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

Field-induced dimer orders in quantum spin chains

Shunsuke C. Furuya Condensed Matter Theory Laboratory, RIKEN, Wako, Saitama 351-0198, Japan
Abstract

Field-induced excitation gaps in quantum spin chains are an interesting phenomenon related to confinements of topological excitations. In this paper, I present a novel type of this phenomenon. I show that an effective magnetic field with a fourfold screw symmetry induces the excitation gap accompanied by dimer orders. The dimer order parameter and the excitation gap exhibit characteristic power-law dependence on the fourfold screw-symmetric field. Moreover, the field-induced dimer order and the field-induced Néel order coexist when the external uniform magnetic field, the fourfold screw-symmetric field, and the twofold staggered field are applied. This situation is in close connection with a compound [Cu(pym)(H2O)4]SiF6H2O\mathrm{[Cu(pym)(H_{2}O)_{4}]SiF_{6}\cdot H_{2}O} [J. Liu et al., Phys. Rev. Lett. 122, 057207 (2019)]. In this paper, I discuss a mechanism of field-induced dimer orders by using a density-matrix renormalization group method, a perturbation theory, and quantum field theories.

I Introduction

Quantum spin-1/21/2 chains do not have a unique gapped ground state in the presence of the time-reversal symmetry unless either the U(1) spin-rotation symmetry or the translation symmetry is broken Lieb et al. (1961); Affleck and Haldane (1987); Furuya and Oshikawa (2017). For example, the spin-1/21/2 Heisenberg antiferromagnetic (HAFM) chain has a unique gapless ground state called the Tomonaga-Luttinger (TL) liquid state Giamarchi (2004). Even when the time-reversal symmetry is broken by the external magnetic field, the TL liquid does not immediately acquire the gap, though it eventually does with the saturated magnetization Metlitski and Thorngren (2018). This is because the external magnetic field is uniform in the scale of spin chains. It breaks neither the U(1) rotation nor the translation symmetry. Interestingly, however, there are several spin-1/21/2 chain compounds where the magnetic field immediately opens the excitation gap Dender et al. (1997); Zvyagin et al. (2004); Umegaki et al. (2009).

This puzzle of the field-induced gap was found in Cu Benzoate Dender et al. (1997) and later solved with quantum field theories Oshikawa and Affleck (1997); Affleck and Oshikawa (1999); Furuya and Oshikawa (2012). Essentially, the field-induced excitation gap in those compounds comes from an absence of a bond-centered inversion symmetry. This low crystalline symmetry allows the gg tensor of electron spins to have a twofold staggered component. The magnetic field, when combined with the low symmetry, generates a twofold staggered magnetic field that breaks the translation symmetry. As a result, the uniform magnetic field induces the excitation gap and also the Néel order in the direction of the effectively generated twofold staggered magnetic field. Thanks to the dimensionality and strong interactions among elementary excitations, the excitation gap and the Néel order exhibit interesting power-law behaviors that deviate from spin-wave predictions Oshikawa and Affleck (1997); Affleck and Oshikawa (1999); Furuya et al. (2011). The phenomenon of the field-induced excitation gap has drawn attention for its connection with confinement of topological excitations Faure et al. (2018); Takayoshi et al. (2018).

In this paper, I discuss a novel field-induced excitation gap phenomenon. That is field-induced dimer orders in quantum spin chains.

Refer to caption
Figure 1: (a) A fourfold screw spin chain. Each ball represents a spin operator. (b) A fourfold screw field symmetric under the bond-centered inversion IbI_{b}. Arrows depict directions of the fourfold screw field.

A key ingredient is a fourfold screw symmetry (Fig. 1). The screw structure can violate two kinds of inversion symmetries at the same time, namely, the bond-centered inversion symmetry and a site-centered inversion symmetry. Such a fourfold structure is indeed incorporated in the gg tensor of spin-chain compounds, BaCo2V2O8\mathrm{BaCo_{2}V_{2}O_{8}} Kimura et al. (2013); Faure et al. (2018) and [Cu(pym)(H2O)4]SiF6H2O\mathrm{[Cu(pym)(H_{2}O)_{4}]SiF_{6}\cdot H_{2}O} Liu et al. (2019). When the uniform magnetic field is externally applied, the fourfold screw symmetry manifests itself as an effective fourfold screw-symmetric magnetic field. The fourfold screw field brings dimer orders to spin chains immediately. I discuss first a mechanism of the dimer-order generation. Next I take the twofold staggered field into account and discuss coexistent growth of the dimer and Néel orders with increase of the uniform magnetic field.

This paper is organized as follows. I define a spin-chain model and show numerical evidence of the field-induced dimer orders in the simplest case in Sec. II. A qualitative mechanism of field-induced dimer orders is discussed in Sec. III, where the spin-chain model is replaced to a spinless fermion model which is smoothly deformed from the original spin-chain model. Here, the low-energy effective Hamiltonian is systematically derived. In Sec. IV, on the basis of observations made in Sec. III, I develop a quantum field theory that explains quantitatively numerical results of Sec. II. The quantum field theory also predicts the coexistence of the dimer and Néel orders both of which grow with the uniform magnetic field. This coexistent growth of the dimer and Néel orders are discussed in Sec. V, which is supported by numerical calculations. I also discuss relevance of theoretical results to experiments in Sec. VI. Finally, I summarize the paper in Sec. VII.

Refer to caption
Figure 2: The lowest-energy excitation gap from the ground state of the Hamiltonian (3) is plotted against the fourfold screw field h4h_{4} for system sizes LL ranging from L=240L=240 to 400400. The gap is extrapolated to the L+L\to+\infty limit by using a formula Δ=a0+a1L+a2L2\Delta=a_{0}+\frac{a_{1}}{L}+\frac{a_{2}}{L^{2}}, where ana_{n} (n=0,1,2n=0,1,2) are fitting parameters. The error of the extrapolated data is estimated to be 11%11~{}\% for h4/J=0.05h_{4}/J=0.05 and <0.28%<0.28~{}\% for h4/J0.1h_{4}/J\geq 0.1. The solid curve is the best fit of the extrapolated data by a function Δ=b0h4b1+b2\Delta=b_{0}{h_{4}}^{b_{1}}+b_{2} with fitting parameters bnb_{n} (n=0,1,2n=0,1,2). Though the fitted result has an unphysical offset b20b_{2}\not=0, its h4h_{4} dependence implies the gap ΔJ(h4/J)1.34\Delta\propto J(h_{4}/J)^{1.34}. This estimation of the agrees with the field-theoretical prediction (32).

II Screw field

II.1 Definition of the model

In this paper I discuss a quantum spin-1/21/2 chain with the following Hamiltonian:

\displaystyle\mathcal{H} =Jj𝑺j𝑺j+1h0jSjzh2j(1)jSjx\displaystyle=J\sum_{j}\bm{S}_{j}\cdot\bm{S}_{j+1}-h_{0}\sum_{j}S_{j}^{z}-h_{2}\sum_{j}(-1)^{j}S_{j}^{x}
h4jδjSjz,\displaystyle\qquad-h_{4}\sum_{j}\delta_{j}S_{j}^{z}, (1)

where 𝑺j\bm{S}_{j} is the S=1/2S=1/2 spin operator, J>0J>0 is the antiferromagnetic exchange coupling, and δj=1,1,1,1\delta_{j}=1,1,-1,-1 for respectively, j=0,1,2,3mod4j=0,1,2,3\mod 4. Parameters h0h_{0}, h2h_{2}, and h4h_{4} denote the uniform magnetic field, the twofold staggered field, and the fourfold screw field, respectively. Note that δj\delta_{j} has a simple expression,

δj\displaystyle\delta_{j} =2cos(π2j14).\displaystyle=\sqrt{2}\cos\biggl{(}\pi\frac{2j-1}{4}\biggr{)}. (2)

Throughout this paper, I employ the unit of =a=1\hbar=a=1 unless otherwise stated, where aa is the lattice spacing.

