ΛΛdp2πdk2π𝑑x𝑑yei(pk)(yx)k2ωk2(ωpωk)24ωp.\hat{Q}_{1}=-\int^{\Lambda}_{-\Lambda}\frac{dp}{2\pi}\int\frac{dk}{2\pi}\int dx\int dye^{i(p-k)(y-x)}\frac{k^{2}}{\omega^{2}_{k}}\frac{\left(\omega_{p}-\omega_{k}\right)^{2}}{4\omega_{p}}. (B.4)

The potential divergence arises from pkp\sim k so let us expand in 1/ϵ1/\epsilon where ϵ=pk\epsilon=p-k. At leading order

ωpωk=ϵωp\omega_{p}-\omega_{k}=\frac{\epsilon}{\omega_{p}} (B.5)

and so at leading order, at pkp\sim k, we have

Q^1=dp2πp24ωp5dϵ2π𝑑x𝑑yeiϵ(yx)ϵ2.\hat{Q}_{1}=-\int\frac{dp}{2\pi}\frac{p^{2}}{4\omega_{p}^{5}}\int\frac{d\epsilon}{2\pi}\int dx\int dye^{i\epsilon(y-x)}\epsilon^{2}. (B.6)

Notice that the exponential is periodic in both xx and yy with period 2π/ϵ2\pi/\epsilon. It is also bounded by 11, as is its norm. Therefore integrating eiϵ(yx)e^{i\epsilon(y-x)} over any rectangle on the xyx-y plane, the integral will never exceed 4π2/ϵ24\pi^{2}/\epsilon^{2}. Thus the integral of ϵ2eiϵ(yx)\epsilon^{2}e^{i\epsilon(y-x)} never exceeds 4π24\pi^{2} and in particular is bounded. When ϵ0\epsilon\neq 0 the integral vanishes in the sense of a distribution as it is periodic. Therefore, after integrating over xx and yy one arrives at a function of ϵ\epsilon which is bounded and vanishes except on the measure zero set ϵ=0\epsilon=0. The integral over ϵ\epsilon is therefore equal to zero. Thus, at leading order in 1/ϵ1/\epsilon, this integral vanishes. We conclude that there is no small ϵ\epsilon divergence, and so the potential IR divergence is not present.

Note that the above derivation did not use the pp integral, it was performed independently at each value of pp. Therefore it is not affected by the cutoff at Λ\Lambda.

Acknowledgement

JE is supported by the CAS Key Research Program of Frontier Sciences grant QYZDY-SSW-SLH006 and the NSFC MianShang grants 11875296 and 11675223. JE also thanks the Recruitment Program of High-end Foreign Experts for support. ABR is supported by NSF grant number PHY-2112781.

