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

Note on the nonrelativistic TT¯T\bar{T} deformation

Bin Chen1,2,3 bchen01@pku.edu.cn    Jue Hou1 houjue@pku.edu.cn    Jia Tian4 wukongjiaozi@ucas.ac.cn 1School of Physics, Peking University, No.5 Yiheyuan Rd, Beijing 100871, P. R. China
2Center for High Energy Physics, Peking University, No.5 Yiheyuan Rd, Beijing 100871, P. R. China
3Peng Huanwu Center for Fundamental Theory, Hefei, Anhui 230026, China
4Kavli Institute for Theoretical Sciences (KITS),
University of Chinese Academy of Science, 100190 Beijing, P. R. China
Abstract

In this paper, we present our study on the TT¯T\bar{T}-deformation of non-relativistic complex scalar field theory. We find the closed form of the deformed Lagrangian by using the perturbation and the method of characteristics. Furthermore we compute the exact energy spectrum of the deformed free theory by using the Brillouin-Wigner perturbation theory in an appropriate regularization scheme.

I Introduction

Solvable irrelevant deformations have attracted many interests in recent years due to their novel features that provide analytical control on the ultraviolet (UV) physics regardless of the difficulties from strong coupling or nonlocality. TT¯T\bar{T}-deformation is the most studied example of such deformations111For a pedagogical review see LectureJiang , and for recent progress see the workshop TTbar2020 . Smirnov:2016lqw ; Cavaglia:2016oda . It is originally defined in two-dimensional relativistic quantum field theories where TT¯T\bar{T} is a well defined composite operator that can be written as the determinant of the energy momentum tensor operator. Many interesting physical quantities such as the energy spectrum, the partition function, the entanglement entropy and the S-matrix have been computed exactly in the deformed theories, see LectureJiang and the references within. These exact results reveal fruitful and meanwhile unexpected structures of solvable irrelevant deformations which we have not fully understood. For example Hage1 ; Hage2 ; Hage3 , the deformed density of states of a deformed field theory shows the Hagedorn behavior which typically happens in string theory. On the other hand, the TT¯T\bar{T}-deformation to holographic conformal field theory opens a new window to study AdS/CFT correspondence verlinde ; Hage1 ; kutasov2 ; dubovsky1 ; dubovsky2 ; bihdtott ; ttandcf ; cutoffads ; commentsontt ; Chen:2018eqk ; Hartman:2018tkw ; Caputa:2019pam ; Murdia:2019fax ; Jeong:2019ylz ; Guica:2019nzm ; Chen:2019mis ; Grieninger:2019zts ; Lewkowycz:2019xse ; Apolo:2019zai ; Hirano:2020nwq ; Jafari:2019qns ; He:2020hhm ; Kruthoff:2020hsi .

Recently similar integrable deformations have been considered in other types of models including integrable lattice models Lattice1 ; Lattice2 and non-relativistic integrable field theories NonRe1 ; NonRe2 ; BoseJ . Interestingly TT¯T\bar{T} (like) deformation can be interpreted as a special type of algebra-preserving deformations studied in the context of integrable spin chains Long1 . The standard integrability technique of solving these integrable deformed models is first to compute the deformed S-matrix in the infinite volume limit where the deformation simply modifies the S-matrix by multiplying a 1. Castillejo-Dalitz-Dyson-like factor Smirnov:2016lqw ; CDD ; CDD2 , and then to substitute the deformed S-matrix into the Bethe equation to solve the theory in the finite volume. Using this integrability technique the deformed one-dimensional Bose gas known as the Lieb-Liniger model was carefully studied in BoseJ where the author showed that the deformed one-dimensional Bose gas shares many qualitative features with the TT¯T\bar{T}-deformed relativistic quantum field theories. For example, the spectrum can also be given by a flow equation and the density of states exhibits the Hagedorn behavior in the thermodynamic limit.

In this work we will use another method, the quantum perturbation theory, to study the spectrum of the deformed non-relativistic field theories which are not necessarily integrable. The model we focus on is the non-relativistic Schrödinger model which can be viewed as the free limit of the Lieb-Liniger model. We find that for models with a degenerate Legendre transformation such as non-relativistic Schrödinger or Lieb-Liniger model, the Lagrangian and Hamiltonian definition of the TT¯T\bar{T}-deformation may not be equivalent. In particular, within the Hamiltonian definition of TT¯T\bar{T}-deformation only the flow equation of the Hamiltonian is not enough to define the deformation, one also needs to specify how the constraints change under the flow. In the first-order deformation, the Legendre transformation is degenerate so one can use the Dirac-Bergmann algorithm to quantize the system and compute the spectrum perturbatively. We find that the perturbative expansion of the deformed spectrum is superficially divergent as expected for the models under an irrelevant deformation. After imposing the Dirichlet regularization and dropping off the transcendental terms which are not supposed to appear in the TT¯T\bar{T}-deformation, the deformed spectrum match the results derived from the flow equation. We will also show that after including the second-order deformation the theory either does not admit a perturbative description or has ambiguities.

The paper is organized as follows: in Sect. II, we derive the closed form of TT¯T\bar{T}-deformed Lagrangian in non-relativistic complex scalar theory; in Sect. III, we compute the deformed energy spectrum perturbatively, and manage to find an exact form by using Brillouin-Wigner perturbation; in Sect. IV, we end with conclusions; we put some technical details into four appendices.


Note added: When we were finishing this project and had obtained the main results in this paper, two interesting papers Italy ; JiangNew appeared on arXiv in the same day. The same closed form of a TT¯T\bar{T}-deformed Lagrangian of non-relativistic models have also been derived. In Italy , the Lagrangian is derived from the dynamical change of coordinates, and in JiangNew it is derived using a geometric method. Our method in this work is similar to the one used in Close , and is somewhat complementary to the ones in222We would like to mention that the closed form of the TT¯T\bar{T}-deformed Lagrangian of non-relativistic models was also found by S. Frolov et.al. from the light-cone gauge interpretation LG . Italy ; JiangNew .

II Non-relativistic TT¯T\bar{T}-deformed Lagrangian

In this section, we consider the TT¯T\bar{T}-deformation of a two-dimensional (2D) non-relativistic field theory whose Lagrangian satisfies333We will work in the convention that gμν=diag(1,1),xμ=(x0,x1)=(t,x).\begin{split}g_{\mu\nu}=\mbox{diag}(1,-1),\hskip 12.91663ptx^{\mu}=(x^{0},x^{1})=(t,x).\\ \end{split}

(λ)λ=det(Tμν(λ)),\begin{split}\frac{\partial\mathcal{L}^{(\lambda)}}{\partial\lambda}=\operatorname{det}\left(T_{\mu\nu}^{(\lambda)}\right),\end{split} (1)

where TμνT_{\mu\nu} is the energy momentum tensor of this theory. Our goal is to solve this flow equation (1) for a non-relativistic complex scalar theory with a potential V(|ϕ|)V(|\phi|).

II.1 One real scalar case

To get some insights about (1) we warm up with a toy model involved with only a free real scalar, whose Lagrangian density reads

0=ϕ0ϕ+(1ϕ)2.\begin{split}\mathcal{L}_{0}=\phi\partial_{0}\phi+(\partial_{1}\phi)^{2}.\end{split} (2)

To solve (1) we expand the Lagrangian with respect to λ\lambda:

=n=0λnn.\begin{split}\mathcal{L}=\sum_{n=0}\lambda^{n}\mathcal{L}_{n}.\end{split} (3)

Substituting this expansion into the flow equation (1) one can read off n\mathcal{L}_{n} term by term, and the first several orders of the Lagrangian are of the forms

0=X+Y,1=Y(X+Y),2=Y(X2+3XY+2Y2),3=Y(X3+6X2Y+10XY2+5Y3),4=Y(X4+10X3Y+30X2Y2+35XY3+14Y4),\begin{split}&\mathcal{L}_{0}=X+Y,\\ &\mathcal{L}_{1}=Y(X+Y),\\ &\mathcal{L}_{2}=Y(X^{2}+3XY+2Y^{2}),\\ &\mathcal{L}_{3}=Y(X^{3}+6X^{2}Y+10XY^{2}+5Y^{3}),\\ &\mathcal{L}_{4}=Y(X^{4}+10X^{3}Y+30X^{2}Y^{2}+35XY^{3}+14Y^{4}),\\ \end{split} (4)

where we have defined the convenient variables:

Xϕ0ϕ,Y(1ϕ)2.\begin{split}X\equiv\phi\partial_{0}\phi,\hskip 12.91663ptY\equiv(\partial_{1}\phi)^{2}.\end{split} (5)

We observe that these terms can be cast into a general form as

n=Yk=0n1kXnkYkCn1k1Cn+kn1=XnYF12(n,n+1,2,YX),n=1,2,3,.\begin{split}\mathcal{L}_{n}&=Y\sum_{k=0}^{n}\frac{1}{k}X^{n-k}Y^{k}C_{n-1}^{k-1}C_{n+k}^{n-1}\\ &=X^{n}Y{}_{2}F_{1}\left(-n,n+1,2,-\frac{Y}{X}\right),\hskip 12.91663ptn=1,2,3,....\end{split} (6)

Performing the summation (3) directly gives the closed form of the Lagrangian

=X+k=0n=knλnY1kXnkYkCn1k1Cn+kn1=12λ(1λX+(1λX)24λY),\begin{split}\mathcal{L}&=X+\sum_{k=0}^{\infty}\sum_{n=k}^{n}\lambda^{n}Y\frac{1}{k}X^{n-k}Y^{k}C_{n-1}^{k-1}C_{n+k}^{n-1}\\ &=-\frac{1}{2\lambda}\left(-1-\lambda X+\sqrt{(1-\lambda X)^{2}-4\lambda Y}\right),\end{split} (7)

which indeed solves the flow equation (1).

Next we add in a potential VV which is just a function of ϕ\phi but does not contain the terms of the derivative of ϕ\phi. As before we expand the Lagrangian with respect to λ\lambda. The first few terms in the deformed Lagrangian are

0=V+X+Y1=V2+VX+Y(X+Y)2=V3+V2XVY2+Y(X2+3XY+2Y2)3=V4+V3XVY2(3X+4Y)+Y(X3+6X2Y+10XY2+5Y3)4=V5+V4X+2V2Y3VY2(6X2+20XY+15Y2)+Y(X4+10X3Y+30X2Y2+35XY3+14Y4),\begin{split}\mathcal{L}_{0}=&-V+X+Y\\ \mathcal{L}_{1}=&-V^{2}+VX+Y(X+Y)\\ \mathcal{L}_{2}=&-V^{3}+V^{2}X-VY^{2}+Y(X^{2}+3XY+2Y^{2})\\ \mathcal{L}_{3}=&-V^{4}+V^{3}X-VY^{2}(3X+4Y)\\ &+Y(X^{3}+6X^{2}Y+10XY^{2}+5Y^{3})\\ \mathcal{L}_{4}=&-V^{5}+V^{4}X+2V^{2}Y^{3}-VY^{2}(6X^{2}+20XY+15Y^{2})\\ &+Y(X^{4}+10X^{3}Y+30X^{2}Y^{2}+35XY^{3}+14Y^{4}),\\ \end{split} (8)

which can be rewritten in a general form as

n=Vn+1+VnX+j=0n(V)jYj+1k=0n2jK(n,j,k),K(n,j,k)=Xn2jkYkCn+knj1Cnj1k1+jCk1jj1k.\begin{split}\mathcal{L}_{n}&=-V^{n+1}+V^{n}X+\sum_{j=0}^{n}(-V)^{j}Y^{j+1}\sum_{k=0}^{n-2j}K(n,j,k),\\ K(n,j,k)&=X^{n-2j-k}Y^{k}C_{n+k}^{n-j-1}C_{n-j-1}^{k-1+j}C_{k-1-j}^{j}\frac{1}{k}.\end{split} (9)

Substituting (9) into (3) and performing the summation we end up with the closed form of the deformed Lagrangian

=V1λV12λ(1λV))[1λX+(1λX)24λ(1λV)Y].\begin{split}\mathcal{L}=&-\frac{V}{1-\lambda V}\\ &-\frac{1}{2\lambda(1-\lambda V))}\left[-1-\lambda X+\sqrt{(1-\lambda X)^{2}-4\lambda(1-\lambda V)Y}\right].\end{split} (10)

Alternatively the deformed Lagrangian can also be obtained by using the method of characteristics with the initial condition

0=ϕα+β2V,withα=0ϕ,β=1ϕ.\begin{split}\mathcal{L}_{0}=\phi\alpha+\beta^{2}-V,\quad\mbox{with}\hskip 8.61108pt\alpha=\partial_{0}\phi,\ \beta=\partial_{1}\phi.\end{split} (11)

In terms of these new variables α\alpha and β\beta the flow equation (1) could be rewritten as

λ=2+ββ+αα,\begin{split}\frac{\partial\mathcal{L}}{\partial\lambda}=-\mathcal{L}^{2}+\beta\mathcal{L}\frac{\partial\mathcal{L}}{\partial\beta}+\alpha\mathcal{L}\frac{\partial\mathcal{L}}{\partial\alpha},\end{split} (12)

which is equivalent to a set of equations:

{dλds=1dαds=αdβds=βdds=2(s=0)=ϕα(0)+β(0)2V{λ=sα=C3s+C2β=C4s+C2=1sC21C2=ϕC3C2+(C4C2)2V.\left\{\begin{aligned} &\frac{d\lambda}{ds}=1\\ &\frac{d\alpha}{ds}=-\alpha\mathcal{L}\\ &\frac{d\beta}{ds}=-\beta\mathcal{L}\\ &\frac{d\mathcal{L}}{ds}=-\mathcal{L}^{2}\\ &\mathcal{L}(s=0)\\ &=\phi\alpha(0)+\beta(0)^{2}-V\end{aligned}\right.\Rightarrow\hskip 8.61108pt\left\{\begin{aligned} &\lambda=s\\ &\alpha=\frac{C_{3}}{-s+C_{2}}\\ &\beta=\frac{C_{4}}{-s+C_{2}}\\ &\mathcal{L}=\frac{1}{s-C_{2}}\\ &-\frac{1}{C_{2}}=\phi\frac{C_{3}}{C_{2}}+(\frac{C_{4}}{C_{2}})^{2}-V.\end{aligned}\right. (13)

Canceling the constants CiC_{i}’s in (13), we get

=V1λV12λ(1λV))[1λX+(1λX)24λ(1λV)Y],\begin{split}\mathcal{L}&=-\frac{V}{1-\lambda V}\\ &-\frac{1}{2\lambda(1-\lambda V))}\left[-1-\lambda X+\sqrt{(1-\lambda X)^{2}-4\lambda(1-\lambda V)Y}\right],\end{split} (14)

which coincides with (10).