The spin-chain model (1) is related to a model proposed for [Cu(pym)(H2O)4]SiF6H2O\mathrm{[Cu(pym)(H_{2}O)_{4}]SiF_{6}\cdot H_{2}O} Liu et al. (2019) but differs from the latter in three points: field directions, a weak exchange anisotropy, and a uniform Dzyaloshinskii-Moriya (DM) interaction. In Ref. Liu et al. (2019), h0h_{0} and h2h_{2} are applied in the xx and the yy directions, respectively. The model used in Ref. Liu et al. (2019) contains a weak XXZ interaction and the uniform DM interaction. Those differences hardly affect the ground state of [Cu(pym)(H2O)4]SiF6H2O\mathrm{[Cu(pym)(H_{2}O)_{4}]SiF_{6}\cdot H_{2}O}, which are to be clarified in Sec. VI.

In the compound [Cu(pym)(H2O)4]SiF6H2O\mathrm{[Cu(pym)(H_{2}O)_{4}]SiF_{6}\cdot H_{2}O} Liu et al. (2019), the twofold staggered field h2h_{2} and the fourfold screw field h4h_{4} originate from the gg tensor of electrons and are thus proportional to the externally applied uniform magnetic field h0h_{0}. In this section, I first deal with an unrealistic but simplest situation with h0=h2=0h_{0}=h_{2}=0 and h40h_{4}\not=0 in Sec. II.2. I will discuss a more realistic situation with h2h0h_{2}\propto h_{0} and h4h0h_{4}\propto h_{0} later in Sec. V.

II.2 Fourfold screw field

Refer to caption
Figure 3: The transverse dimer order parameter (5) of the ground state of the Hamiltonian (3) is plotted against the fourfold screw field h4h_{4} for system sizes from L=240L=240 to 400400. The dimer is extrapolated to the L+L\to+\infty limit by using a formula D=a0+a1L+a2LD_{\perp}=a^{\prime}_{0}+\frac{a^{\prime}_{1}}{\sqrt{L}}+\frac{a^{\prime}_{2}}{L} because the scaling dimension of the dimer order parameter is 1/21/2. The error of the extrapolated data is estimated to be 6.5%6.5~{}\% for h4/J=0.025h_{4}/J=0.025 and <0.21%<0.21~{}\% for h4/J0.1h_{4}/J\geq 0.1. The error monotonically decreases with incraese of h4/Jh_{4}/J. The solid curve is the best fit of the extrapolated data by a function D=b0h4b1+b2D_{\perp}=b^{\prime}_{0}{h_{4}}^{b^{\prime}_{1}}+b^{\prime}_{2} with fitting parameters bnb^{\prime}_{n} (n=0,1,2n=0,1,2). Though DD_{\perp} shows an unphysical offset b20b^{\prime}_{2}\not=0, it implies a power law D(h4/J)0.672D_{\perp}\propto(h_{4}/J)^{0.672}. It agrees with the field-theoretical prediction (34).
Refer to caption
Figure 4: The longitudinal dimer order parameter (6) of the ground state of the Hamiltonian (3) is plotted against the fourfold screw field h4h_{4} for system sizes from L=240L=240 to 400400. The dimer is extrapolated to the L+L\to+\infty limit in the same way as DD_{\perp}. The error of the extrapolated data is estimated to be 6.4%6.4~{}\% for h4/J=0.025h_{4}/J=0.025 and <0.20%<0.20~{}\% for h4/J0.1h_{4}/J\geq 0.1. The solid curve is the best fit of the extrapolated data. It implies a power law D(h4/J)1.07D_{\parallel}\propto(h_{4}/J)^{1.07}, which differs from that for DD_{\perp} [Eq. (7)].

The fourfold screw field h4h_{4} can generate an excitation gap all by itself to the ground state of the spin chain. To show this, I set h0=h2=0h_{0}=h_{2}=0 and discuss the h4h_{4} dependence of the lowest-energy excitation gap from the ground state. The Hamiltonian is thus simplified as

4\displaystyle\mathcal{H}_{4} =Jj𝑺j𝑺j+1h4jδjSjz.\displaystyle=J\sum_{j}\bm{S}_{j}\cdot\bm{S}_{j+1}-h_{4}\sum_{j}\delta_{j}S_{j}^{z}. (3)

When h4=0h_{4}=0, the ground state of the model (3) is gapless Giamarchi (2004). Figure 2 shows numerical results on the excitation gap obtained by using the density-matrix renormalization group (DMRG) method with the ITensor C++ library ITe , where I used the bond dimension χ=400\chi=400 and the truncation error cutoff 1×10101\times 10^{-10}. Note that all the DMRG calculations in this paper were performed with the open boundary condition. The DMRG result implies that an infinitesimal h4/Jh_{4}/J immediately opens the excitation gap between the ground state and the lowest-energy excited state,

ΔJ(h4J)1.34.\displaystyle\Delta\propto J\biggl{(}\frac{h_{4}}{J}\biggr{)}^{1.34}. (4)

Similarly to the twofold staggered field, the fourfold screw field induces an excitation gap with a power law. However, the power 1.341.34 differs from that, 2/32/3, of the twofold staggered field Oshikawa and Affleck (1997); Affleck and Oshikawa (1999).

Unlike the twofold staggered field, the fourfold screw field h4h_{4} induces no Néel order. Instead, h4h_{4} induces dimer orders (Figs. 3 and 4),

D\displaystyle D_{\perp} =1Lj(1)j(SjxSj+1x+SjySj+1y),\displaystyle=\frac{1}{L}\sum_{j}\braket{(-1)^{j}(S_{j}^{x}S_{j+1}^{x}+S_{j}^{y}S_{j+1}^{y})}, (5)
D\displaystyle D_{\parallel} =1Lj(1)jSjzSj+1z,\displaystyle=\frac{1}{L}\sum_{j}\braket{(-1)^{j}S_{j}^{z}S_{j+1}^{z}}, (6)

where LL is the length of the spin chain. The dimer order parameters (5) and (6) show different power-law dependence on h4h_{4}. DMRG results (Figs. 3 and 4) imply

D\displaystyle D_{\perp} (h4J)0.672,\displaystyle\propto\biggl{(}\frac{h_{4}}{J}\biggr{)}^{0.672}, (7)
D\displaystyle D_{\parallel} (h4J)1.07.\displaystyle\propto\biggl{(}\frac{h_{4}}{J}\biggr{)}^{1.07}. (8)

Induction of DD_{\parallel} by h4h_{4} is easily understandable. Let us recall that h4h_{4} is coupled to an operator,

fjz\displaystyle f_{j}^{z} =2cos(π2j14)Sjz.\displaystyle=\sqrt{2}\cos\biggl{(}\pi\frac{2j-1}{4}\biggr{)}S_{j}^{z}. (9)

The fourfold screw field induces the uniform fzf^{z} order:

jfjz0.\displaystyle\sum_{j}\braket{f_{j}^{z}}\not=0. (10)

The longitudinal dimer order parameter DD_{\parallel} is written in terms of fjzf_{j}^{z} as

D\displaystyle D_{\parallel} =1Ljfjzfj+1z.\displaystyle=\frac{1}{L}\sum_{j}\braket{f_{j}^{z}f_{j+1}^{z}}. (11)

Nonzero DD_{\parallel} follows immediately from the uniform fjzf_{j}^{z} order (10). However, the induction of the transverse dimer order (7) is nontrivial.