References

  • [1] R. F. Dashen, B. Hasslacher and A. Neveu, “Nonperturbative Methods and Extended Hadron Models in Field Theory 2. Two-Dimensional Models and Extended Hadrons,” Phys. Rev. D 10 (1974) 4130. doi:10.1103/PhysRevD.10.4130
  • [2] R. Rajaraman, “Some Nonperturbative Semiclassical Methods in Quantum Field Theory: A Pedagogical Review,” Phys. Rept.  21 (1975) 227. doi:10.1016/0370-1573(75)90016-2
  • [3] A. S. Goldhaber, A. Rebhan, P. van Nieuwenhuizen and R. Wimmer, “Quantum corrections to mass and central charge of supersymmetric solitons,” Phys. Rept.  398 (2004) 179 doi:10.1016/j.physrep.2004.05.001 [hep-th/0401152].
  • [4] I. Takyi, M. K. Matfunjwa and H. Weigel, “Quantum corrections to solitons in the Φ8\Phi^{8} model,” Phys. Rev. D 102 (2020) no.11, 116004 doi:10.1103/PhysRevD.102.116004 [arXiv:2010.07182 [hep-th]].
  • [5] J. L. Gervais, A. Jevicki and B. Sakita, “Perturbation Expansion Around Extended Particle States in Quantum Field Theory. 1.,” Phys. Rev. D 12 (1975), 1038 doi:10.1103/PhysRevD.12.1038
  • [6] A. Rebhan and P. van Nieuwenhuizen, “No saturation of the quantum Bogomolnyi bound by two-dimensional supersymmetric solitons,” Nucl. Phys. B 508 (1997) 449 doi:10.1016/S0550-3213(97)00625-1, 10.1016/S0550-3213(97)80021-1 [hep-th/9707163].
  • [7] H. Nastase, M. A. Stephanov, P. van Nieuwenhuizen and A. Rebhan, “Topological boundary conditions, the BPS bound, and elimination of ambiguities in the quantum mass of solitons,” Nucl. Phys. B 542 (1999), 471-514 doi:10.1016/S0550-3213(98)00773-1 [arXiv:hep-th/9802074 [hep-th]].
  • [8] A. Litvintsev and P. van Nieuwenhuizen, “Once more on the BPS bound for the SUSY kink,” [arXiv:hep-th/0010051 [hep-th]].
  • [9] S. D. Drell, M. Weinstein and S. Yankielowicz, “Variational Approach to Strong Coupling Field Theory. 1. Φ4\Phi^{4} Theory,” Phys. Rev. D 14 (1976), 487 doi:10.1103/PhysRevD.14.487
  • [10] E. Tomboulis, “Canonical Quantization of Nonlinear Waves,” Phys. Rev. D 12 (1975) 1678 doi:10.1103/PhysRevD.12.1678
  • [11] K. E. Cahill, A. Comtet and R. J. Glauber, “Mass Formulas for Static Solitons,” Phys. Lett. B 64 (1976), 283-285 doi:10.1016/0370-2693(76)90202-1
  • [12] K. Hepp, “The Classical Limit for Quantum Mechanical Correlation Functions,” Commun. Math. Phys.  35 (1974) 265. doi:10.1007/BF01646348
  • [13] J. Sato and T. Yumibayashi, “Quantum-classical correspondence via coherent state in integrable field theory,” [arXiv:1811.03186 [quant-ph]].
  • [14] R. B. Pearson, “Variational Methods and Bounds for Lattice Field Theories,” Ph.D. thesis, SLAC report, 1975 (unpublished); R. Blankenbecler and R. B. Pearson, (unpublished)
  • [15] S. D. Drell, M. Weinstein and S. Yankielowicz, “Variational Approach to Strong Coupling Field Theory. 1. Phi**4 Theory,” Phys. Rev. D 14 (1976), 487 doi:10.1103/PhysRevD.14.487
  • [16] S. R. Coleman, “The Quantum Sine-Gordon Equation as the Massive Thirring Model,” Phys. Rev. D 11 (1975) 2088. doi:10.1103/PhysRevD.11.2088
  • [17] J. Evslin, “Manifestly Finite Derivation of the Quantum Kink Mass,” JHEP 11 (2019), 161 doi:10.1007/JHEP11(2019)161 [arXiv:1908.06710 [hep-th]].
  • [18] J. Evslin, “Well-defined quantum soliton masses without supersymmetry,” Phys. Rev. D 101 (2020) no.6, 065005 doi:10.1103/PhysRevD.101.065005 [arXiv:2002.12523 [hep-th]].
  • [19] J. Evslin and H. Guo, “Two-Loop Scalar Kinks,” [arXiv:2012.04912 [hep-th]].
  • [20] J. Evslin, “ϕ4\phi^{4} kink mass at two loops,” Phys. Rev. D 104 (2021) no.8, 085013 doi:10.1103/PhysRevD.104.085013 [arXiv:2104.07991 [hep-th]].
  • [21] J. Evslin, “The two-loop ϕ4\phi^{4} kink mass,” Phys. Lett. B 822 (2021), 136628 doi:10.1016/j.physletb.2021.136628 [arXiv:2109.05852 [hep-th]].
  • [22] M. Serone, G. Spada and G. Villadoro, “λϕ24\lambda\phi_{2}^{4} theory — Part II. the broken phase beyond NNNN(NNNN)LO,” JHEP 05 (2019), 047 doi:10.1007/JHEP05(2019)047 [arXiv:1901.05023 [hep-th]].
  • [23] H. Guo and J. Evslin, “Finite derivation of the one-loop sine-Gordon soliton mass,” JHEP 02 (2020), 140 doi:10.1007/JHEP02(2020)140 [arXiv:1912.08507 [hep-th]].
  • [24] M. A. Shifman, A. I. Vainshtein and M. B. Voloshin, “Anomaly and quantum corrections to solitons in two-dimensional theories with minimal supersymmetry,” Phys. Rev. D 59 (1999), 045016 doi:10.1103/PhysRevD.59.045016 [arXiv:hep-th/9810068 [hep-th]].
  • [25] G. ’t Hooft, “Topology of the Gauge Condition and New Confinement Phases in Nonabelian Gauge Theories,” Nucl. Phys. B 190 (1981) 455. doi:10.1016/0550-3213(81)90442-9
  • [26] S. Mandelstam, “Vortices and Quark Confinement in Nonabelian Gauge Theories,” Phys. Rept.  23 (1976) 245. doi:10.1016/0370-1573(76)90043-0
  • [27] J. Greensite, “The Confinement problem in lattice gauge theory,” Prog. Part. Nucl. Phys. 51 (2003), 1 doi:10.1016/S0146-6410(03)90012-3 [arXiv:hep-lat/0301023 [hep-lat]].
  • [28] Hörmander, Lars, The analysis of linear partial differential operators. I, Springer-Verlag, Berlin (2003), doi:10.1007/978-3-642-61497-2.
  • [29] D. Lee, N. Salwen and D. Lee, “The Diagonalization of quantum field Hamiltonians,” Phys. Lett. B 503 (2001), 223-235 doi:10.1016/S0370-2693(01)00197-6 [arXiv:hep-th/0002251 [hep-th]].
  • [30] S. Rychkov and L. G. Vitale, “Hamiltonian truncation study of the ϕ4\phi^{4} theory in two dimensions,” Phys. Rev. D 91 (2015), 085011 doi:10.1103/PhysRevD.91.085011 [arXiv:1412.3460 [hep-th]].
  • [31] Z. Bajnok and M. Lajer, “Truncated Hilbert space approach to the 2d ϕ4\phi^{4} theory,” JHEP 10 (2016), 050 doi:10.1007/JHEP10(2016)050 [arXiv:1512.06901 [hep-th]].
  • [32] T. D. Karanikolaou, P. Emonts and A. Tilloy, “Gaussian Continuous Tensor Network States for Simple Bosonic Field Theories,” [arXiv:2006.13143 [cond-mat.str-el]].