II.2 One complex scalar case

Now let us turn to the complex scalar case. The Lagrangian density of the free complex scalar theory is

0=i2(ϕ0ϕϕ0ϕ)1ϕ1ϕ\begin{split}\mathcal{L}_{0}=\frac{i}{2}(\phi^{\ast}\partial_{0}\phi-\phi\partial_{0}\phi^{\ast})-\partial_{1}\phi\partial_{1}\phi^{\ast}\end{split} (15)

Define α(0ϕ,0ϕ)\vec{\alpha}\equiv\left(\partial_{0}\phi,\partial_{0}\phi^{\ast}\right) and β(1ϕ,1ϕ)\vec{\beta}\equiv\left(\partial_{1}\phi,\partial_{1}\phi^{\ast}\right). The TT¯T\bar{T}-flow equation is given by

λ=2+ββ+αα(αα)(ββ)+(αβ)(βα).\begin{split}\frac{\partial\mathcal{L}}{\partial\lambda}=&-\mathcal{L}^{2}+\mathcal{L}\vec{\beta}\cdot\frac{\partial\mathcal{L}}{\partial\vec{\beta}}+\mathcal{L}\vec{\alpha}\cdot\frac{\partial\mathcal{L}}{\partial\vec{\alpha}}\\ &-\left(\frac{\partial\mathcal{L}}{\partial\vec{\alpha}}\cdot\vec{\alpha}\right)\left(\frac{\partial\mathcal{L}}{\partial\vec{\beta}}\cdot\vec{\beta}\right)+\left(\frac{\partial\mathcal{L}}{\partial\vec{\alpha}}\cdot\vec{\beta}\right)\left(\frac{\partial\mathcal{L}}{\partial\vec{\beta}}\cdot\vec{\alpha}\right).\end{split} (16)

Without the last two terms this flow equation (16) can be solved by using the method of characteristics with the initial condition

(s=0)=i2(ϕ(α)1ϕ(α)2)(α)1(β)2\begin{split}\mathcal{L}(s=0)=\frac{i}{2}\left(\phi^{\ast}(\vec{\alpha})_{1}-\phi(\vec{\alpha})_{2}\right)-(\vec{\alpha})_{1}(\vec{\beta})_{2}\end{split} (17)

in exactly the same way as we described in last section. The solution is

=12λ[1λX+(1λX)24λY],\begin{split}\mathcal{L}&=-\frac{1}{2\lambda}\left[-1-\lambda X+\sqrt{(1-\lambda X)^{2}-4\lambda Y}\right],\end{split} (18)

where X=i2(ϕ0ϕϕ0ϕ)X=\frac{i}{2}(\phi^{\ast}\partial_{0}\phi-\phi\partial_{0}\phi^{\ast}) and Y=1ϕ1ϕY=-\partial_{1}\phi\partial_{1}\phi^{\ast}.

To include the modifications of the last two terms in (16), we make an ansatz on the exact Lagrangian and expand it with respect to λ\lambda . By matching the expanded terms444The explicit forms of the Lagrangians i\mathcal{L}_{i} are quite involved, and we list them in Appendix A. to the first few orders we can fix our ansatz completely, and in the end we check our result with the flow equation. The final result we get is

=12λ[1λX+(1λX)24λY+8iλ2A4λ3B]\begin{split}\mathcal{L}&=-\frac{1}{2\lambda}\left[-1-\lambda X+\sqrt{(1-\lambda X)^{2}-4\lambda Y+8i\lambda^{2}A-4\lambda^{3}B}\right]\end{split} (19)

where

A=(ϕ1ϕ+ϕ1ϕ)(1ϕ0ϕ0ϕ1ϕ),B=ϕϕ(1ϕ0ϕ0ϕ1ϕ)2.\begin{split}&A=(\phi^{\ast}\partial_{1}\phi+\phi\partial_{1}\phi^{\ast})(\partial_{1}\phi\partial_{0}\phi^{\ast}-\partial_{0}\phi\partial_{1}\phi^{\ast}),\\ &B=\phi\phi^{\ast}(\partial_{1}\phi\partial_{0}\phi^{\ast}-\partial_{0}\phi\partial_{1}\phi^{\ast})^{2}.\\ \end{split} (20)

In a similar way, we can find the deformed Lagrangian in the case that the complex scalar has a potential VV,

=V1λV12λ(1λV)(1λX+(1λX)24λ(1λV)Y+8iλ2(1λV)A4λ3(1λV)B).\begin{split}&\mathcal{L}=-\frac{V}{1-\lambda V}-\frac{1}{2\lambda(1-\lambda V)}(-1-\lambda X+\\ &\sqrt{(1-\lambda X)^{2}-4\lambda(1-\lambda V)Y+8i\lambda^{2}(1-\lambda V)A-4\lambda^{3}(1-\lambda V)B}).\end{split} (21)

The detailed derivation is straightforward but too tedious to be presented here.

III Deformed energy from quantum perturbative theory

In this section, we will compute the spectrum of the TT¯T\bar{T}-deformed non-relativistic free boson (an effective Schrödinger model). The model we are considering is the free complex scalar defined on a compact region x[0,R]x\in[0,R], whose Lagrangian density and Hamiltonian are respectively

=i2(ϕtϕtϕϕ)xϕxϕ,H=dxxϕ(x)xϕ(x).\begin{split}&\mathcal{L}=\frac{i}{2}\left(\phi^{\ast}\partial_{t}\phi-\partial_{t}\phi^{\ast}\phi\right)-\partial_{x}\phi^{\ast}\partial_{x}\phi,\\ &H=\int\mathrm{d}x\,\partial_{x}\phi^{\ast}(x)\partial_{x}\phi(x).\end{split} (22)

This simple free theory can be though as the free boson limit of the Lieb-Liniger model whose TT¯T\bar{T}-deformation has been well studied in BoseJ . The momentum operator is

P=i2[ϕ(x)xϕ(x)xϕ(x)ϕ(x)].P=-\frac{i}{2}\int\left[\phi^{\ast}(x)\partial_{x}\phi(x)-\partial_{x}\phi^{\ast}(x)\phi(x)\right]. (23)

The complex scalar field satisfies the equal-time commutation relations,

[ϕ(x,t),ϕ(y,t)]=0,[ϕ(x,t),ϕ(y,t)]=0,[ϕ(x,t),ϕ(y,t)]=δ(xy).\begin{split}&[\phi(x,t),\phi(y,t)]=0,\quad\left[\phi^{\ast}(x,t),\phi^{\ast}(y,t)\right]=0,\\ &\left[\phi(x,t),\phi^{\ast}(y,t)\right]=\delta(x-y).\end{split} (24)

The Hilbert space is spanned by NN-particle states,

|x=ϕ(x1)ϕ(xN)|0.|\vec{x}\rangle=\phi^{\ast}\left(x_{1}\right)\ldots\phi^{\ast}\left(x_{N}\right)|0\rangle. (25)

Here, x=(x1,x2,,xN)\vec{x}=(x_{1},x_{2},...,x_{N}) and |0|0\rangle is the vacuum of Fock space, which is annihilated by ϕ(x)\phi(x). The NN-particle eigenfunction of this free model is simply given by the plane waves

x|u=ψN(ux)=1N!σSNexp[ij=1Nxjuσj],|uN=1N!dNxψN(ux)|x,\begin{split}&\langle\vec{x}|\vec{u}\rangle=\psi_{N}(\vec{u}\mid\vec{x})=\frac{1}{\sqrt{N!}}\sum_{\sigma\in\mathrm{S}_{N}}\exp\left[i\sum_{j=1}^{N}x_{j}u_{\sigma_{j}}\right],\\ &\left|\vec{u}_{N}\right\rangle=\frac{1}{\sqrt{N!}}\int\mathrm{d}^{N}x\psi_{N}(\vec{u}\mid\vec{x})|\vec{x}\rangle,\end{split} (26)

with the eigenvalues

EN(u)=j=1Nuj2,PN(u)=j=1Nuj,E_{N}(\vec{u})=\sum_{j=1}^{N}u_{j}^{2},\quad P_{N}(\vec{u})=\sum_{j=1}^{N}u_{j}, (27)

and

x|y=σSNδ(xiyσi).\langle\vec{x}|\vec{y}\rangle=\sum_{\sigma\in\mathrm{S}_{N}}\delta(x_{i}-y_{\sigma_{i}}).

III.1 Lagrangian vs. Hamiltonian definition

The TT¯T\bar{T}-deformation is formally defined as a flow of the Lagrangian density of the theory:

λ=ϵabϵcdTac(λ)Tbd(λ)𝒪TT¯(λ)\begin{split}\partial_{\lambda}\mathcal{L}=-\epsilon^{ab}\epsilon^{cd}T_{ac}^{(\lambda)}T_{bd}^{(\lambda)}\equiv-\mathcal{O}^{(\lambda)}_{T\bar{T}}\end{split} (28)

where Tab(λ)T_{ab}^{(\lambda)} is the stress tensor of the theory with the flow parameter λ\lambda. It is generally believed that at least in the classical level the deformation can be equivalently defined as a flow of the Hamiltonian density of theory:

λ=𝒪TT¯(λ).\begin{split}\partial_{\lambda}\mathcal{H}=\mathcal{O}^{(\lambda)}_{T\bar{T}}.\end{split} (29)

The classical equivalence has been examined in the leading order in Kruthoff:2020hsi and more generally in Jorjadze:2020ili . However the equivalence is built on an implicit assumption that the Lagrangian is not singular or equivalently the Legendre transformation is invertible such that the TT¯T\bar{T}-deformation and the Legendre transformation are commutative. The subtlety of a singular Lagrangian is that it leads to a constrained Hamiltonian. In this section we want to show that in the Hamiltonian formalism the flow equation (29) itself is not enough to defined the TT¯T\bar{T}-deformation.

To illustrate our point, let us consider the simple model (22) with the Lagrangian

=0+λ1,\begin{split}\mathcal{L}=&\mathcal{L}_{0}+\lambda\mathcal{L}_{1},\end{split} (30)

where the leading-order deformation of the Lagrangian is

1=i2(ϕtϕ(xϕ)2ϕtϕ(xϕ)2)+(xϕxϕ)2.\begin{split}\mathcal{L}_{1}=\frac{i}{2}\left(\phi\partial_{t}\phi(\partial_{x}\phi^{\ast})^{2}-\phi^{\ast}\partial_{t}\phi^{\ast}(\partial_{x}\phi)^{2}\right)+\left(\partial_{x}\phi^{\ast}\partial_{x}\phi\right)^{2}.\end{split} (31)

The energy density is now:

T00=(tϕ)tϕ+(tϕ)tϕ=xϕxϕλ(xϕxϕ)2.\begin{split}T_{00}=&\frac{\partial\mathcal{L}}{\partial(\partial_{t}\phi)}\partial_{t}\phi+\frac{\partial\mathcal{L}}{\partial(\partial_{t}\phi^{\ast})}\partial_{t}\phi^{\ast}-\mathcal{L}\\ =&\partial_{x}\phi\partial_{x}\phi^{\ast}-\lambda\left(\partial_{x}\phi^{\ast}\partial_{x}\phi\right)^{2}.\end{split} (32)

This energy density is supposed to be identified with Hamiltonian density \mathcal{H} after rewriting tϕ,tϕ\partial_{t}\phi,~{}\partial_{t}\phi^{\ast} in terms of ϕ,ϕ\phi,~{}\phi^{\ast} and the canonical momenta Π,Π\Pi,\Pi^{\ast}, which are given by

Π=(tϕ)=i2ϕ+iλ2ϕ(xϕ)2,Π=(tϕ)=i2ϕiλ2ϕ(xϕ)2.\begin{split}&\Pi=\frac{\partial\mathcal{L}}{\partial(\partial_{t}\phi)}=\frac{i}{2}\phi^{\ast}+\frac{i\lambda}{2}\phi(\partial_{x}\phi^{\ast})^{2},\\ &\Pi^{\ast}=\frac{\partial\mathcal{L}}{\partial(\partial_{t}\phi^{\ast})}=-\frac{i}{2}\phi-\frac{i\lambda}{2}\phi^{\ast}(\partial_{x}\phi)^{2}.\end{split} (33)

The problem is that the Lagrangian density \mathcal{L} is singular so that the rewriting can not be performed. Since the energy density (32) does not depend on tϕ,tϕ\partial_{t}\phi,~{}\partial_{t}\phi^{\ast} explicitly, one tends to claim

=T00=xϕxϕλ(xϕxϕ)2.\begin{split}\mathcal{H}=T_{00}=\partial_{x}\phi\partial_{x}\phi^{\ast}-\lambda\left(\partial_{x}\phi^{\ast}\partial_{x}\phi\right)^{2}.\end{split} (34)