III Free spinless fermion theory

This section is devoted to a qualitative explanation on a mechanism of the field-induced transverse dimer order (7). For this purpose, I rewrite the spin chain (3) in terms of spinless fermions with the aid of the Jordan-Wigner transformation Giamarchi (2004). Let cjc_{j}^{\dagger} and cjc_{j} be creation and annihilation operators of the spinless fermion at the site jj, respectively. The spin-chain model (3) is equivalent to the following model of interacting spinless fermions:

4\displaystyle\mathcal{H}_{4} =J2j(cjcj+1+H.c.)h4jδj(cjcj12)\displaystyle=-\frac{J}{2}\sum_{j}(c_{j}^{\dagger}c_{j+1}+\mathrm{H.c.})-h_{4}\sum_{j}\delta_{j}\biggl{(}c_{j}^{\dagger}c_{j}-\frac{1}{2}\biggr{)}
+Jj(cjcj12)(cj+1cj+112),\displaystyle\qquad+J\sum_{j}\biggl{(}c_{j}^{\dagger}c_{j}-\frac{1}{2}\biggr{)}\biggl{(}c_{j+1}^{\dagger}c_{j+1}-\frac{1}{2}\biggr{)}, (12)

where H.c.\mathrm{H.c.} denotes the Hermitian conjugate.

The interaction of spinless fermions, the second line of Eq. (12), comes from the longitudinal component of the exchange interaction, JjSjzSj+1zJ\sum_{j}S_{j}^{z}S_{j+1}^{z}. Even if this interaction is ignored, qualitative aspects of the ground state are kept intact since the HAFM chain and the XY chain belong to the same TL-liquid phase Giamarchi (2004). Therefore, I discuss in this section the free spinless fermion model,

XY\displaystyle\mathcal{H}_{\rm XY} =Jj(cjcj+1+H.c.)h4jδj(cjcj12),\displaystyle=-J\sum_{j}(c_{j}^{\dagger}c_{j+1}+\mathrm{H.c.})-h_{4}\sum_{j}\delta_{j}\biggl{(}c_{j}^{\dagger}c_{j}-\frac{1}{2}\biggr{)}, (13)

instead of the model (12). In terms of spins, it is the XY chain in the fourfold screw field,

XY\displaystyle\mathcal{H}_{\rm XY} =Jj(SjxSj+1x+SjySj+1y)+h4jδjSjz.\displaystyle=J\sum_{j}(S_{j}^{x}S_{j+1}^{x}+S_{j}^{y}S_{j+1}^{y})+h_{4}\sum_{j}\delta_{j}S_{j}^{z}. (14)

III.1 Particle-hole excitations

Performing the Fourier transformation on Eq. (14), I obtain

XY\displaystyle\mathcal{H}_{\rm XY} =kϵ(k)ckck\displaystyle=\sum_{k}\epsilon(k)c_{k}^{\dagger}c_{k}
h42k(eπi/4ckck+π2+eπi/4ck+π2ck)\displaystyle\qquad-\frac{h_{4}}{\sqrt{2}}\sum_{k}\bigl{(}e^{-\pi i/4}c_{k}^{\dagger}c_{k+\frac{\pi}{2}}+e^{\pi i/4}c_{k+\frac{\pi}{2}}^{\dagger}c_{k}\bigr{)}
+const.\displaystyle\qquad+\mathrm{const.} (15)

where ϵ(k)=(J/2)cosk\epsilon(k)=-(J/2)\cos k and k(π,π]k\in(-\pi,\pi] is the wave number. When J4=0J_{4}=0, the spinless fermion is free and has the simple cosine dispersion. Since the total magnetization is zero in the XY chain, the cosine band is half occupied and the Fermi points are located at ±π/2\pm\pi/2.

Refer to caption
Figure 5: Particle-hole excitations are schematically drawn. Filled and empty circles depict a particle and a hole, respectively. (a) The twofold staggered field h2h_{2} and the bond alternation generate low-energy particle-hole excitations with a wave number q=πq=\pi. (b) The fourfold screw field h4h_{4} generates high-energy excitations with q=π/2q=\pi/2 when they act on the ground state once. (c) The screw field can generate low-energy excitations with q=πq=\pi when they act on the ground state twice.

Let us make some observations on effects of the fourfold screw field on the free spinless fermion chain in the TL-liquid phase. Particle-hole excitations are the fundamental low-energy excitation in the TL-liquid phase. An operator ρq=kck+qck\rho_{q}=\sum_{k}c_{k+q}^{\dagger}c_{k}, which creates a particle-hole excitation with a wave number qq, can be written as a superposition of bosonic creation and annihilation operators of the TL liquid Giamarchi (2004). When the second line of Eq. (15) acts on the ground state of the XY model, particle-hole excitations with wave numbers q=±π/2q=\pm\pi/2 are generated [Fig. 5 (b)]. Apparently, such an h4h_{4} term hardly affects the low-energy physics of the XY model because these particle-hole excitations have large excitation energies of O(J)O(J). However, applying the h4h_{4} term twice to the ground state, I can generate low-energy particle-hole excitations with the wave number q=πq=\pi [Figs. 5 (a,c)]. These observations show that though the fourfold screw field is highly irrelevant, it will generates a relevant interaction in a second-order perturbation process.

III.2 Low-energy effective Hamiltonian

To confirm the perturbative generation of the relevant interaction, I derive a simple effective Hamiltonian that governs the low-energy physics of the fermion chain (15). Among several options to derive such a low-energy effective Hamiltonian is to use the Schrieffer-Wolff canonical transformation Schrieffer and Wolff (1966); Slagle and Kim (2017). The generic theory is explained in Appendix A.1. Here, I briefly summarize the derivation. First, I perform a canonical transformation,

XY:=eηXYeη\displaystyle\mathcal{H}^{\prime}_{\rm XY}:=e^{\eta}\mathcal{H}_{\rm XY}e^{-\eta} (16)

with an antiunitary operator η\eta. Two Hamiltonians XY\mathcal{H}_{\rm XY} and XY\mathcal{H}^{\prime}_{\rm XY} have one-to-one corresponding lists of eigenstates with exactly the same eigenenergies. Next, I perform a perturbative expanison by using a projection operator PP onto a low-energy subspace {|ϕk}k\{\ket{\phi_{k}}\}_{k} with kRk\in R and

R={k(π,π]| 0||k||kF||<Λ},\displaystyle R=\bigl{\{}k\in(-\pi,\pi]|\,0\leq\bigl{|}|k|-|k_{F}|\bigr{|}<\Lambda\bigr{\}}, (17)

where Λ\Lambda is a cutoff in the wave number and assumed as Λ1\Lambda\ll 1. Here, |ϕk\ket{\phi_{k}} is an eigenstate of the unperturbed Hamiltonian, Eq. (15) for h4=0h_{4}=0, with the total wave number kk. PP can explicitly be written as P=kR|ϕkϕk|P=\sum_{k\in R}\ket{\phi_{k}}\bra{\phi_{k}}. The projection onto the low-energy subspace leads to the effective Hamiltonian

~XY\displaystyle\tilde{\mathcal{H}}_{\rm XY} =PXYP.\displaystyle=P\mathcal{H}^{\prime}_{\rm XY}P. (18)

Choosing η\eta properly, I can simplify the perturbative expansion of the right hand side of Eq. (18). The effective Hamiltonian up to the second order of h4/Jh_{4}/J is then given by (see Appendix A.2)

~XY\displaystyle\tilde{\mathcal{H}}_{\rm XY}
=kRϵ(k)ckckih424kR(ck+πckckck+π)\displaystyle=\sum_{k\in R}\epsilon(k)c_{k}^{\dagger}c_{k}-i\frac{h_{4}^{2}}{4}\sum_{k\in R}(c_{k+\pi}^{\dagger}c_{k}-c_{k}^{\dagger}c_{k+\pi})
×(1ϵ(k)ϵ(k+π2)+1ϵ(k+π)ϵ(k+π2))\displaystyle\qquad\times\biggl{(}\frac{1}{\epsilon(k)-\epsilon(k+\frac{\pi}{2})}+\frac{1}{\epsilon(k+\pi)-\epsilon(k+\frac{\pi}{2})}\biggr{)}
+h422kRckck(1ϵ(k)ϵ(kπ2)+1ϵ(k)ϵ(k+π2)).\displaystyle\quad+\frac{h_{4}^{2}}{2}\sum_{k\in R}c_{k}^{\dagger}c_{k}\biggl{(}\frac{1}{\epsilon(k)-\epsilon(k-\frac{\pi}{2})}+\frac{1}{\epsilon(k)-\epsilon(k+\frac{\pi}{2})}\biggr{)}. (19)

Since the cutoff Λ\Lambda is small enough, the kinetic term of Eq. (19) can be linearized around the Fermi surface Giamarchi (2004). Creation operators ckc_{k}^{\dagger} at kπ/2k\approx\pi/2 and kπ/2k\approx-\pi/2 are replaced to those of different species, which I denote as ck,Rc_{k,R}^{\dagger} and ck,Lc_{k,L}^{\dagger}, respectively. RR and LL refer to right movers and left movers of fermions. The low-energy Hamiltonian thus turns out to be

~XY\displaystyle\tilde{\mathcal{H}}_{\rm XY} kR{vF(kkF)ck,Rck,RvF(k+kF)ck,Lck,L}\displaystyle\approx\sum_{k\in R}\{v_{F}(k-k_{F})c_{k,R}^{\dagger}c_{k,R}-v_{F}(k+k_{F})c_{k,L}^{\dagger}c_{k,L}\}
+ih42JkR(ck,Rck,Lck,Lck,R),\displaystyle\qquad+i\frac{h_{4}^{2}}{J}\sum_{k\in R}(c_{k,R}^{\dagger}c_{k,L}-c_{k,L}^{\dagger}c_{k,R}), (20)

where vFv_{F} is the Fermi velocity. Note that the last term of Eq. (19) was discarded in the linearized Hamiltonian (20) because the coefficient of ckckc_{k}^{\dagger}c_{k} is O(Λh42/J)O(\Lambda h_{4}^{2}/J) and thus negligibly small for kRk\in R. The Hamiltonian (20) is diagonalized in terms of Majorana fermions,

ξk,ν\displaystyle\xi_{k,\nu} =ck,ν+ck,ν2,χk,ν=ck,νck,ν2i\displaystyle=\frac{c_{k,\nu}+c_{k,\nu}^{\dagger}}{\sqrt{2}},\quad\chi_{k,\nu}=\frac{c_{k,\nu}-c_{k,\nu}^{\dagger}}{\sqrt{2}i} (21)

The second line of Eq. (20) becomes mass terms,

ih42JkR(ck,Rck,Lck,Lck,R)\displaystyle i\frac{h_{4}^{2}}{J}\sum_{k\in R}(c_{k,R}^{\dagger}c_{k,L}-c_{k,L}^{\dagger}c_{k,R}) =ih42JkR(ξRξL+χRχL),\displaystyle=i\frac{h_{4}^{2}}{J}\sum_{k\in R}(\xi_{R}\xi_{L}+\chi_{R}\chi_{L}), (22)

which indicate that these Majorana fermions have the excitation gap Δ=h42/J\Delta=h_{4}^{2}/J Shelton et al. (1996). Note that the gap does not reproduce the power law (4). This discrepancy of the power is attributed to interactions of fermions. I will come back to this point in Sec. IV.

The mass terms (22) are nothing but the bond alternation Jδj(1)j(SjxSj+1x+SjySj+1y)J\delta_{\perp}\sum_{j}(-1)^{j}(S_{j}^{x}S_{j+1}^{x}+S_{j}^{y}S_{j+1}^{y}) with δ=(h4/J)2\delta_{\perp}=(h_{4}/J)^{2} Giamarchi (2004):

j\displaystyle\sum_{j} (1)j(SjxSj+1x+SjySj+1y)\displaystyle(-1)^{j}(S_{j}^{x}S_{j+1}^{x}+S_{j}^{y}S_{j+1}^{y})
=12j(1)j(cjcj+1+cj+1cj)\displaystyle=-\frac{1}{2}\sum_{j}(-1)^{j}(c_{j}^{\dagger}c_{j+1}+c_{j+1}^{\dagger}c_{j})
ikR(ck,Rck,Lck,Lck,R).\displaystyle\approx i\sum_{k\in R}(c_{k,R}^{\dagger}c_{k,L}-c_{k,L}^{\dagger}c_{k,R}). (23)

For comparison, the following is the twofold staggered field term in terms of fermions.

h2j(1)jSjzh2kR(ck,Rck,L+ck,Lck,R).\displaystyle h_{2}\sum_{j}(-1)^{j}S_{j}^{z}\approx h_{2}\sum_{k\in R}(c_{k,R}^{\dagger}c_{k,L}+c_{k,L}^{\dagger}c_{k,R}). (24)

Here, I comment on effects of the canonical transformation (16) on observables. The canonical transformation also transforms an operator, say, 𝒪\mathcal{O}, in the original model XY\mathcal{H}_{\rm XY} to

𝒪~\displaystyle\tilde{\mathcal{O}} =eη𝒪eη.\displaystyle=e^{\eta}\mathcal{O}e^{-\eta}. (25)

Note that one observes 𝒪\braket{\mathcal{O}} in experiments, not 𝒪~\braket{\tilde{\mathcal{O}}}. Equations (23) and (24) refer to the latter. According to the generic framework of Appendix A.1, the operator eηe^{\eta} can be expanded with h4/Jh_{4}/J,

𝒪~=𝒪+[η,𝒪]+[η,𝒪]+12[η,[η,𝒪]]+.\displaystyle\tilde{\mathcal{O}}=\mathcal{O}+[\eta,\mathcal{O}]+[\eta,\mathcal{O}]+\frac{1}{2}[\eta,[\eta,\mathcal{O}]]+\cdots. (26)

A relation 𝒪~𝒪\braket{\tilde{\mathcal{O}}}\approx\braket{\mathcal{O}} is valid for h4/J1h_{4}/J\ll 1. I can thus basically identify 𝒪\mathcal{O} and 𝒪~\tilde{\mathcal{O}} but their small discrepancy, [η1,𝒪][\eta_{1},\mathcal{O}], would affect dynamics of spin chains (see Appendix. B).