However clearly (34) is not consistent with the definition (29) even in the leading order

=0+λ𝒪TT¯(λ)=xϕxϕλ(xϕxϕ)2iλ2(ϕtϕ(xϕ)2ϕtϕ(xϕ)2).\begin{split}&\mathcal{H}=\mathcal{H}_{0}+\lambda\mathcal{O}^{(\lambda)}_{T\bar{T}}\\ &=\partial_{x}\phi\partial_{x}\phi^{\ast}-\lambda\left(\partial_{x}\phi^{\ast}\partial_{x}\phi\right)^{2}-\frac{i\lambda}{2}\left(\phi\partial_{t}\phi(\partial_{x}\phi^{\ast})^{2}-\phi^{\ast}\partial_{t}\phi^{\ast}(\partial_{x}\phi)^{2}\right).\end{split} (35)

Strictly speaking, this is not the expression of Hamiltonian density because it contains the time-derivative terms tϕ,tϕ\partial_{t}\phi,~{}\partial_{t}\phi^{\ast} which come from TabT_{ab}. Using the equations of motion

itϕ+x2ϕ=itϕx2ϕ=0+𝒪(λ)\begin{split}i\partial_{t}\phi+\partial_{x}^{2}\phi=i\partial_{t}\phi^{\ast}-\partial_{x}^{2}\phi^{\ast}=0+\mathcal{O}(\lambda)\end{split} (36)

one can rewrite (LABEL:HTt) as

=xϕxϕλ(xϕxϕ)2+λ2((xϕ)2x2ϕϕ+(xϕ)2x2ϕϕ).\begin{split}&\mathcal{H}=\partial_{x}\phi\partial_{x}\phi^{\ast}-\lambda\left(\partial_{x}\phi^{\ast}\partial_{x}\phi\right)^{2}\\ &+\frac{\lambda}{2}\left((\partial_{x}\phi^{\ast})^{2}\partial_{x}^{2}\phi\phi+(\partial_{x}\phi)^{2}\partial_{x}^{2}\phi^{\ast}\phi^{\ast}\right).\end{split} (37)

It is obvious that (34) and (LABEL:HT001) are in mismatch. This mismatch is problematic when one tries to compute the spectrum of the theory perturbatively. This problem does not appear in a relativistic theory, for example in a theory of the relativistic free boson. In the relativistic case, the canonical momentum depends on the flow parameter as well such that the deformed Hamiltonian densities are the same at the classical level in two different ways, as we show in the Appendix B. A more severe problem about the definition (29) is that it is not enough to define the deformation. For example if want to obtain the second-order correction of (LABEL:HT001) we need to first introduce the Lagrangian density

=tϕΠ+tϕΠ,\begin{split}\mathcal{L}=\partial_{t}\phi\Pi+\partial_{t}^{\ast}\phi\Pi^{\ast}-\mathcal{H},\end{split} (38)

with the primary constraints

χ1=Πi2ϕ=0,χ2=Π+i2ϕ=0,\begin{split}\chi_{1}=\Pi-\frac{i}{2}\phi^{\ast}=0,\quad\chi_{2}=\Pi^{\ast}+\frac{i}{2}\phi=0,\end{split} (39)

then derive the first-order canonical stress tensor through

Tba=(aϕ)aϕ+(aϕ)aϕδba,\begin{split}T^{a}_{b}=\frac{\partial\mathcal{L}}{\partial(\partial_{a}\phi)}\partial_{a}\phi+\frac{\partial\mathcal{L}}{\partial(\partial_{a}\phi^{\ast})}\partial_{a}\phi^{\ast}-\delta^{a}_{b}\mathcal{L},\end{split} (40)

and in the end substitute it into (29). Therefore besides the flow equation (29) one also needs to know how the constraints χ1(λ)\chi_{1}^{(\lambda)} and χ2(λ)\chi_{2}^{(\lambda)} change under the flow.

Assuming that we use (29) together with χ1(λ)\chi_{1}^{(\lambda)} and χ2(λ)\chi_{2}^{(\lambda)} to define the TT¯T\bar{T}-deformation of the non-relativistic free boson gas, we can use the Hellmann-Feynman theorem and factorization formula to derive the flow equation of the spectrum

dn|H|ndλ=n|T00|nn|T11|nn|T01|nn|T10|n.\begin{split}\frac{d\langle n|H|n\rangle}{d\lambda}=\langle n|T_{00}|n\rangle\langle n|T_{11}|n\rangle-\langle n|T_{01}|n\rangle\langle n|T_{10}|n\rangle.\end{split} (41)

But because the mismatch of (34) and (LABEL:HT001) we can not identify n|T00|n\langle n|T_{00}|n\rangle with En/RE_{n}/R and n|H|n=En\langle n|H|n\rangle=E_{n} at the same time if we use (LABEL:HT001) as the Hamiltonian density, where RR is the size of the system.

It is not hard to check that the two constraints (39) are the second-class constraints so one may wonder how about to solve them to reduce the dimensions of the phase space first and then perform the TT¯T\bar{T}-deformation. The reduced Hamiltonian density is given in Gergely:2003zy

r=i2xϕ~xΠ~,ϕ~=ϕ2+iΠ,Π~=Π+iϕ2.\begin{split}\mathcal{H}_{r}=-\frac{i}{2}\partial_{x}\tilde{\phi}\partial_{x}\tilde{\Pi},\quad\tilde{\phi}=\frac{\phi}{2}+i\Pi^{\ast},\quad\tilde{\Pi}=\Pi+i\frac{\phi^{\ast}}{2}.\end{split} (42)

However this does not solve the problem: we still can not transform the Hamiltonian density into a Lagrangian density which only depends on ϕ~\tilde{\phi} and tϕ~\partial_{t}\tilde{\phi} so that the TT¯T\bar{T}-deformation can not be defined. A caveat of our analysis is that there may exist other ways to describe the reduced Hamiltonian system such that the velocity tϕ\partial_{t}\phi could be solved in terms of the momentum.

The singular Lagrangian density can be cured by including the higher-order deformations. In particular the exact Lagrangian is not singular so in principle we can use the Legendre transformation to obtain a Hamiltonian density without any constraints. But it turns out the resulting Hamiltonian has a singularity at λ=0\lambda=0, indicating that it can not be expanded in a Taylor series of the flow parameter λ\lambda. Requiring the vanishing of these singularities introduces the constraints. Since there are different ways to introduce the constraints, there will be ambiguities in the constrained Hamiltonian. In short, the mismatch (34) and (LABEL:HT001) is not just a problem in the first-order deformation, instead it is a problem of the definition of the TT¯T\bar{T}-deformation in a system with singular Lagrangian density or equivalently in a constrained Hamiltonian system.

We conclude this section by stressing that the above argument is purely classical, without considering the problem of Feynman path integral and the ambiguities in the operator ordering.

III.2 Dirac-Bergmann Algorithm

The standard technique to solve the constrained Hamiltonian system is through the Dirac-Bergman Algorithm (DBA). In this section, we will focus on the first-order deformation and study how the deformation affects the commutation relations using DBA. First let us demonstrate the algorithm in the free theory with the action (22). The conjugate momentum of ϕ\phi and ϕ\phi^{\ast} are

Π=i2ϕ,Π=i2ϕ,\begin{split}\Pi=\frac{i}{2}\phi^{\ast},\quad\Pi^{\ast}=-\frac{i}{2}\phi,\end{split} (43)

satisfying the standard equal-time commutation relation

[ϕ(x),Π(y)]=iδ(xy),[ϕ(x),Π(y)]=iδ(xy).\begin{split}[\phi(x),\Pi(y)]=i\delta(x-y),\quad[\phi^{\ast}(x),\Pi^{\ast}(y)]=i\delta(x-y).\end{split} (44)

The Hamiltonian system has two constraints (39) satisfying

[χ1,χ2]C1x,2y=δ(x,y).\begin{split}[\chi_{1},\chi_{2}]\equiv C_{1_{x},2_{y}}=\delta(x,y).\end{split} (45)

These two constraints are the second class, so we can introduce the Dirac bracket which is defined by

[ϕ(x),ϕ(y)]D=[ϕ(x),ϕ(y)]𝑑x𝑑y[ϕ(x),χa(x)]Cax,by1[χb(y),ϕ(y)]=0𝑑x𝑑y(iδ(xx)(δ(xy))(iδ(yy)))=δ(xy)\begin{split}&[\phi(x),\phi^{\ast}(y)]_{D}\\ &=[\phi(x),\phi^{\ast}(y)]-\int dx^{\prime}dy^{\prime}[\phi(x),\chi_{a}(x^{\prime})]C_{a_{x}^{\prime},b_{y}^{\prime}}^{-1}[\chi_{b}(y^{\prime}),\phi^{\ast}(y)]\\ &=0-\int dx^{\prime}dy^{\prime}\left(i\delta(x-x^{\prime})(-\delta(x^{\prime}-y^{\prime}))(-i\delta(y^{\prime}-y))\right)\\ &=\delta(x-y)\end{split} (46)

as expected, recalling that the inverse of the matrix Ca,b1C^{-1}_{a,b} is defined as

𝑑xCax,bxCbx,cy1=δacδ(xy).\begin{split}\int dx^{\prime}C_{a_{x},b_{x}^{\prime}}C^{-1}_{b_{x}^{\prime},c_{y}}=\delta_{ac}\delta(x-y).\end{split} (47)

Now we consider the first-order deformed theory with the action

=0+λ1,1=(xϕϕxϕ)2+12(iϕtϕ(xϕ)2iϕtϕ(xϕ)2).\begin{split}&\mathcal{L}=\mathcal{L}_{0}+\lambda\mathcal{L}_{1},\\ &\mathcal{L}_{1}=(\partial_{x}\phi^{\ast}\phi_{x}\phi)^{2}+\frac{1}{2}\left(i\phi\partial_{t}\phi(\partial_{x}\phi^{\ast})^{2}-i\phi^{\ast}\partial_{t}\phi^{\ast}(\partial_{x}\phi)^{2}\right).\end{split} (48)

Correspondingly the two constraints become

χ1=Πi2ϕiλ2ϕ(xϕ)2,χ2=Π+i2ϕ+iλ2(ϕ(xϕ)2).\begin{split}&\chi_{1}=\Pi-\frac{i}{2}\phi^{\ast}-\frac{i\lambda}{2}\phi(\partial_{x}\phi^{\ast})^{2},\\ &\chi_{2}=\Pi^{\ast}+\frac{i}{2}\phi+\frac{i\lambda}{2}(\phi^{\ast}(\partial_{x}\phi)^{2}).\end{split} (49)

From them we can compute

[χ1(x),χ2(y)]C1x,2y=C2y,1x=δ(xy)+λG(x,y).\begin{split}[\chi_{1}(x),\chi_{2}(y)]&\equiv C_{1_{x},2_{y}}=-C_{2_{y},1_{x}}\\ &=\delta(x-y)+\lambda G(x,y).\end{split} (50)

where

G(x,y)=yδ(xy)ϕ(y)yϕ(y)+xδ(xy)xϕ(x)ϕ(x).\begin{split}G(x,y)=\partial_{y}\delta(x-y)\phi^{\ast}(y)\partial_{y}\phi(y)+\partial_{x}\delta(x-y)\partial_{x}\phi^{\ast}(x)\phi(x).\end{split} (51)

The inverse of the matrix C1x,2yC_{1_{x},2_{y}} up to the first order of λ\lambda is simply

C1x,2y1=δ(xy)+λG(y,x).\begin{split}C_{1_{x},2_{y}}^{-1}=-\delta(x-y)+\lambda G(y,x).\end{split} (52)

Then at the first order of λ\lambda the non-vanishing commutation relations between ϕ\phi and ϕ\phi^{\ast} is

[ϕ(x),ϕ(y)]=0C1x,2y1=δ(xy)λG(y,x).\begin{split}[\phi(x),\phi^{\ast}(y)]=0-C_{1_{x},2_{y}}^{-1}=\delta(x-y)-\lambda G(y,x).\end{split} (53)

This implies that ϕ\phi and ϕ\phi^{\ast} are not canonical variables so that ϕ\phi^{\ast} is not the creation operator to generate the Fock space. This fact is a reflection of the fact that TT¯T\bar{T}-deformation is equivalent to a dynamical coordinate transformation.

At the leading order we can introduce the canonical variables

φ=ϕ+λ2ϕ(xϕ)2,φ=ϕ+λ2(xϕ)2ϕ,\begin{split}\varphi=\phi+\frac{\lambda}{2}\phi^{\ast}(\partial_{x}\phi)^{2},\quad\varphi^{\ast}=\phi^{\ast}+\frac{\lambda}{2}(\partial_{x}\phi^{\ast})^{2}\phi,\end{split} (54)

and conversely

ϕ=φλ2φ(xφ)2,ϕ=φλ2φ(xφ)2,\begin{split}\phi=\varphi-\frac{\lambda}{2}\varphi^{\ast}(\partial_{x}\varphi)^{2},\quad\phi^{\ast}=\varphi^{\ast}-\frac{\lambda}{2}\varphi(\partial_{x}\varphi^{\ast})^{2},\end{split} (55)

such that

x1x2|x3x4=φ1φ2φ3φ4=δ(x13)δ(x24)+δ(x14)δ(x23).\begin{split}\langle x_{1}x_{2}|x_{3}x_{4}\rangle=\langle\varphi_{1}\varphi_{2}\varphi_{3}^{\ast}\varphi_{4}^{\ast}\rangle=\delta(x_{13})\delta(x_{24})+\delta(x_{14})\delta(x_{23}).\end{split} (56)

The relations (54) are not canonical transformation and they are not the state flow which was introduced in Kruthoff:2020hsi . In terms of the new canonical variables the Hamiltonian density (34) up to the first order reads

T00=H0+H1+𝒪(λ2)\begin{split}T_{00}=\mathcal{H}\equiv H_{0}+H_{1}+\mathcal{O}(\lambda^{2})\end{split} (57)

with

H0=xφxφH1=2λ(xφxφ)2λ(φxφxφx2φ+φxφxφx2φ).\begin{split}H_{0}=&\partial_{x}\varphi\partial_{x}\varphi^{\ast}\\ H_{1}=&-2\lambda(\partial_{x}\varphi\partial_{x}\varphi^{\ast})^{2}\\ &-\lambda\left(\varphi\partial_{x}\varphi\partial_{x}\varphi^{\ast}\partial_{x}^{2}\varphi^{\ast}+\varphi^{\ast}\partial_{x}\varphi^{\ast}\partial_{x}\varphi\partial_{x}^{2}\varphi\right).\end{split} (58)

As a consistency check we also work out the first-order deformation of the number operator and momentum operator

N=i2(tϕϕtϕϕ)=ϕϕ+λ2(ϕ2(xϕ)2+(xϕ)2ϕ2)=φφ+𝒪(λ2),\begin{split}N&=-\frac{i}{2}\left(\frac{\partial\mathcal{L}}{\partial\partial_{t}\phi}\phi-\frac{\partial\mathcal{L}}{\partial\partial_{t}\phi^{\ast}}\phi^{\ast}\right)\\ &=\phi^{\ast}\phi+\frac{\lambda}{2}\left({\phi^{\ast}}^{2}(\partial_{x}\phi)^{2}+(\partial_{x}\phi^{\ast})^{2}\phi^{2}\right)\\ &=\varphi^{\ast}\varphi+\mathcal{O}(\lambda^{2}),\end{split} (59)
P=tϕxϕ+tϕxϕ=i2(xϕϕxϕϕ)(1λxϕxϕ),=i2(xφφxφφ)+total derivatives+𝒪(λ2)\begin{split}P&=\frac{\partial\mathcal{L}}{\partial\partial_{t}\phi}\partial_{x}\phi+\frac{\partial\mathcal{L}}{\partial\partial_{t}\phi^{\ast}}\partial_{x}\phi^{\ast}\\ &=\frac{i}{2}(\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi)(1-\lambda\,\partial_{x}\phi\partial_{x}\phi^{\ast}),\quad\\ &=\frac{i}{2}(\partial_{x}\varphi\varphi^{\ast}-\partial_{x}\varphi^{\ast}\varphi)+\text{total derivatives}+\mathcal{O}(\lambda^{2})\end{split} (60)

which are indeed undeformed as expected. We expect that the spectrum of (57) can be derived by performing the second quantization and applying the standard quantum perturbation theory.

III.3 Two-particle sector without momentum

The explicit form of the Hamiltonian (57) implies that there is only two-to-two scattering, so without losing any generality we can focus on the two particle sector. When the total momentum is zero, the TT¯T\bar{T}-deformed spectrum satisfies a very simple equation BoseJ

EN(R,λ)=EN(R+λEN,0).\begin{split}E_{N}(R,\lambda)=E_{N}(R+\lambda E_{N},0).\end{split} (61)

To illustrate the procedure of our calculation, we start from the two-particle sector with zero momentum, u1=u2=u>0u_{1}=-u_{2}=u>0. In this case, a general two-particle eigenfunction is

ψ=2Rcos(u(x1x2)),\begin{split}\psi=\frac{\sqrt{2}}{R}\cos\left(u(x_{1}-x_{2})\right),\end{split} (62)

with

uI=2πRI,I𝐙+.\begin{split}u_{I}=\frac{2\pi}{R}I,\quad I\in\mathbf{Z}^{+}.\end{split} (63)

First we compute the matrix elements of the normal ordered Hamiltonian (57) of the field theory in the position space

y|H0+H1|x.\begin{split}\langle\vec{y}|H_{0}+H_{1}|\vec{x}\rangle.\end{split} (64)

For example the undeformed Hamiltonian density 0\mathcal{H}_{0} is evaluated to be

y|0|x=0|φ(y1)φ(y2)xφxφφ(x1)φ(x2)|0=xδ(xx1)xδ(xy2)δ(y1x2)+xδ(xx2)xδ(xy2)δ(y1x1)+xδ(xx1)xδ(xy1)δ(y2x2)+xδ(xx2)xδ(xy1)δ(y2x1).\begin{split}\langle\vec{y}|\mathcal{H}_{0}|\vec{x}\rangle&=\langle 0|\varphi(y_{1})\varphi(y_{2})\partial_{x}\varphi^{\ast}\partial_{x}\varphi\varphi^{\ast}(x_{1})\varphi^{\ast}(x_{2})|0\rangle\\ &=\partial_{x}\delta(x-x_{1})\partial_{x}\delta(x-y_{2})\delta(y_{1}-x_{2})\\ &+\partial_{x}\delta(x-x_{2})\partial_{x}\delta(x-y_{2})\delta(y_{1}-x_{1})\\ &+\partial_{x}\delta(x-x_{1})\partial_{x}\delta(x-y_{1})\delta(y_{2}-x_{2})\\ &+\partial_{x}\delta(x-x_{2})\partial_{x}\delta(x-y_{1})\delta(y_{2}-x_{1}).\\ \end{split} (65)

Then we use the wavefunction (26) to obtain the Hamiltonian

H0IIuI|H0|uI=12𝑑y𝑑x𝑑xψI(y)ψI(x)y|H0|x=𝑑xψI(x)(x12x22)ψI(x)=2uI2γI2,\begin{split}H_{0II}&\equiv\langle u_{I}|H_{0}|u_{I}\rangle=\frac{1}{2}\int d\vec{y\,}d\vec{x}\,dx\,\psi_{I}^{\ast}(\vec{y})\psi_{I}(\vec{x})\langle\vec{y}|H_{0}|\vec{x}\rangle\\ &=\int d\vec{x}\,\psi^{\ast}_{I}(\vec{x})(-\partial^{2}_{x_{1}}-\partial^{2}_{x_{2}})\psi_{I}(\vec{x})\\ &=2u_{I}^{2}\equiv\gamma I^{2},\end{split} (66)

where γ=8π2R2\gamma=\frac{8\pi^{2}}{R^{2}}. Similarly we find

H1IJ=λ2(2R)2(8uI2uJ2R))=2λγ2I2J2R.\begin{split}H_{1IJ}=-\frac{\lambda}{2}(\frac{\sqrt{2}}{R})^{2}\left(8u_{I}^{2}u_{J}^{2}R)\right)=-2\lambda\frac{\gamma^{2}I^{2}J^{2}}{R}.\end{split} (67)

The result H1II=2λ/RH0II2H_{1II}=-2\lambda/RH_{0II}^{2} coincides with the subleading term in the λ\lambda expansion of the exact spectrum:

EN(R,λ)=EN(0)2λREN(0)2+7λ2R2EN(0)330λ3R3EN(0)4+143λ4R4EN(0)5728λ5R5EN(0)6\begin{split}&E_{N}(R,\lambda)=E_{N}^{(0)}-\frac{2\lambda}{R}{E_{N}^{(0)}}^{2}+\frac{7\lambda^{2}}{R^{2}}{E_{N}^{(0)}}^{3}-\frac{30\lambda^{3}}{R^{3}}{E_{N}^{(0)}}^{4}\\ &+\frac{143\lambda^{4}}{R^{4}}{E_{N}^{(0)}}^{5}-\frac{728\lambda^{5}}{R^{5}}{E_{N}^{(0)}}^{6}\dots\end{split} (68)

which is solved from the flow equation in BoseJ .

Rayleigh-Schrödinger perturbation

The higher-order perturbative results can be obtained by using the Rayleigh-Schrödinger perturbation:

EI(2)=I4(2λ)2γ3R2gI(1),EI(3)=I4(2λ)3γ4R3(gI(1)2gI(2))EI(4)=I4(2λ)4γ5R4(gI(1)33gI(1)gI(2)+gI(3))EI(5)=I4(2λ)5γ6R5(gI(1)46gI(1)2gI(2)+2gI(2)2+4gI(1)gI(3)gI(4)).\begin{split}E^{(2)}_{I}=&I^{4}(-2\lambda)^{2}\frac{\gamma^{3}}{R^{2}}g_{I}(1),\\ E^{(3)}_{I}=&I^{4}(-2\lambda)^{3}\frac{\gamma^{4}}{R^{3}}\left(g_{I}(1)^{2}-g_{I}(2)\right)\\ E^{(4)}_{I}=&I^{4}(-2\lambda)^{4}\frac{\gamma^{5}}{R^{4}}\left(g_{I}(1)^{3}-3g_{I}(1)g_{I}(2)+g_{I}(3)\right)\\ E^{(5)}_{I}=&I^{4}(-2\lambda)^{5}\frac{\gamma^{6}}{R^{5}}\left(g_{I}(1)^{4}-6g_{I}(1)^{2}g_{I}(2)\right.\\ &\left.+2g_{I}(2)^{2}+4g_{I}(1)g_{I}(3)-g_{I}(4)\right)\\ &....\end{split}

where for convenience we have introduced the function

gI(a)JII4(a1)J4(I2J2)a.\begin{split}g_{I}(a)\equiv\sum_{J\neq I}^{\infty}\frac{I^{4(a-1)}J^{4}}{(I^{2}-J^{2})^{a}}.\end{split} (69)

The functions gI(1)g_{I}(1) and gI(2)g_{I}(2) are divergent, and proper regularization has to be adopted. The first few terms of gI(a)g_{I}(a) are given by

gI(1)=7I24,gI(2)=11I416+π2I612,gI(3)=I6325π2I848,gI(4)=3I8256+π2I1096+π4I12720,.\begin{split}&g_{I}(1)=\frac{7I^{2}}{4},\quad g_{I}(2)=-\frac{11I^{4}}{16}+\frac{\pi^{2}I^{6}}{12},\\ &g_{I}(3)=-\frac{I^{6}}{32}-\frac{5\pi^{2}I^{8}}{48},\\ &g_{I}(4)=-\frac{3I^{8}}{256}+\frac{\pi^{2}I^{10}}{96}+\frac{\pi^{4}I^{12}}{720},....\end{split} (70)

where we have used the Dirichlet regularization to regularize gI(1)g_{I}(1) and gI(2)g_{I}(2). For example, to compute g(1)g(1) we first separate it into the divergent piece and convergent piece

gI(1)=JI([I2J2]+I4I2J2)\begin{split}g_{I}(1)=\sum_{J\neq I}^{\infty}\left([-I^{2}-J^{2}]+\frac{I^{4}}{I^{2}-J^{2}}\right)\end{split} (71)

then regularize the divergent summation with the Dirichlet regularization. Substituting the regularized functions into (III), we obtain the higher-order quantum corrections to the spectrum:

EI(2)=I6λ2γ3R27,EI(3)=I8(λ)3γ4R3(302π2I23),EI(4)=I10(λ)4γ5R4(14326π2I23),EI(5)=I12(λ)5γ6R5(7282π4I4400π2I25),.\begin{split}&E_{I}^{(2)}=I^{6}\lambda^{2}\frac{\gamma^{3}}{R^{2}}7,\quad E_{I}^{(3)}=I^{8}(-\lambda)^{3}\frac{\gamma^{4}}{R^{3}}\left(30-\frac{2\pi^{2}I^{2}}{3}\right),\\ &E^{(4)}_{I}=I^{10}(-\lambda)^{4}\frac{\gamma^{5}}{R^{4}}\left({143}-\frac{26\pi^{2}I^{2}}{3}\right),\\ &E^{(5)}_{I}=I^{12}(-\lambda)^{5}\frac{\gamma^{6}}{R^{5}}\left(728-\frac{2\pi^{4}I^{4}-400\pi^{2}I^{2}}{5}\right),....\end{split} (72)

Surprisingly, these expressions are in exact match with (LABEL:Exact) if all the transcendental terms are discarded. It was shown in Zamolodchikov:1991vx ; Caselle:2013dra that the appearance of the transcendental terms is the signal of the contributions from other irrelevant operators instead of TT¯T\bar{T}, so they should be discarded. If we follow this rule, we can use the Brillouin-Wigner perturbation to obtain the exact spectrum which coincides with the one solved from the flow equation, as we show below.