III.3 Symmetries

I showed that the fourfold screw field yields the bond alternation instead of the twofold staggered field. Actually, the bond-centered inversion symmetry of the spin chain (3) forbids the twofold staggered field from emerging in the effective Hamiltonian (20).

The uniform spin chain is symmetric under two types of spatial inversions: the site-centered inversion IsI_{s} and the bond-centered inversion IbI_{b}. These spatial inversions act on spins as Is:𝑺j𝑺jI_{s}:\bm{S}_{j}\mapsto\bm{S}_{-j} and Ib:𝑺j𝑺1jI_{b}:\bm{S}_{j}\mapsto\bm{S}_{1-j}. The twofold staggered field is invariant under IsI_{s} but not under IbI_{b}. On the other hand, the bond alternation and the fourfold screw field are invariant under IbI_{b} but not under IsI_{s}. In general, a low-energy effective Hamiltonian keeps the symmetries that the original Hamiltonian possesses. In this sense, the low-energy effective Hamiltonian of the spin chain (3) cannot have the twofold staggered field term that breaks the bond-centered inversion symmetry of the original Hamiltonian (14).

IV Interacting boson theory

The second-order perturbation turned out to give rise to the bond alternation in the low-energy effective Hamiltonian of the XY model in the screw field (14). However, the free spinless fermion theory does not explain the power-law behavior of the excitation gap. In this section, I present a simple theoretical explanation for the numerically found power law, incorporating the interaction of spinless fermions.

Discussions in the previous section prompt us to make an ansatz that the low-energy effective Hamiltonian of the HAFM model in the fourfold screw field (3) should be

~4\displaystyle\tilde{\mathcal{H}}_{4} :=Peη4eηP\displaystyle:=Pe^{\eta}\mathcal{H}_{4}e^{-\eta}P
=Jj𝑺j𝑺j+1\displaystyle=J\sum_{j}\bm{S}_{j}\cdot\bm{S}_{j+1}
+Jδj(1)j(SjxSj+1x+SjySj+1y).\displaystyle\qquad+J\delta_{\perp}\sum_{j}(-1)^{j}(S_{j}^{x}S_{j+1}^{x}+S_{j}^{y}S_{j+1}^{y}). (27)

Here, the effective bond alternation is characterized by the parameter δ(h4/J)2\delta_{\perp}\propto(h_{4}/J)^{2}.

Let us investigate whether the ansatz (27) explains numerical results. For small enough h4/Jh_{4}/J, one can bosonize the spin operator Affleck and Haldane (1987),

Sjz\displaystyle S_{j}^{z} =12πxϕ+(1)ja1sin(2πϕ),\displaystyle=\frac{1}{\sqrt{2\pi}}\partial_{x}\phi+(-1)^{j}a_{1}\sin(\sqrt{2\pi}\phi), (28)
Sj+\displaystyle S_{j}^{+} =ei2πθ[(1)jb0+b1sin(2πϕ)],\displaystyle=e^{-i\sqrt{2\pi}\theta}\bigl{[}(-1)^{j}b_{0}+b_{1}\sin(\sqrt{2\pi}\phi)\bigr{]}, (29)

where Sj+=Sjx+iSjyS_{j}^{+}=S_{j}^{x}+iS_{j}^{y} is the ladder operator. Coefficients a1a_{1}, b0b_{0}, b1b_{1} depend on details of the lattice model and are thus nonuniversal. They are numerically estimated Hikihara and Furusaki (2004). The Hamiltonian is then bosonized as

~4\displaystyle\tilde{\mathcal{H}}_{4} =v2𝑑x{(xθ)2+(xϕ)2}\displaystyle=\frac{v}{2}\int dx\bigl{\{}(\partial_{x}\theta)^{2}+(\partial_{x}\phi)^{2}\bigr{\}}
+dxyJδ𝑑xcos(2πϕ).\displaystyle\qquad+d_{xy}J\delta_{\perp}\int dx\cos(\sqrt{2\pi}\phi). (30)

Here, vv is the spinon velocity and the coefficient dxyd_{xy} is a nonuniversal constant Takayoshi and Sato (2010); Hikihara et al. (2017). This bosonic field theory (30) is interacting but, fortunately, integrable Essler and Konik (2005).

The lowest-energy excitation gap of the sine-Gordon model (30) is exactly given by Lukyanov and Zamolodchikov (1997); Zamolodchikov (1995); Eßler (1999)

Δ\displaystyle\Delta =2vπΓ(1/6)Γ(2/3)(dxyπJ2vΓ(3/4)Γ(1/4)δ)2/3.\displaystyle=\frac{2v}{\sqrt{\pi}}\frac{\Gamma(1/6)}{\Gamma(2/3)}\biggl{(}\frac{d_{xy}\pi J}{2v}\frac{\Gamma(3/4)}{\Gamma(1/4)}\delta_{\perp}\biggr{)}^{2/3}. (31)

I thus find

Δδ2/3(h4J)4/3.\displaystyle\Delta\propto\delta_{\perp}^{2/3}\propto\biggl{(}\frac{h_{4}}{J}\biggr{)}^{4/3}. (32)

The power 4/34/3 shows an excellent agreement with the numerical estimation (4).

The sine-Gordon theory also explains the power-law behavior of the transverse dimer order (7). In terms of the sine-Gordon theory, the transverse dimer order is an average of the vertex operator Lukyanov and Zamolodchikov (1997),

D\displaystyle D_{\perp} =dxycos(2πϕ)\displaystyle=d_{xy}\braket{\cos(\sqrt{2\pi}\phi)}
=dxy[ΔπΓ(2/3)vΓ(1/6)]1/2exp[0dtt{12e2t\displaystyle=d_{xy}\biggl{[}\frac{\Delta\sqrt{\pi}\Gamma(2/3)}{v\Gamma(1/6)}\biggr{]}^{1/2}\exp\biggl{[}\int_{0}^{\infty}\frac{dt}{t}\biggl{\{}-\frac{1}{2}e^{-2t}
+sinh2(t/2)2sinh(t/4)sinhtcosh(3t/4)}].\displaystyle\qquad+\frac{\sinh^{2}(t/2)}{2\sinh(t/4)\sinh t\cosh(3t/4)}\biggr{\}}\biggr{]}. (33)

It immediately follows from Eq. (32) that

DΔ1/2(h4J)2/3.\displaystyle D_{\perp}\propto\Delta^{1/2}\propto\biggl{(}\frac{h_{4}}{J}\biggr{)}^{2/3}. (34)

The power 2/32/3 also agrees excellenetly with the numerical estimation (7).