Exact form: Brillouin-Wigner perturbation

A technical introduction to the Brillouin-Wigner perturbation can be found in Appendix C. The exact energy satisfies

Ek=Ek(0)+i=1j=0i1H1mjmj+1j=1i1(EkEmj(0)),m0=mi=k.\begin{split}E_{k}=E_{k}^{(0)}+\sum_{i=1}^{\infty}\frac{\prod_{j=0}^{i-1}H_{1m_{j}m_{j+1}}}{\prod_{j=1}^{i-1}(E_{k}-E^{(0)}_{m_{j}})},\quad m_{0}=m_{i}=k.\end{split} (73)

In our case, En(0)=γn2,H1nm=2λγ2Rn2m2E_{n}^{(0)}=\gamma n^{2},\ \ H_{1nm}=-2\lambda\frac{\gamma^{2}}{R}n^{2}m^{2} so that

Ek=γk2+γk4i=1(2γλR)ig(k,bk)i1\begin{split}E_{k}=\gamma k^{2}+\gamma k^{4}\sum_{i=1}^{\infty}\left(-\frac{2\gamma\lambda}{R}\right)^{i}g(k,b_{k})^{i-1}\end{split} (74)

where bk=Ek/γb_{k}=E_{k}/\gamma and

g(k,bk)=mk,m=1m4bkm2=k4k2bk+12bk3/2πcot(πbk).\begin{split}g(k,b_{k})&=\sum_{m\neq k,m=1}^{\infty}\frac{m^{4}}{b_{k}-m^{2}}=\frac{k^{4}}{k^{2}-b_{k}}+\frac{1}{2}b_{k}^{3/2}\pi\cot(\pi\sqrt{b_{k}}).\end{split} (75)

The equation of the energy can be recast into the following form

bk=k2βk41+β(k4k2bk+12bk3/2πcot(πbk))\begin{split}&b_{k}=k^{2}-\frac{\beta k^{4}}{1+\beta\left(\frac{k^{4}}{k^{2}-b_{k}}+\frac{1}{2}b_{k}^{3/2}\pi\cot(\pi\sqrt{b_{k}})\right)}\\ \end{split} (76)

where β=γλR\beta=\frac{\gamma\lambda}{R}. As in the last subsection, if we discard the π\pi-dependent terms this result will match the exact result found in BoseJ . Notice that we must do this carefully. Naively, it seems necessary to throw away the term πcot(πbk)\pi\cot(\pi\sqrt{b_{k}}) completely. However, because of bk=k2+O(λ)b_{k}=k^{2}+O(\lambda), we can expand the function and find

πcot(πbk)=πcot(π(bkk))=1bkk+(π-dependent terms).\begin{split}&\pi\cot(\pi\sqrt{b_{k}})=\pi\cot(\pi(\sqrt{b_{k}}-k))\\ &=\frac{1}{\sqrt{b_{k}}-k}+(\mbox{$\pi$-dependent terms}).\end{split} (77)

Then the equation of the energy becomes

bkk2=1βk21+βk2(11bk/k2+12(bk/k2)3/2bk/k21).\begin{split}\frac{b_{k}}{k^{2}}=1-\frac{\beta k^{2}}{1+\beta k^{2}\left(\frac{1}{1-b_{k}/k^{2}}+\frac{1}{2}\frac{(b_{k}/k^{2})^{3/2}}{\sqrt{b_{k}/k^{2}}-1}\right)}.\\ \end{split} (78)

This is an algebraic equation, and is equivalent to

β~kb~k3/2+b~k1=0\begin{split}\tilde{\beta}_{k}\tilde{b}_{k}^{3/2}+\sqrt{\tilde{b}_{k}}-1=0\end{split} (79)

where

β~k=βk2/2=γλk22R,b~k=Ekγk2.\begin{split}\tilde{\beta}_{k}=\beta k^{2}/2=\frac{\gamma\lambda k^{2}}{2R},\hskip 12.91663pt\tilde{b}_{k}=\frac{E_{k}}{\gamma k^{2}}.\end{split} (80)

Multiplying (79) by β~kb~k3/2+b~k+1\tilde{\beta}_{k}\tilde{b}_{k}^{3/2}+\sqrt{\tilde{b}_{k}}+1, we get

β~k2b~k3+2β~kb~k2+b~k1=0,\begin{split}\tilde{\beta}_{k}^{2}\tilde{b}_{k}^{3}+2\tilde{\beta}_{k}\tilde{b}_{k}^{2}+\tilde{b}_{k}-1=0,\end{split} (81)

which is the Burgers’ equation in BoseJ . A solution of (79) is simply

b~k=23β~k[cosh(23arcsinh(33β~k2))1].\begin{split}\tilde{b}_{k}=\frac{2}{3\tilde{\beta}_{k}}\left[\cosh\left(\frac{2}{3}{\rm arcsinh}\left(\frac{3\sqrt{3\tilde{\beta}_{k}}}{2}\right)\right)-1\right].\end{split} (82)

Expanding the above relation, we can reproduce the perturbative results in the last subsection.

III.4 Two-particle sector with momentum

In this subsection, we consider the two-particle sector with momentum. The general two-particle eigenfunction is

ψ=12R(eiu1x1+iu2x2+eiu1x2+iu2x1),\begin{split}\psi=\frac{1}{\sqrt{2}R}\left(e^{iu_{1}x_{1}+iu_{2}x_{2}}+e^{iu_{1}x_{2}+iu_{2}x_{1}}\right),\end{split} (83)

with zero-order eigenvalue

Eu1,u2(0)=u12+u22.\begin{split}E^{(0)}_{u_{1},u_{2}}=u_{1}^{2}+u_{2}^{2}.\end{split} (84)

When u1±u2u_{1}\neq\pm u_{2}, the eigenvalues have a four-fold degeneracy Eu1,u2(0)=Eu1,u2(0)=Eu1,u2(0)=Eu1,u2(0)E^{(0)}_{u_{1},u_{2}}=E^{(0)}_{-u_{1},u_{2}}=E^{(0)}_{u_{1},-u_{2}}=E^{(0)}_{-u_{1},-u_{2}}. When u1=u2u_{1}=u_{2}, the degeneracy is two-fold: Eu1,u2(0)=Eu1,u2(0)E^{(0)}_{u_{1},u_{2}}=E^{(0)}_{-u_{1},-u_{2}}. The degenerate spectrum is an obstacle to apply the flow equation (41) to derive the deformed spectrum because the factorization formula fails. However in this case, because the total momentum is conserved so one can focus on a sector with some fixed total momentum u=u1+u2u=u_{1}+u_{2}. Within this sector the spectrum will not be degenerate anymore and the flow equation still works. The flow equation can be solved by a formal series expansion in λ\lambda. The first few terms of the deformed spectrum are given by555There is a typo in equation (122)(122) of BoseJ : EN(1)=1R(2M22+2M1M3)E_{N}^{(1)}=\frac{1}{R}(-2M_{2}^{2}+2M_{1}M_{3}).

u1,u2(1)=λR2u1u2(u1u2)2,u1,u2(2)=λ2R22u1u2(u1u2)2(u125u1u2+u22),u1,u2(3)=λ3R32u1u2(u1u2)2(u1210u1u2+u22)(u123u1u2+u22),\begin{split}&\mathcal{E}_{u_{1},u_{2}}^{(1)}=\frac{\lambda}{R}2u_{1}u_{2}(u_{1}-u_{2})^{2},\\ &\mathcal{E}_{u_{1},u_{2}}^{(2)}=-\frac{\lambda^{2}}{R^{2}}2u_{1}u_{2}(u_{1}-u_{2})^{2}(u_{1}^{2}-5u_{1}u_{2}+u_{2}^{2}),\\ &\mathcal{E}_{u_{1},u_{2}}^{(3)}=\frac{\lambda^{3}}{R^{3}}2u_{1}u_{2}(u_{1}-u_{2})^{2}(u_{1}^{2}-10u_{1}u_{2}+u_{2}^{2})\\ &~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}(u_{1}^{2}-3u_{1}u_{2}+u_{2}^{2}),\\ &\dots\end{split} (85)

Below we will try to compute the deformed spectrum within quantum perturbation theory. In the basis of the wavefunction (83), we find

u1,u2|H1|u3,u4\displaystyle\langle u_{1},u_{2}|H_{1}|u_{3},u_{4}\rangle (86)
=\displaystyle= λRδ(u1+u2u3u4)\displaystyle-\frac{\lambda}{R}\delta(u_{1}+u_{2}-u_{3}-u_{4})
×(8u1u2u3u4(u1+u2)2(u1u2+u3u4)),\displaystyle\times\left(8u_{1}u_{2}u_{3}u_{4}-(u_{1}+u_{2})^{2}(u_{1}u_{2}+u_{3}u_{4})\right),

where the Dirac delta function is due to the momentum conservation. In particular we have

Eu1,u2(1)=u1,u2|H1|u1,u2=λR2u1u2(u1u2)2=u1,u2(1).\begin{split}E_{u_{1},u_{2}}^{(1)}=\langle u_{1},u_{2}|H_{1}|u_{1},u_{2}\rangle=\frac{\lambda}{R}2u_{1}u_{2}(u_{1}-u_{2})^{2}=\mathcal{E}_{u_{1},u_{2}}^{(1)}.\end{split} (87)

The second-order correction is given by a divergent summation666Without losing generality we have assumed u1u2u_{1}\geq u_{2}.

Eu1,u2(2)=u3,u4|u1u2|H1|u3,u4|2|u12+u22u32u42=λ2R2u3u1,u3=8u1u2(u1+u2u3)u3+(u1+u2)2(u1u2+(u1+u2u3)u3)22(u1u3)(u3u2).\begin{split}E_{u_{1},u_{2}}^{(2)}&=\sum_{u_{3},u_{4}}\frac{|\langle u_{1}u_{2}|H_{1}|u_{3},u_{4}\rangle|^{2}|}{u_{1}^{2}+u_{2}^{2}-u_{3}^{2}-u_{4}^{2}}\\ &=\frac{\lambda^{2}}{R^{2}}\sum_{u_{3}\neq u_{1},u_{3}=-\infty}^{\infty}\frac{-8u_{1}u_{2}(u_{1}+u_{2}-u_{3})u_{3}+(u_{1}+u_{2})^{2}(u_{1}u_{2}+(u_{1}+u_{2}-u_{3})u_{3})^{2}}{2(u_{1}-u_{3})(u_{3}-u_{2})}.\end{split} (88)

As before to regularize it we first separate out the divergent piece by a 1/u31/u_{3} Taylor expansion

(88)=u3=[D0(u1,u2)+D1(u1,u2)u3+D2(u1,u2)u32](D0(u1,u2)+D1(u1,u2)u1+D2(u1,u2)u12)+u3u1,u3=C(u1,u2,u3),\begin{split}\eqref{TwoMoH2}=&\sum_{u_{3}=-\infty}^{\infty}[D_{0}(u_{1},u_{2})+D_{1}(u_{1},u_{2})u_{3}+D_{2}(u_{1},u_{2})u_{3}^{2}]\\ &-(D_{0}(u_{1},u_{2})+D_{1}(u_{1},u_{2})u_{1}+D_{2}(u_{1},u_{2})u_{1}^{2})\\ &+\sum_{u_{3}\neq u_{1},u_{3}=-\infty}^{\infty}C(u_{1},u_{2},u_{3}),\end{split} (89)

with

C(u1,u2,u3)=2u12(u1u2)4u22(u1u3)(u3u2).\begin{split}&C(u_{1},u_{2},u_{3})=\frac{2u_{1}^{2}(u_{1}-u_{2})^{4}u_{2}^{2}}{(u_{1}-u_{3})(u_{3}-u_{2})}.\end{split} (90)

Then using the Dirichlet regularization

u3=1=u3=u3=u3=u32=0,\begin{split}\sum_{u_{3}=-\infty}^{\infty}1=\sum_{u_{3}=-\infty}^{\infty}u_{3}=\sum_{u_{3}=-\infty}^{\infty}u_{3}^{2}=0,\end{split} (91)

we find the final result of Eu1,u2(2)E_{u_{1},u_{2}}^{(2)}

u1,u2(2)=λ2R2(0(2u1(u1u2)2u2(u126u1u2+u22))+2u12(u1u2)2u22).\begin{split}\mathcal{E}_{u_{1},u_{2}}^{(2)}=&\frac{\lambda^{2}}{R^{2}}\left(0-\left(2u_{1}(u_{1}-u_{2})^{2}u_{2}(u_{1}^{2}-6u_{1}u_{2}+u_{2}^{2})\right)\right.\\ &\left.+2u_{1}^{2}(u_{1}-u_{2})^{2}u_{2}^{2}\right).\end{split} (92)

Similarly, we can read the third-order correction

Eu1,u2(3)=u1,u2(3)π22λ33R3u13u23(u1u2)4,\begin{split}E_{u_{1},u_{2}}^{(3)}=\mathcal{E}_{u_{1},u_{2}}^{(3)}-\pi^{2}\frac{2\lambda^{3}}{3R^{3}}u_{1}^{3}u_{2}^{3}(u_{1}-u_{2})^{4},\end{split} (93)

which reduces to (LABEL:TwoSpectrum) by taking u2=u1u_{2}=-u_{1}. The calculation of higher-order corrections is much involved, but the feature is very similar so we will not present the analysis here.

III.5 Three-particle sector

To analyze the spectrum of the three-particle sector one has to include the second order deformation. However as we discuss before, after including the second order deformation the theory either does not admit a perturbative description or has ambiguities. The detail of this analysis can be found in Appendix D. In this subsection we only compute the spectrum up to the first-order deformation. The eigenfunction of the free three-particle theory with I=(I1,I2,I3)I=(I_{1},I_{2},I_{3}) is given by

ψI(x1,x2,x3)=13!R3σSexpij=13xjuIσj\begin{split}\psi_{I}(x_{1},x_{2},x_{3})=\frac{1}{\sqrt{3!R^{3}}}\sum_{\sigma\in S}\exp^{i\sum_{j=1}^{3}x_{j}u_{I\sigma_{j}}}\end{split} (94)

As before one can find the undeformed Hamiltonian in the position space

x1,x2,x3|H0|y1,y2,y3=i,jS𝑑xxδ(xxi1)xδ(xyj1)δ(xi2yj2)δ(xi3yj3)\begin{split}&\langle x_{1},x_{2},x_{3}|H_{0}|y_{1},y_{2},y_{3}\rangle\\ &=\sum_{i,j\in S}\int\,dx\partial_{x}\delta(x-x_{i_{1}})\partial_{x}\delta(x-y_{j_{1}})\delta(x_{i_{2}}-y_{j_{2}})\delta(x_{i_{3}}-y_{j_{3}})\end{split} (95)

and in the momentum space

uI|H0|uJ=(2πR)2Ii1Jj1δJj1Ii1δJj2Ii2δJj3Ii3,\begin{split}\langle u_{I}|H_{0}|u_{J}\rangle=\left(\frac{2\pi}{R}\right)^{2}\sum I_{i_{1}}J_{j_{1}}\delta^{I_{i_{1}}}_{J_{j_{1}}}\delta^{I_{i_{2}}}_{J_{j_{2}}}\delta^{I_{i_{3}}}_{J_{j_{3}}},\end{split} (96)

where SS denotes the permutation of three particles (1,2,3)(1,2,3), and uI1=2πRI1u_{I_{1}}=\frac{2\pi}{R}I_{1}. Therefore the zeroth-order energy is given by777Here we assume that I1>I2>I3I_{1}>I_{2}>I_{3}.

EI(0)=uI|H0|uI=(2πR)2(I12+I22+I32).\begin{split}E_{I}^{(0)}=\langle u_{I}|H_{0}|u_{I}\rangle=\left(\frac{2\pi}{R}\right)^{2}(I_{1}^{2}+I_{2}^{2}+I_{3}^{2}).\end{split} (97)

Similarly, we can get the first-order result,

x1,x2,x3|H1|y1,y2,y3=i,jS𝑑xδ(xxi1)δ(xxi2)δ(xyj1)δ(xyj2)δ(xi3yj3)(2xi1xi2yj1yj2xi1xi22yj1xi1yj1yj22)\begin{split}&\langle x_{1},x_{2},x_{3}|H_{1}|y_{1},y_{2},y_{3}\rangle=\\ &\sum_{i,j\in S}\int\,dx\delta(x-x_{i_{1}})\delta(x-x_{i_{2}})\delta(x-y_{j_{1}})\delta(x-y_{j_{2}})\delta(x_{i_{3}}-y_{j_{3}})\\ &~{}~{}~{}\left(-2\partial_{x_{i_{1}}}\partial_{x_{i_{2}}}\partial_{y_{j_{1}}}\partial_{y_{j_{2}}}-\partial_{x_{i_{1}}}\partial_{x_{i_{2}}}^{2}\partial_{y_{j_{1}}}-\partial_{x_{i_{1}}}\partial_{y_{j_{1}}}\partial_{y_{j_{2}}}^{2}\right)\end{split} (98)

and

uI|H1|uJ=1R(2πR)4i,jSδJi3Ii3(2Ii1Ii2Jj1Jj2+Ii1Ii22Jj1+Ii1Jj1Jj22)\begin{split}&\langle u_{I}|H_{1}|u_{J}\rangle\\ &=\frac{1}{R}\left(\frac{2\pi}{R}\right)^{4}\sum_{i,j\in S}\delta^{I_{i_{3}}}_{J_{i_{3}}}\left(-2I_{i_{1}}I_{i_{2}}J_{j_{1}}J_{j_{2}}+I_{i_{1}}I_{i_{2}}^{2}J_{j_{1}}+I_{i_{1}}J_{j_{1}}J_{j_{2}}^{2}\right)\end{split} (99)

The first order correction of the energy is given by

EI(1)=λuI|H1|uI=λ1R(2πR)4(2I1I2(I1I2)2+2I2I3(I2I3)2+2I3I1(I3I1)2)\begin{split}E^{(1)}_{I}=&\lambda\langle u_{I}|H_{1}|u_{I}\rangle\\ =&\lambda\frac{1}{R}\left(\frac{2\pi}{R}\right)^{4}\left(2I_{1}I_{2}(I_{1}-I_{2})^{2}\right.\\ &\left.+2I_{2}I_{3}(I_{2}-I_{3})^{2}+2I_{3}I_{1}(I_{3}-I_{1})^{2}\right)\end{split} (100)

It is the same as the one in BoseJ .

The second-order correction of the energy should come from uI|H1|uJ\langle u_{I}|H_{1}|u_{J}\rangle and uI|H2|uI\langle u_{I}|H_{2}|u_{I}\rangle

EI(2)=λ2JI|uI|H1|uJ|2EI(0)EJ(0)+λ2uI|H2|uI.\begin{split}E_{I}^{(2)}=\lambda^{2}\sum_{J\neq I}\frac{|\langle u_{I}|H_{1}|u_{J}\rangle|^{2}}{E_{I}^{(0)}-E_{J}^{(0)}}+\lambda^{2}\langle u_{I}|H_{2}|u_{I}\rangle.\end{split} (101)

The first part of EI(2)E_{I}^{(2)} is

JI|uI|H1|uJ|2EI(0)EJ(0)=1R2(2πR)6JI1I12+I22+I32J12J22J32i3,j3δJi3Ii3(8Ii1Ii2Jj1Jj2+(Ii1Ii2+Jj1Jj2)(Ii1+Ii2)(Jj1+Jj2))2\begin{split}&\sum_{J\neq I}\frac{|\langle u_{I}|H_{1}|u_{J}\rangle|^{2}}{E_{I}^{(0)}-E_{J}^{(0)}}\\ &=\frac{1}{R^{2}}\left(\frac{2\pi}{R}\right)^{6}\sum_{J\neq I}\frac{1}{I_{1}^{2}+I_{2}^{2}+I_{3}^{2}-J_{1}^{2}-J_{2}^{2}-J_{3}^{2}}\sum_{i_{3},j_{3}}\delta^{I_{i_{3}}}_{J_{i_{3}}}\left(-8I_{i_{1}}I_{i_{2}}J_{j_{1}}J_{j_{2}}+(I_{i_{1}}I_{i_{2}}+J_{j_{1}}J_{j_{2}})(I_{i_{1}}+I_{i_{2}})(J_{j_{1}}+J_{j_{2}})\right)^{2}\end{split}

Notice that there is no term like δJ1I2δJ2I3\delta_{J_{1}}^{I_{2}}\delta_{J_{2}}^{I_{3}} due to I1+I2+I3=J1+J2+J3I_{1}+I_{2}+I_{3}=J_{1}+J_{2}+J_{3} and JIJ\neq I. After some manipulations, we read

JI|uI|H1|uJ|2EI(0)EJ(0)\displaystyle\sum_{J\neq I}\frac{|\langle u_{I}|H_{1}|u_{J}\rangle|^{2}}{E_{I}^{(0)}-E_{J}^{(0)}}
=\displaystyle= 1R2(2πR)6(2I1I2(I1I2)(I125I1I2+I22)2I1I3(I1I3)(I125I1I3+I32)2I3I2(I3I2)(I325I3I2+I22))\displaystyle\frac{1}{R^{2}}\left(\frac{2\pi}{R}\right)^{6}\left(-2I_{1}I_{2}(I_{1}-I_{2})(I_{1}^{2}-5I_{1}I_{2}+I_{2}^{2})-2I_{1}I_{3}(I_{1}-I_{3})(I_{1}^{2}-5I_{1}I_{3}+I_{3}^{2})-2I_{3}I_{2}(I_{3}-I_{2})(I_{3}^{2}-5I_{3}I_{2}+I_{2}^{2})\right)

Matching EI(2)E_{I}^{(2)} with the result in BoseJ , it implies that

uI|H2|uI=2I1I2I3(5I136I12I26I1I22+5I236I12I3+21I1I2I36I22I36I1I326I2I32+5I33).\begin{split}\langle u_{I}|H_{2}|u_{I}\rangle=&2I_{1}I_{2}I_{3}(5I_{1}^{3}-6I_{1}^{2}I_{2}-6I_{1}I_{2}^{2}+5I_{2}^{3}-6I_{1}^{2}I_{3}\\ &+21I_{1}I_{2}I_{3}-6I_{2}^{2}I_{3}-6I_{1}I_{3}^{2}-6I_{2}I_{3}^{2}+5I_{3}^{3}).\end{split} (103)

IV Conclusion

In this work we have derived a closed form (LABEL:ttbarL) of the Lagrangian of the TT¯T\bar{T}-deformed non-relativistic model by using the perturbative computations and the method of characteristics. In addition, we computed the deformed spectrum by using two different perturbation theories. In particular, by using the Brillouin-Wigner perturbation, we managed to find the exact form of the deformed energy spectrum, which can match the one obtained in BoseJ after some appropriate regularization.

One subtle issue in our study is on the regularization. The perturbative expansion of the deformed energy is a summation of infinite series, which is superficially divergent. We applied the zeta-function regularization and found that the results could match the ones in BoseJ if we discard the π\pi-dependent terms. The similar π\pi-dependent terms, which were referred to as the terms with nonzero transcendentality, appeared also in the study on other models Zamolodchikov:1991vx ; Caselle:2013dra . They could come from the perturbations of higher order of TT¯T\bar{T}, say T2T¯2T^{2}\bar{T}^{2}. It would be certainly interesting to understand why they appear in our perturbative analysis.

Another interesting finding is that the TT¯T\bar{T}-deformation on the Lagrangian and the one on the Hamiltonian is not equivalent if the theory has a degenerate Legendre transformation. The deformation may remove the degeneracy but the resulting deformed theory would not admit a perturbative description. More importantly, we want to stress that to define the TT¯T\bar{T}-deformation in the Hamiltonian formalism one has to also specify how the constraints are deformed.

Acknowledgements

We would like to thank Reiko Liu, Shi-Lin Zhu, Florian Loebbert and Sergey Frolov for valuable discussions and comments. The work is in part supported by NSFC Grants No. 11325522, No. 11735001, No. 11947301 and No. 12047502. J. T. is also supported by the UCAS program of special research associate and by the internal funds of the KITS.

Appendix A Perturbative Lagrangian

Expanding the TT¯T\bar{T}-deformed Lagrangian by λ\lambda, =λnn\mathcal{L}=\sum\lambda^{n}\mathcal{L}_{n}, one can get a recursion relation from the flow equation (1),

i+1=1i+1k=0i[kik+kβikβ+ikαkα(kxα)(ikyβ)+(ikαβ)(kβα)].\begin{split}\mathcal{L}_{i+1}=&\frac{1}{i+1}\sum_{k=0}^{i}\left[-\mathcal{L}_{k}\mathcal{L}_{i-k}+\mathcal{L}_{k}\vec{\beta}\cdot\frac{\partial\mathcal{L}_{i-k}}{\partial\vec{\beta}}+\mathcal{L}_{i-k}\vec{\alpha}\cdot\frac{\partial\mathcal{L}_{k}}{\partial\vec{\alpha}}-\left(\frac{\partial\mathcal{L}_{k}}{\partial\vec{x}}\cdot\vec{\alpha}\right)\left(\frac{\partial\mathcal{L}_{i-k}}{\partial\vec{y}}\cdot\vec{\beta}\right)+\left(\frac{\partial\mathcal{L}_{i-k}}{\partial\vec{\alpha}}\cdot\vec{\beta}\right)\left(\frac{\partial\mathcal{L}_{k}}{\partial\vec{\beta}}\cdot\vec{\alpha}\right)\right].\end{split} (104)

For one complex scalar, by the initial condition,

0=V+X+Y\begin{split}\mathcal{L}_{0}=-V+X+Y\end{split} (105)

one can get the first several k\mathcal{L}_{k}’s,

1=2iA(VXY)(V+Y)2=BV3+V2XVY22iA(X+2Y)+Y(X2+3XY+2Y2)3=4A2V4+V3X+B(X+2Y)VY2(3X+4Y)2iA(X2+6XY2(V3Y)Y)+Y(X3+6X2Y+10XY2+5Y3)4=V5+V4X+2V2Y3+B(X2+6XY+6Y2)+Y(X4+10X3Y+30X2Y2+35XY3+14Y4)+4A2(V3(X+2Y))VY(2B+Y(6X2+20XY+15Y2))2iA(2B+(X+2Y)(X2+10XY+2Y(3V+5Y)))\begin{split}\mathcal{L}_{1}=&-2iA-(V-X-Y)(V+Y)\\ \mathcal{L}_{2}=&B-V^{3}+V^{2}X-VY^{2}-2iA(X+2Y)\\ &+Y(X^{2}+3XY+2Y^{2})\\ \mathcal{L}_{3}=&-4A^{2}-V^{4}+V^{3}X+B(X+2Y)-VY^{2}(3X+4Y)\\ &-2iA(X^{2}+6XY-2(V-3Y)Y)\\ &+Y(X^{3}+6X^{2}Y+10XY^{2}+5Y^{3})\\ \mathcal{L}_{4}=&-V^{5}+V^{4}X+2V^{2}Y^{3}+B(X^{2}+6XY+6Y^{2})\\ &+Y(X^{4}+10X^{3}Y+30X^{2}Y^{2}+35XY^{3}+14Y^{4})\\ &+4A^{2}(V-3(X+2Y))\\ &-VY(2B+Y(6X^{2}+20XY+15Y^{2}))\\ &-2iA(2B+(X+2Y)(X^{2}+10XY+2Y(-3V+5Y)))\end{split} (106)

Appendix B Difference between two deformation formalisms

To illustrate the difference between the deformations in the Lagrangian formalism and the Hamiltonian formalism, let us consider the well studied TT¯T\bar{T}-deformed free scalar theory whose action was found in Cavaglia:2016oda to take Nambu-Goto form

λ=1λ(1+λ(xϕ)2λ(tϕ)21).\begin{split}\mathcal{L}_{\lambda}=-\frac{1}{\lambda}\left(\sqrt{1+\lambda(\partial_{x}\phi)^{2}-\lambda(\partial_{t}\phi)^{2}}-1\right).\end{split} (107)

Taking the derivative with respect to tϕ\partial_{t}\phi we can obtain the conjugate momentum

Π=tϕ1+λ(xϕ)2λ(tϕ)2.\begin{split}\Pi=\frac{\partial_{t}\phi}{\sqrt{1+\lambda(\partial_{x}\phi)^{2}-\lambda(\partial_{t}\phi)^{2}}}.\end{split} (108)

The standard Legendre transformation leads to the deformed Hamiltonian

Hλ=1λ(1+λ(Πλ2+(xϕ)2)+λ2Πλ2(xϕ)21).\begin{split}H_{\lambda}=\frac{1}{\lambda}\left(\sqrt{1+\lambda(\Pi_{\lambda}^{2}+(\partial_{x}\phi)^{2})+\lambda^{2}\Pi_{\lambda}^{2}(\partial_{x}\phi)^{2}}-1\right).\end{split} (109)

Now let us compare the first-order expansion of (107) and (109):

λ=12((tϕ)2(xϕ)2)+18((tϕ)2(xϕ)2)2λ+𝒪(λ2)0+λdet(Tμν),λ=12(Π2+(xϕ)2)18((Π)2(xϕ)2)2λ+𝒪(λ2).\begin{split}\mathcal{L}_{\lambda}=&\frac{1}{2}\left((\partial_{t}\phi)^{2}-(\partial_{x}\phi)^{2}\right)+\frac{1}{8}((\partial_{t}\phi)^{2}-(\partial_{x}\phi)^{2})^{2}\lambda+\mathcal{O}(\lambda^{2})\\ &\sim\mathcal{L}_{0}+\lambda\,\mbox{det}(T_{\mu\nu}),\\ \mathcal{H}_{\lambda}=&\frac{1}{2}\left(\Pi^{2}+(\partial_{x}\phi)^{2}\right)-\frac{1}{8}((\Pi)^{2}-(\partial_{x}\phi)^{2})^{2}\lambda+\mathcal{O}(\lambda^{2}).\end{split} (110)

The first-order correction of HλH_{\lambda} can be written as

λ=0λdet(Tμν)\begin{split}\mathcal{H}_{\lambda}=\mathcal{H}_{0}-\lambda\,\mbox{det}(T_{\mu\nu})\end{split} (111)

if and only if we use the zeroth-order result of Π\Pi in (108). It implies that if one wants to find the first-order expansion of the Hamiltonian from the Lagrangian via Legendre transformation, he has to rewrite the time derivative terms in terms of the conjugate momentum first, then perform the expansion while keep the conjugate momentum. However in the non-relativistic model we are studying, this can not be done. This fact leaves us with two possible choices: one may take (111) as the first order correction to the Hamiltonian or one may take the Legendre transformation of (110) as the first order correction to the Hamiltonian. We have shown explicitly that these two choice give rise to different results.