The bosonization approach predicts the same power law for the longitudinal dimer order. When I naively bosonizes the operator (1)jSjzSj+1z(-1)^{j}S_{j}^{z}S_{j+1}^{z}, I obtain

(1)jSjzSj+1z\displaystyle(-1)^{j}S_{j}^{z}S_{j+1}^{z} a12πxϕ(x)sin[2πϕ(x+a)].\displaystyle\approx\frac{a_{1}}{\sqrt{2\pi}}\partial_{x}\phi(x)\sin[\sqrt{2\pi}\phi(x+a)]. (35)

An operator-product expansion on the right hand side Cardy (1996) yields a more relevant interaction Starykh et al. (2005),

(1)jSjzSj+1z\displaystyle(-1)^{j}S_{j}^{z}S_{j+1}^{z} dzcos[2πϕ(x)]\displaystyle\approx d_{z}\cos[\sqrt{2\pi}\phi(x)]
+aa12πxϕ(x)sin[2πϕ(x)]+.\displaystyle\qquad+\frac{aa_{1}}{\sqrt{2\pi}}\partial_{x}\phi(x)\sin[\sqrt{2\pi}\phi(x)]+\cdots. (36)

Here, dzd_{z} is a nonuniversal constant and precisely estimated Takayoshi and Sato (2010); Hikihara et al. (2017). Note that dzd_{z} and dxyd_{xy} satisfy the following relation for small h4/Jh_{4}/J Takayoshi and Sato (2010); Hikihara et al. (2017),

2dz\displaystyle 2d_{z} =dxy,\displaystyle=d_{xy}, (37)

which reflects the SU(2) symmetry of the exchange interaction. The bosonization formula (36) indicates

D\displaystyle D_{\parallel} =dzcos(2πϕ)(h4J)2/3.\displaystyle=d_{z}\braket{\cos(\sqrt{2\pi}\phi)}\propto\biggl{(}\frac{h_{4}}{J}\biggr{)}^{2/3}. (38)

Nevertheless, the DMRG result (Fig. 4) implies Eq. (8). This discrepancy remains unclear unfortunately. This will be because the low-energy Hamiltonian fails to capture the uniform fzf^{z} order (10) properly.

Refer to caption
Figure 6: The transverse dimer order parameter (5) for system sizes L=100,200,400,800L=100,200,400,800 and its extrapolated value to the L+L\to+\infty limited are plotted. The error in the extrapolation is estimated as <1×103%<1\times 10^{-3}~{}\%. The twofold staggered field h2=0.8h0h_{2}=0.8h_{0} and the screw field h4=0.4h0h_{4}=0.4h_{0} are increased linearly with the uniform magnetic field h0h_{0}.
Refer to caption
Figure 7: The longitudinal dimer order parameter (6) for system sizes L=100,200,400,800L=100,200,400,800 and its extrapolated value to L+L\to+\infty are plotted. The error in the extrapolation is estimated as <1×103%<1\times 10^{-3}~{}\%. Parameters (α2,α4)=(0.8,0.4)(\alpha_{2},\alpha_{4})=(0.8,0.4) are used.
Refer to caption
Figure 8: The Néel order parameter (51) for system sizes L=100,200,400,800L=100,200,400,800, and its extrapolated value to L+L\to+\infty are plotted. The error in the extrapolation is estimated as <1×106%<1\times 10^{-6}~{}\%. Parameters (α2,α4)=(0.8,0.4)(\alpha_{2},\alpha_{4})=(0.8,0.4) are used.

V Coexistence of Néel and dimer orders

V.1 Renormalization groups

On the basis of the fact that the fourfold screw field induces the transverse dimer order (34), here, I investigate the realistic case with h2h_{2} and h4h_{4} proportional to the uniform field h0h_{0}. I assume two proportional coefficients

α2=h2/h0,α4=h4/h0,\displaystyle\alpha_{2}=h_{2}/h_{0},\qquad\alpha_{4}=h_{4}/h_{0}, (39)

are both constant. DMRG results for the dimer order parameters and the Néel order parameter are shown in Figs. 6, 7, and 8 for (α2,α4)=(0.8,0.4)(\alpha_{2},\alpha_{4})=(0.8,0.4).

To understand DMRG results, I replace the Hamiltonian (1) to the low-energy effective Hamiltonian,

~\displaystyle\tilde{\mathcal{H}} =Jj𝑺j𝑺j+1h0jSjzh2j(1)jSjx\displaystyle=J\sum_{j}\bm{S}_{j}\cdot\bm{S}_{j+1}-h_{0}\sum_{j}S_{j}^{z}-h_{2}\sum_{j}(-1)^{j}S_{j}^{x}
+Jδj(1)j(SjxSj+1x+SjySj+1y)\displaystyle\qquad+J\delta_{\perp}\sum_{j}(-1)^{j}(S_{j}^{x}S_{j+1}^{x}+S_{j}^{y}S_{j+1}^{y}) (40)

with δ(h4/J)2\delta_{\perp}\propto(h_{4}/J)^{2}. I can immediately bosonize it.

~\displaystyle\tilde{\mathcal{H}} =v2𝑑x{(xθ)2+(xϕ)2}h02π𝑑xxϕ\displaystyle=\frac{v}{2}\int dx\bigl{\{}(\partial_{x}\theta)^{2}+(\partial_{x}\phi)^{2}\bigr{\}}-\frac{h_{0}}{\sqrt{2\pi}}\int dx\,\partial_{x}\phi
g2𝑑xcos(2πθ)+g4𝑑xcos(2πϕ),\displaystyle\qquad-g_{2}\int dx\,\cos(\sqrt{2\pi}\theta)+g_{4}\int dx\,\cos(\sqrt{2\pi}\phi), (41)

where g2=b0h2g_{2}=b_{0}h_{2} and g4=dxyJδg_{4}=d_{xy}J\delta_{\perp}. This complex Hamiltonian consists of two parts. The first line of Eq. (41) favors the gapless TL-liquid ground state for small h0/Jh_{0}/J. The second line represents potential terms of ϕ\phi and θ\theta that give rise to an excitation gap. In general, the scaling dimensions of cos(2πθ)\cos(\sqrt{2\pi}\theta) and cos(2πϕ)\cos(\sqrt{2\pi}\phi) are 1/4K1/4K and KK, respectively. Here, KK is a parameter called the Luttinger parameter that signifies strength of interactions Giamarchi (2004). The XY and the Heisenberg chains have K=1K=1 and K=1/2K=1/2, respectively. In the latter case, two cosine interactions are equally relevant. Therefore, Néel and dimer orders can coexist in the ground state from the viewpoint of the renormalization group (RG).

The coupling constant g2g_{2} of cos(2πθ)\cos(\sqrt{2\pi}\theta), whose bare value is b0h2b_{0}h_{2}, is increasing in the course of iterative RG transformations. g2g_{2} follows the RG equation,

dg2()d\displaystyle\frac{dg_{2}(\ell)}{d\ell} 32g2().\displaystyle\approx\frac{3}{2}g_{2}(\ell). (42)

Here, \ell characterizes the effective short-distance cutoff a()=aea(\ell)=ae^{\ell}. Note that aa is the lattice spacing which was set to be unity. The RG transformation of Eq. (42) is terminated when a()a(\ell) reaches a correlation length of the lowest-energy excitation, v/Δv/\Delta.

Despite the same value of scaling dimensions, behaviors of the RG transformation of g4g_{4} differ from that of g2g_{2}. This is due to the Zeeman energy which competes with the transverse bond alternation cos(2πϕ)\cos(\sqrt{2\pi}\phi). I can absorb the Zeeman energy in Eq. (41) into the kinetic term by shifting ϕϕ+h0/2πv\phi\to\phi+h_{0}/\sqrt{2\pi}v. The ϕ\phi shift introduces an incommensurate oscillation to the transverse bond alternation term, cos(2πϕ)cos(2πϕ+h0x/v)\cos(\sqrt{2\pi}\phi)\to\cos(\sqrt{2\pi}\phi+h_{0}x/v). When the wave length v/h0v/h_{0} is much longer than the short-distance cutoff a()a(\ell), the incommensurate oscillation is negligible and then the RG equation of g4g_{4} is simply

dg4()d\displaystyle\frac{dg_{4}(\ell)}{d\ell} 32g4().\displaystyle\approx\frac{3}{2}g_{4}(\ell). (43)

Otherwise, the incommensurate oscillation is rapid enough to eliminate g4g_{4}:

g4()=0.\displaystyle g_{4}(\ell)=0. (44)

The same argument can be found in Ref. Affleck and Oshikawa (1999).