Appendix C Brillouin-Wigner Perturbation

In the textbookShankar , Rayleigh-Schrödinger perturbation theory has been discussed. However, it is too complicated to go to a higher order. If we want to know the correction of the energy, we have to calculate the energy and wavefunctions simultaneously. And its physical meaning is not clear. In contrast, the Brillouin-Wigner perturbation is easy to calculate to higher orders. One can calculate the energy alone without knowing the wavefunctions. It is the Feynman diagrams in quantum mechanics. One can find a good lecture in Sun:2016 .

For a theory whose Hamiltonian is H=H0+λHH=H_{0}+\lambda H^{\prime}, where λ\lambda is small. The energy of the kk-th eigenstate is given by

En=En(0)+λHnn+λ2HnmHmnEnEm(0)+λ3HnmHmmHmn(EnEm(0))(EnEm(0))+\begin{split}E_{n}=&E_{n}^{(0)}+\lambda H^{\prime}_{\mathrm{nn}}+\lambda^{2}\frac{H^{\prime}_{nm}H^{\prime}_{mn}}{E_{n}-E_{m}^{(0)}}\\ &+\lambda^{3}\frac{H^{\prime}_{nm}H^{\prime}_{mm^{\prime}}H^{\prime}_{m^{\prime}n}}{\left(E_{n}-E_{m}^{(0)}\right)\left(E_{n}-E_{m^{\prime}}^{(0)}\right)}+...\end{split} (112)

where

Hij=ψi(0)|H|ψj(0).\begin{split}H^{\prime}_{\mathrm{ij}}=\left\langle\psi_{i}^{(0)}\left|H^{\prime}\right|\psi_{j}^{(0)}\right\rangle.\end{split} (113)

The superscript (0)(0) denotes the zeroth correction and ψi\psi_{i} denotes the ii-th eigenstate. One need to sum over the repeated indices. Eq. (112) is an equation of energy EnE_{n} alone. One can use the iteration to get the perturbative formula of the energy, which is just Rayleigh-Schrödinger perturbation formula. By solving the equation, one can get the exact energy spectrum in principle. However, the equation has infinite number of terms and one cannot solve it in most cases. Fortunately, in the case at hand, we can solve it exactly.

There is a physical interpretation for Brillouin-Wigner perturbation. Eq. (112) could be understood as the summation of Feynman diagrams: HijH^{\prime}_{ij} denotes the vertex and 1EnEm(0)\frac{1}{E_{n}-E_{m}^{(0)}} denotes the propagator. For example,

Hnm1EnEm(0)Hmm1EnEm(0)Hmn=[Uncaptioned image]\begin{split}&H^{\prime}_{nm}\frac{1}{E_{n}-E_{m}^{(0)}}H^{\prime}_{mm^{\prime}}\frac{1}{E_{n}-E_{m^{\prime}}^{(0)}}H^{\prime}_{m^{\prime}n}=\vbox{\includegraphics[scale={0.25}]{1.png}}\\ \end{split} (114)
En= [Uncaptioned image]
E_{n}=\vbox{\includegraphics[scale={0.4}]{2.png}}\\
(115)

Appendix D Towards the second order deformation

Expanding the exact deformed Lagrangian to the second-order of λ\lambda gives

=0+λ1+λ22\begin{split}\mathcal{L}=\mathcal{L}_{0}+\lambda\mathcal{L}_{1}+\lambda^{2}\mathcal{L}_{2}\end{split} (116)

with (48) and

2=2xϕ2xϕ2+ixϕxϕ(tϕxϕ2ϕtϕxϕ2ϕ)12tϕtϕxϕxϕϕϕ+i2xϕ2xϕ2(tϕϕtϕϕ)+14tϕtϕ(xϕ2ϕ2+xϕ2ϕ2).\begin{split}&\mathcal{L}_{2}=-2\partial_{x}\phi^{\ast 2}\partial_{x}\phi^{2}+i\partial_{x}\phi\partial_{x}\phi^{\ast}\left(\partial_{t}\phi^{\ast}\partial_{x}\phi^{2}\phi^{\ast}-\partial_{t}\phi\partial_{x}\phi^{\ast 2}\phi\right)\\ &\qquad-\frac{1}{2}\partial_{t}\phi\partial_{t}\phi^{\ast}\partial_{x}\phi\partial_{x}\phi^{\ast}\phi\phi^{\ast}+\frac{i}{2}\partial_{x}\phi^{2}\partial_{x}\phi^{\ast 2}\left(\partial_{t}\phi\phi^{\ast}-\partial_{t}\phi^{\ast}\phi\right)\\ &\qquad+\frac{1}{4}\partial_{t}\phi\partial_{t}\phi^{\ast}(\partial_{x}\phi^{\ast 2}\phi^{2}+\partial_{x}\phi^{2}\phi^{\ast 2}).\end{split} (117)

This Lagrangian density is not singular so we can solve tϕ\partial_{t}\phi and tϕ\partial_{t}\phi^{\ast}

tϕ=4Π+2i(ϕ+λϕxϕ2+λ2ϕxϕ2xϕ22xϕ3ϕxϕ)λ2(xϕϕxϕϕ),tϕ=4Π2i(ϕ+λϕxϕ2+λ2ϕxϕ2xϕ22xϕ3ϕxϕ)λ2(xϕϕxϕϕ).\begin{split}&\partial_{t}\phi=\frac{4\Pi^{\ast}+2i(\phi+\lambda\phi^{\ast}\partial_{x}\phi^{2}+\lambda^{2}\phi\partial_{x}\phi^{2}\partial_{x}\phi^{\ast 2}-2\partial_{x}\phi^{3}\phi^{\ast}\partial_{x}\phi^{\ast})}{\lambda^{2}(\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi)},\\ &\partial_{t}\phi^{\ast}=\frac{4\Pi-2i(\phi^{\ast}+\lambda\phi\partial_{x}\phi^{\ast 2}+\lambda^{2}\phi^{\ast}\partial_{x}\phi^{2}\partial_{x}\phi^{\ast 2}-2\partial_{x}\phi^{\ast 3}\phi\partial_{x}\phi)}{\lambda^{2}(\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi)}.\end{split} (118)

Therefore the Hamiltonian density is given by the Legendre transformation

=tϕΠ+tϕΠ=(2Π+iϕ)(2Πiϕ)λ2(xϕϕxϕϕ)2+ixϕ2ϕ(2Πiϕ)ixϕ2ϕ(2Π+iϕ)λ(xϕϕxϕϕ)2+2iΠ(ϕxϕxϕ2xϕ2ϕ)2iΠ(ϕxϕxϕ2ϕxϕ2)xϕ2ϕ2+ϕϕxϕxϕxϕ2ϕ2(xϕϕxϕϕ)22λϕϕxϕ3xϕ3(xϕϕxϕϕ)2+λ2ϕϕxϕ4xϕ4(xϕϕxϕϕ)2+𝒪(λ3).\begin{split}\mathcal{H}=&\partial_{t}\phi\Pi+\partial_{t}\phi^{\ast}\Pi^{\ast}-\mathcal{L}\\ =&\frac{(2\Pi^{\ast}+i\phi)(2\Pi-i\phi^{\ast})}{\lambda^{2}(\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi)^{2}}+\frac{i\partial_{x}\phi^{2}\phi^{\ast}(2\Pi-i\phi^{\ast})-i\partial_{x}\phi^{\ast 2}\phi(2\Pi^{\ast}+i\phi)}{\lambda(\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi)^{2}}\\ &+\frac{2i\Pi(\phi\partial_{x}\phi\partial_{x}\phi^{\ast}-2\partial_{x}\phi^{2}\phi^{\ast})-2i\Pi^{\ast}(\phi^{\ast}\partial_{x}\phi\partial_{x}\phi^{\ast}-2\phi\partial_{x}\phi^{\ast 2})-\partial_{x}\phi^{2}\phi^{\ast 2}+\phi\phi^{\ast}\partial_{x}\phi\partial_{x}\phi^{\ast}-\partial_{x}\phi^{\ast 2}\phi^{2}}{(\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi)^{2}}\\ &-\frac{2\lambda\phi\phi^{\ast}\partial_{x}\phi^{3}\partial_{x}\phi^{\ast 3}}{(\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi)^{2}}+\frac{\lambda^{2}\phi\phi^{\ast}\partial_{x}\phi^{4}\partial_{x}\phi^{\ast 4}}{(\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi)^{2}}+\mathcal{O}(\lambda^{3}).\end{split} (119)

Because the Hamiltonian density is singular at λ=0\lambda=0 so the perturbation theory is not applicable. The appearance of a non-trivial denominator is also problematic. To proceed one may introduce proper constraints as before to reduce the Hamiltonian system even though it is ad hoc. Based on our analysis of the first-order deformation, it seems that the most obvious constraints are

χ1=Πtϕ=0,χ2=Πtϕ=0.\begin{split}&\chi_{1}=\Pi-\frac{\partial\mathcal{L}}{\partial\partial_{t}\phi}=0,\quad\chi_{2}=\Pi^{\ast}-\frac{\partial\mathcal{L}}{\partial\partial_{t}\phi^{\ast}}=0.\end{split} (120)

Imposing these two constraints and using the equation of motion to get rid of tϕ\partial_{t}\phi and tϕ\partial_{t}\phi^{\ast} we end up with a Hamiltonian system:

2=xϕxϕλ(xϕxϕ)2+λ24(8[xϕxϕ]3+x2ϕx2ϕ[xϕϕxϕϕ]2),χ1=Πiϕ2iλϕxϕ22+iλ24{(xϕϕxϕϕ)2x2ϕ2(xϕxϕ)2ϕ+4xϕϕ(xϕ)3}χ1(0)+λχ1(1)+λ2χ1(2),χ2=Π+iϕ2+iλϕxϕ22iλ24{(xϕϕxϕϕ)2x2ϕ2(xϕxϕ)2ϕ+4xϕϕ(xϕ)3}χ2(0)+λχ2(1)+λ2χ2(2).\begin{split}\mathcal{H}_{2}&=\partial_{x}\phi\partial_{x}\phi^{\ast}-\lambda(\partial_{x}\phi\partial_{x}\phi^{\ast})^{2}+\frac{\lambda^{2}}{4}\left(8[\partial_{x}\phi\partial_{x}\phi^{\ast}]^{3}+\partial^{2}_{x}\phi\partial^{2}_{x}\phi^{\ast}[\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi]^{2}\right),\\ \chi_{1}&=\Pi-\frac{i\phi^{\ast}}{2}-\frac{i\lambda\phi\partial_{x}\phi^{\ast 2}}{2}+\frac{i\lambda^{2}}{4}\Big{\{}(\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi)^{2}\partial^{2}_{x}\phi^{\ast}-2(\partial_{x}\phi\partial_{x}\phi^{\ast})^{2}\phi^{\ast}+4\partial_{x}\phi\phi(\partial_{x}\phi^{\ast})^{3}\Big{\}}\\ &\equiv\chi_{1}^{(0)}+\lambda\chi_{1}^{(1)}+\lambda^{2}\chi_{1}^{(2)},\\ \chi_{2}&=\Pi^{\ast}+\frac{i\phi}{2}+\frac{i\lambda\phi^{\ast}\partial_{x}\phi^{2}}{2}-\frac{i\lambda^{2}}{4}\Big{\{}(\partial_{x}\phi\phi^{\ast}-\partial_{x}\phi^{\ast}\phi)^{2}\partial^{2}_{x}\phi-2(\partial_{x}\phi\partial_{x}\phi^{\ast})^{2}\phi+4\partial_{x}\phi^{\ast}\phi^{\ast}(\partial_{x}\phi)^{3}\Big{\}}\\ &\equiv\chi_{2}^{(0)}+\lambda\chi_{2}^{(1)}+\lambda^{2}\chi_{2}^{(2)}.\\ \end{split} (121)

Using the Dirac-Bergmann Algorithm one can find

[ϕ(x),ϕ(y)]DB=δ(xy)λC1(x,y)λ2(C2(x,y)𝑑xC1(x,x)C1(x,y)),Ci(x,y)=[Π(y),χ2(i)(x)]+[χ1(i)(y),Π(x)],i=1,2.\begin{split}&[\phi(x),\phi^{\ast}(y)]_{DB}=\delta(x-y)\\ &-\lambda C_{1}(x,y)-\lambda^{2}\left(C_{2}(x,y)-\int dx^{\prime}C_{1}(x,x^{\prime})C_{1}(x^{\prime},y)\right),\\ &C_{i}(x,y)=[\Pi(y),\chi_{2}^{(i)}(x)]+[\chi_{1}^{(i)}(y),\Pi^{\ast}(x)],\quad i=1,2.\end{split} (122)

Taking the ansatz of the canonical variables to be

φ=ϕiλχ2(1)iλ2χ2(2)+λ2ϕ2,φ=ϕ+iλχ1(1)+iλ2χ2(2)+λ2ϕ2,\begin{split}&\varphi=\phi-i\lambda\chi_{2}^{(1)}-i\lambda^{2}\chi_{2}^{(2)}+\lambda^{2}\phi_{2},\\ &\varphi^{\ast}=\phi^{\ast}+i\lambda\chi_{1}^{(1)}+i\lambda^{2}\chi_{2}^{(2)}+\lambda^{2}\phi_{2}^{\ast},\end{split} (123)

and imposing the condition

[φ(x),φ(y)]DB=δ(xy)\begin{split}[\varphi(x),\varphi^{\ast}(y)]_{DB}=\delta(x-y)\end{split} (124)

we end up with the equation for ϕ2\phi_{2} and ϕ2\phi_{2}^{\ast}

[ϕ(x),ϕ2(y)]+[ϕ2(x),ϕ(y)]14δ(xy)(xϕ(x))2(xϕ(x))2+xy(δ(xy)xϕϕxϕϕ)=0.\begin{split}&[\phi(x),\phi_{2}^{\ast}(y)]+[\phi_{2}(x),\phi^{\ast}(y)]-\frac{1}{4}\delta(x-y)(\partial_{x}\phi(x))^{2}(\partial_{x}\phi^{\ast}(x))^{2}\\ &+\partial_{x}\partial_{y}\left(\delta(x-y)\partial_{x}\phi^{\ast}\phi^{\ast}\partial_{x}\phi\phi\right)=0.\\ \end{split} (125)

Unfortunately we do not find local solutions to this equation.

References

  • (1) Y. Jiang, “Lectures on solvable irrelevant deformations of 2d quantum field theory,” [arXiv:1904.13376 [hep-th]].
  • (2) https://www.apctp.org/plan.php/ttbar2020/2951
  • (3) F. A. Smirnov and A. B. Zamolodchikov, “On space of integrable quantum field theories,” Nucl. Phys. B 915, 363-383 (2017) doi:10.1016/j.nuclphysb.2016.12.014 [arXiv:1608.05499 [hep-th]].
  • (4) A. Giveon, N. Itzhaki and D. Kutasov, “TT¯\mathrm{T}\overline{\mathrm{T}} and LST,” JHEP 07, 122 (2017) doi:10.1007/JHEP07(2017)122 [arXiv:1701.05576 [hep-th]].
  • (5) A. Cavaglià, S. Negro, I. M. Szécsényi and R. Tateo, “TT¯T\bar{T}-deformed 2D Quantum Field Theories,” JHEP 10, 112 (2016) doi:10.1007/JHEP10(2016)112 [arXiv:1608.05534 [hep-th]].
  • (6) S. Datta and Y. Jiang, “TT¯T\bar{T} deformed partition functions,” JHEP 08, 106 (2018) doi:10.1007/JHEP08(2018)106 [arXiv:1806.07426 [hep-th]].
  • (7) O. Aharony, S. Datta, A. Giveon, Y. Jiang and D. Kutasov, “Modular invariance and uniqueness of TT¯T\bar{T} deformed CFT,” JHEP 01, 086 (2019) doi:10.1007/JHEP01(2019)086 [arXiv:1808.02492 [hep-th]].
  • (8) L. McGough, M. Mezei and H. Verlinde, “Moving the CFT into the bulk with TT¯T\overline{T},” JHEP 04 (2018) 010, arXiv:1611.03470 [hep-th].
  • (9) S. Chakraborty, A. Giveon, N. Itzhaki and D. Kutasov, “Entanglement Beyond AdS,” Nucl. Phys. B 935, 290-309 (2018),arXiv:1805.06286 [hep-th].
  • (10) S. Dubovsky, V. Gorbenko and M. Mirbabayi, “Asymptotic fragility, near AdS2AdS_{2} holography and TT¯T\bar{T},” JHEP 09 (2017) 136, arXiv:1706.06604 [hep-th].
  • (11) S. Dubovsky, V. Gorbenko and G. Hernndez-Chifflet, “TT¯T\bar{T} Partition Function from Topological Gravity,”JHEP 09, 158 (2018), arXiv:1805.07386 [hep-th].
  • (12) V. Shyam, “Background independent holographic dual to TT¯T\bar{T} deformed CFT with large central charge in 2 dimensions,” JHEP 10 (2017) 108, arXiv:1707.08118 [hep-th].
  • (13) G. Giribet, “TT¯T\bar{T}-deformations, AdS/CFT and correlation functions,” JHEP 02 (2018) 114, arXiv:1711.02716 [hep-th].
  • (14) P. Kraus, J. Liu and D. Marolf, “Cutoff AdS3AdS_{3} versus the TT¯T\bar{T} deformation,” JHEP 07, 027 (2018), arXiv:1801.02714 [hep-th].
  • (15) W. Cottrell and A. Hashimoto, “Comments on TT¯T\bar{T} double trace deformations and boundary conditions,”Phys. Lett. B 789, 251-255 (2019), arXiv:1801.09708 [hep-th].
  • (16) B. Chen, L. Chen and P. X. Hao, “Entanglement entropy in TT¯T\overline{T}-deformed CFT,” Phys. Rev. D 98, no.8, 086025 (2018) doi:10.1103/PhysRevD.98.086025 [arXiv:1807.08293 [hep-th]].
  • (17) T. Hartman, J. Kruthoff, E. Shaghoulian and A. Tajdini, “Holography at finite cutoff with a T2T^{2} deformation,” JHEP 03, 004 (2019) doi:10.1007/JHEP03(2019)004 [arXiv:1807.11401 [hep-th]].
  • (18) P. Caputa, S. Datta and V. Shyam, “Sphere partition functions \& cut-off AdS,” JHEP 05, 112 (2019) doi:10.1007/JHEP05(2019)112 [arXiv:1902.10893 [hep-th]].
  • (19) C. Murdia, Y. Nomura, P. Rath and N. Salzetta, “Comments on holographic entanglement entropy in TTTT deformed conformal field theories,” Phys. Rev. D 100, no.2, 026011 (2019) doi:10.1103/PhysRevD.100.026011 [arXiv:1904.04408 [hep-th]].
  • (20) H. S. Jeong, K. Y. Kim and M. Nishida, “Entanglement and Rényi entropy of multiple intervals in TT¯T\overline{T}-deformed CFT and holography,” Phys. Rev. D 100, no.10, 106015 (2019) doi:10.1103/PhysRevD.100.106015 [arXiv:1906.03894 [hep-th]].
  • (21) M. Guica and R. Monten, “TT¯T\bar{T} and the mirage of a bulk cutoff,” [arXiv:1906.11251 [hep-th]].
  • (22) B. Chen, L. Chen and C. Y. Zhang, “Surface/state correspondence and TT¯T\overline{T} deformation,” Phys. Rev. D 101, no.10, 106011 (2020) doi:10.1103/PhysRevD.101.106011 [arXiv:1907.12110 [hep-th]].
  • (23) S. Grieninger, “Entanglement entropy and TT¯T\overline{T} deformations beyond antipodal points from holography,” JHEP 11, 171 (2019) doi:10.1007/JHEP11(2019)171 [arXiv:1908.10372 [hep-th]].
  • (24) A. Lewkowycz, J. Liu, E. Silverstein and G. Torroba, “TT¯T\overline{T} and EE, with implications for (A)dS subregion encodings,” JHEP 04, 152 (2020) doi:10.1007/JHEP04(2020)152 [arXiv:1909.13808 [hep-th]].
  • (25) L. Apolo, S. Detournay and W. Song, “TsT, TT¯T\bar{T} and black strings,” JHEP 06, 109 (2020) doi:10.1007/JHEP06(2020)109 [arXiv:1911.12359 [hep-th]].
  • (26) S. Hirano and M. Shigemori, “Random boundary geometry and gravity dual of TT¯T\overline{T} deformation,” JHEP 11, 108 (2020) doi:10.1007/JHEP11(2020)108 [arXiv:2003.06300 [hep-th]].
  • (27) J. Kruthoff and O. Parrikar, “On the flow of states under TT¯T\overline{T},” [arXiv:2006.03054 [hep-th]].
  • (28) G. Jafari, A. Naseh and H. Zolfi, “Path Integral Optimization for TT¯T\bar{T} Deformation,” Phys. Rev. D 101, no.2, 026007 (2020) doi:10.1103/PhysRevD.101.026007 [arXiv:1909.02357 [hep-th]].
  • (29) M. He and Y. h. Gao, “TT¯/JT¯T\bar{T}/J\bar{T}-deformed WZW models from Chern-Simons AdS3 gravity with mixed boundary conditions,” Phys.Rev.D 103, no.12, 126019 (2021) [arXiv:2012.05726 [hep-th]].
  • (30) B. Pozsgay, Y. Jiang and G. Takács, “TT¯T\bar{T}-deformation and long range spin chains,” JHEP 03, 092 (2020) doi:10.1007/JHEP03(2020)092 [arXiv:1911.11118 [hep-th]].
  • (31) E. Marchetto, A. Sfondrini and Z. Yang, “TT¯T\bar{T} Deformations and Integrable Spin Chains,” Phys. Rev. Lett. 124, no.10, 100601 (2020) doi:10.1103/PhysRevLett.124.100601 [arXiv:1911.12315 [hep-th]].
  • (32) J. Cardy, “TT¯T\overline{T} deformations of non-Lorentz invariant field theories,” [arXiv:1809.07849 [hep-th]].
  • (33) J. Cardy and B. Doyon, “TT¯T{\overline{T}} deformations and the width of fundamental particles,” [arXiv:2010.15733 [hep-th]].
  • (34) Y. Jiang, “TT¯\mathrm{T}\overline{\mathrm{T}}-deformed 1d Bose gas,” [arXiv:2011.00637 [hep-th]].
  • (35) T. Bargheer, N. Beisert and F. Loebbert, “Long-Range Deformations for Integrable Spin Chains,” J. Phys. A 42, 285205 (2009) doi:10.1088/1751-8113/42/28/285205 [arXiv:0902.0956 [hep-th]].
  • (36) S. Dubovsky, V. Gorbenko and M. Mirbabayi, “Asymptotic fragility, near AdS2 holography and TT¯T\overline{T},” JHEP 09, 136 (2017) doi:10.1007/JHEP09(2017)136 [arXiv:1706.06604 [hep-th]].
  • (37) G. Mussardo and P. Simon, “Bosonic type S matrix, vacuum instability and CDD ambiguities,” Nucl. Phys. B 578, 527-551 (2000) doi:10.1016/S0550-3213(99)00806-8 [arXiv:hep-th/9903072 [hep-th]].
  • (38) P. Ceschin, R. Conti and R. Tateo, “TT¯\text{T}\bar{\text{T}}-deformed Nonlinear Schr\”odinger,” [arXiv:2012.12760 [hep-th]].
  • (39) D. Hansen, Y. Jiang and J. Xu, “Geometrizing non-relativistic bilinear deformations,” [arXiv:2012.12290 [hep-th]].
  • (40) G. Bonelli, N. Doroud and M. Zhu, “TT¯T\bar{T}-deformations in closed form,” JHEP 06, 149 (2018) doi:10.1007/JHEP06(2018)149 [arXiv:1804.10967 [hep-th]].
  • (41) S. Frolov, “TT¯T\overline{T} Deformation and the Light-Cone Gauge,” Proc. Steklov Inst. Math. 309, 107-126 (2020) doi:10.1134/S0081543820030098 [arXiv:1905.07946 [hep-th]].
  • (42) G. Jorjadze and S. Theisen, “Canonical maps and integrability in TT¯T\bar{T} deformed 2d CFTs,” [arXiv:2001.03563 [hep-th]].
  • (43) L. A. Gergely, “On Hamiltonian formulations of the Schrodinger system,” Annals Phys. 298, 394-402 (2002) doi:10.1006/aphy.2002.6262 [arXiv:hep-th/0301052 [hep-th]].
  • (44) A. B. Zamolodchikov, “From tricritical Ising to critical Ising by thermodynamic Bethe ansatz,” Nucl. Phys. B 358, 524-546 (1991) doi:10.1016/0550-3213(91)90423-U
  • (45) M. Caselle, D. Fioravanti, F. Gliozzi and R. Tateo, “Quantisation of the effective string with TBA,” JHEP 07, 071 (2013) doi:10.1007/JHEP07(2013)071 [arXiv:1305.1278 [hep-th]].
  • (46) R. Shankar, “Principles of Quantum Mechanics,” Springe 1994.
  • (47) Kai Sun, Time-Independent Perturbation Theory, 2016. http://www-personal.umich.edu/~sunkai/teaching/Winter_2016/chapter02.pdf