Now I can classify into two cases the strong-coupling limit that the RG flow eventually reaches. (i) When v/Δv/h0v/\Delta\ll v/h_{0}, the coupling constants g2()g_{2}(\ell) and g4()g_{4}(\ell) grow equally following Eqs. (42) and (43) and eventually reach O(1)O(1). Then the Néel and the dimer orders coexist in the ground state. (ii) When v/Δv/h0v/\Delta\gg v/h_{0}, the coupling constant g4()g_{4}(\ell) vanishes because of the rapid incommensurate oscillation [Eq. (44)]. Then, the ground state has only the Néel order.

The correlation length v/Δv/\Delta and the wave length v/h0v/h_{0} are easily compared. If α2/α4=0\alpha_{2}/\alpha_{4}=0, the gap becomes Δ(h4/J)2/3(h0/J)4/3\Delta\propto(h_{4}/J)^{2/3}\propto(h_{0}/J)^{4/3}. When h0/J1h_{0}/J\ll 1, the gap Δ(h0/J)4/3\Delta\propto(h_{0}/J)^{4/3} never exceeds h0/Jh_{0}/J, in other words, v/Δv/h0v/\Delta\gg v/h_{0}. Then the ground state does not have the dimer order. On the other hand, if α2/α4\alpha_{2}/\alpha_{4} is finite, the gap is a complex function of h0/Jh_{0}/J. Still, in the limit h0/J0h_{0}/J\to 0, the gap is reduced to the simple form of Δ(h0/J)2/3\Delta\propto(h_{0}/J)^{2/3}, which is much larger than h0/Jh_{0}/J. In other words, v/Δv/h0v/\Delta\ll v/h_{0} is valid and the first scenario comes true. Therefore, finite |α2/α4||\alpha_{2}/\alpha_{4}| is necessary for the coexistence of the Néel and the dimer orders.

V.2 Non-Abelian bosonization

There is one remaining problem in the RG analysis on the coexistence of the Néel and the dimer orders. The transverse Néel order (1)jSjxcos(2πθ)(-1)^{j}S_{j}^{x}\approx\cos(\sqrt{2\pi}\theta) and the transverse dimer order cos(2πϕ)\cos(\sqrt{2\pi}\phi) seem to compete with each other since ϕ\phi and θ\theta are noncommutative. However, this competition is an artifact of the Abelian bosonization and these orders are cooperative Garate and Affleck (2010); Chan et al. (2017); Jin and Starykh (2017).

To avoid the artifact, I rewrite the Hamiltonian (40) as

~\displaystyle\tilde{\mathcal{H}} =Jj{1+2δ3(1)j}𝑺j𝑺j+1h2j(1)jSjx\displaystyle=J\sum_{j}\biggl{\{}1+\frac{2\delta_{\perp}}{3}(-1)^{j}\biggr{\}}\bm{S}_{j}\cdot\bm{S}_{j+1}-h_{2}\sum_{j}(-1)^{j}S_{j}^{x}
Jδ3j(1)j(2SjzSj+1zSjxSj+1xSjySj+1y),\displaystyle\qquad-\frac{J\delta_{\perp}}{3}\sum_{j}(-1)^{j}(2S_{j}^{z}S_{j+1}^{z}-S_{j}^{x}S_{j+1}^{x}-S_{j}^{y}S_{j+1}^{y}), (45)

where I assume finite α2/α4\alpha_{2}/\alpha_{4}. According to the RG analysis, the uniform Zeeman energy is negligible for finite α2/α4\alpha_{2}/\alpha_{4}. Here, I simply put h0=0h_{0}=0 from the beginning. Note that the second line of Eq. (45) yields only irrelevant interactions for the relation (37) and is discarded hereafter.

Instead of the Abelian bosonization, I employ the non-Abelian bosonization approach Affleck and Haldane (1987); Gogolin et al. (2004). In the non-Abelian bosonization language, the effective Hamiltonian (41) for h0=0h_{0}=0 is written as

~\displaystyle\tilde{\mathcal{H}} =2πv3𝑑x(𝑱R𝑱R+𝑱L𝑱L)\displaystyle=\frac{2\pi v}{3}\int dx\,(\bm{J}_{R}\cdot\bm{J}_{R}+\bm{J}_{L}\cdot\bm{J}_{L})
+dxyJδ3𝑑xtr(g)ib0h22𝑑xtr(gσx).\displaystyle\qquad+\frac{d_{xy}J\delta_{\perp}}{3}\int dx\,\operatorname{tr}(g)-i\frac{b_{0}h_{2}}{2}\int dx\,\operatorname{tr}(g\sigma^{x}). (46)

Here, the spin operator 𝑺j\bm{S}_{j} is represented as

𝑺j=𝑱R+𝑱Lib02tr(g𝝈),\displaystyle\bm{S}_{j}=\bm{J}_{R}+\bm{J}_{L}-\frac{ib_{0}}{2}\operatorname{tr}(g\bm{\sigma}), (47)

with SU(2)\mathrm{SU(2)} currents 𝑱R\bm{J}_{R} and 𝑱L\bm{J}_{L}, a fundamental field gSU(2)g\in\mathrm{SU(2)}, the Pauli matrices 𝝈=(σxσyσz)T\bm{\sigma}=(\sigma^{x}\,\sigma^{y}\,\sigma^{z})^{T} Affleck and Haldane (1987); Gogolin et al. (2004). The matrix gSU(2)g\in\mathrm{SU(2)} is simply related to the U(1) bosons,

g\displaystyle g =(ei2πϕiei2πθiei2πθei2πϕ).\displaystyle=\begin{pmatrix}e^{i\sqrt{2\pi}\phi}&ie^{-i\sqrt{2\pi}\theta}\\ ie^{i\sqrt{2\pi}\theta}&e^{-i\sqrt{2\pi}\phi}\end{pmatrix}. (48)

Since global rotations keep the excitation spectrum unchanged, I perform a global π/2\pi/2 rotation in the spin space as (σx,σy,σz)(σz,σy,σx)(\sigma^{x},\,\sigma^{y},\,\sigma^{z})\to(\sigma^{z},\,\sigma^{y},\,-\sigma^{x}). The rotation transforms the Hamiltonian (46) into

~\displaystyle\tilde{\mathcal{H}} =2πv3𝑑x(𝑱R𝑱R+𝑱L𝑱L)\displaystyle=\frac{2\pi v}{3}\int dx\,(\bm{J}_{R}\cdot\bm{J}_{R}+\bm{J}_{L}\cdot\bm{J}_{L})
+dxyJδ3𝑑xtr(g)ib0h22𝑑xtr(gσz).\displaystyle\qquad+\frac{d_{xy}J\delta_{\perp}}{3}\int dx\,\operatorname{tr}(g)-i\frac{b_{0}h_{2}}{2}\int dx\,\operatorname{tr}(g\sigma^{z}). (49)

Translating it to the Abelian bosonizaion language, I can express this Hamiltonian as

~\displaystyle\tilde{\mathcal{H}} =v2𝑑x{(xθ)2+(xϕ)2}\displaystyle=\frac{v}{2}\int dx\bigl{\{}(\partial_{x}\theta)^{2}+(\partial_{x}\phi)^{2}\bigr{\}}
+2dxyJδ3𝑑xcos(2πϕ)b0h2𝑑xsin(2πϕ)\displaystyle\quad+\frac{2d_{xy}J\delta_{\perp}}{3}\int dx\cos(\sqrt{2\pi}\phi)-b_{0}h_{2}\int dx\,\sin(\sqrt{2\pi}\phi)
=v2𝑑x{(xθ)2+(xϕ)2}+g𝑑xcos(2πϕ+α),\displaystyle=\frac{v}{2}\int dx\bigl{\{}(\partial_{x}\theta)^{2}+(\partial_{x}\phi)^{2}\bigr{\}}+g\int dx\cos(\sqrt{2\pi}\phi+\alpha), (50)

with the coupling constant g=(2dxyJδ/3)2+(b0h2)2g=\sqrt{(2d_{xy}J\delta_{\perp}/3)^{2}+(b_{0}h_{2})^{2}} and the phase shift α=tan1(3b0h2/2dxyJδ)\alpha=\tan^{-1}(3b_{0}h_{2}/2d_{xy}J\delta_{\perp}). The complex model (40) is thus reduced to the simple sine-Gordon model (50).

The incommensurate phase shift α\alpha realizes the coexistence of the Néel order,

Nx:=j(1)jSjx/L,\displaystyle N_{x}:=\sum_{j}(-1)^{j}\braket{S_{j}^{x}}/L, (51)

and the transverse dimer order (5). Their ground-state averages are given by

Nx\displaystyle N_{x} Δ/Jsinα(g/J)1/3sinα,\displaystyle\propto\sqrt{\Delta/J}\sin\alpha\propto(g/J)^{1/3}\sin\alpha, (52)
D\displaystyle D_{\perp} Δ/Jcosα(g/J)1/3cosα.\displaystyle\propto\sqrt{\Delta/J}\cos\alpha\propto(g/J)^{1/3}\cos\alpha. (53)

For h0/J0h_{0}/J\to 0, the angle α\alpha approaches π/2\pi/2. Therefore, at low fields h0/J1h_{0}/J\ll 1, NxN_{x} and DD_{\perp} are expected to follow power laws:

Nx\displaystyle N_{x} (h0/J)1/3,D(h0/J)4/3.\displaystyle\propto(h_{0}/J)^{1/3},\quad D_{\perp}\propto(h_{0}/J)^{4/3}. (54)

The power law (54) is qualitatively consistent with Figs. 6 and 8 at low fields. For weak magnetic fields, the longitudinal and the transverse dimer order parameters are almost equal. On the other hand, they are much smaller than the Néel order parameter NxN_{x}. If we assume δ=Ch42\delta_{\perp}=C_{\perp}{h_{4}}^{2}, the bosonized theory (50) predicts a ratio D/NxD_{\perp}/N_{x} given by

DNx\displaystyle\frac{D_{\perp}}{N_{x}} =dxyb0cotα\displaystyle=\frac{d_{xy}}{b_{0}}\cot\alpha
=2C3(dxyb0)2α42α2h0J.\displaystyle=\frac{2C_{\perp}}{3}\biggl{(}\frac{d_{xy}}{b_{0}}\biggr{)}^{2}\frac{{\alpha_{4}}^{2}}{\alpha_{2}}\frac{h_{0}}{J}. (55)

The right hand side is approximately estimated as 0.028Ch0/J0.028C_{\perp}h_{0}/J for the parameters used in DMRG. If C0.7C_{\perp}\approx 0.7, the ratio (55) is roughly consistent with the DMRG data of Figs. 6 and 8 for h0/J<0.3h_{0}/J<0.3.

VI Experimental relevance

VI.1 [Cu(pym)(H2O)4]SiF6{}_{6}\cdot H2O

The model (1) that I have dealt with so far is similar to a model proposed for the spin-chain compound [Cu(pym)(H2O)4]SiF6H2O\mathrm{[Cu(pym)(H_{2}O)_{4}]SiF_{6}\cdot H_{2}O} with the following Hamiltonian: Liu et al. (2019)

exp\displaystyle\mathcal{H}_{\rm exp} =Jj(SjxSj+1x+SjySj+1y+λSjzSj+1z)\displaystyle=J\sum_{j}(S_{j}^{x}S_{j+1}^{x}+S_{j}^{y}S_{j+1}^{y}+\lambda S_{j}^{z}S_{j+1}^{z})
+h0jSjx+h2j(1)jSjy+h4jδjSjz\displaystyle\qquad+h_{0}\sum_{j}S_{j}^{x}+h_{2}\sum_{j}(-1)^{j}S_{j}^{y}+h_{4}\sum_{j}\delta_{j}S_{j}^{z}
+Duj(SjxSj+1ySjySj+1x),\displaystyle\qquad+D_{u}\sum_{j}(S_{j}^{x}S_{j+1}^{y}-S_{j}^{y}S_{j+1}^{x}), (56)

where λ1\lambda\approx 1. There are three differences in two models (1) and (56): field directions, the weak exchange anisotropy, and the uniform DM interaction. In this section, I investigate effects of these differences one by one and discuss an experimental feasibility of the field-induced transverse dimer order.

VI.2 Field directions

In the model (56), the magnetic field h0h_{0}, the twofold staggered field h2h_{2}, and the fourfold screw field h4h_{4} are applied in different directions. On the other hand, the model (1) has the uniform field and the fourfold screw field in the same direction. This difference in field directions is actually insignificant in low-energy physics for small h0/Jh_{0}/J. When the uniform DM interaction is absent (Du=0D_{u}=0), the model (56) has the following low-energy effective Hamiltonian:

~exp\displaystyle\tilde{\mathcal{H}}_{\rm exp} Jj(SjxSj+1x+SjySj+1y+λSjzSj+1z)\displaystyle\approx J\sum_{j}(S_{j}^{x}S_{j+1}^{x}+S_{j}^{y}S_{j+1}^{y}+\lambda S_{j}^{z}S_{j+1}^{z})
+h0jSjx+h2j(1)jSjy\displaystyle\qquad+h_{0}\sum_{j}S_{j}^{x}+h_{2}\sum_{j}(-1)^{j}S_{j}^{y}
+Jδj(1)j(SjxSj+1x+SjySj+1y),\displaystyle\qquad+J\delta_{\perp}\sum_{j}(-1)^{j}(S_{j}^{x}S_{j+1}^{x}+S_{j}^{y}S_{j+1}^{y}), (57)

with δ(h4/J)2\delta_{\perp}\propto(h_{4}/J)^{2}. Relabeling the spins (Sjx,Sjy,Sjz)(Sjz,Sjx,Sjy)(S_{j}^{x},\,S_{j}^{y},\,S_{j}^{z})\to(S_{j}^{z},\,S_{j}^{x},\,S_{j}^{y}), I rewrite this Hamiltonian as

~exp\displaystyle\tilde{\mathcal{H}}_{\rm exp} Jj(SjxSj+1x+λSjySj+1y+SjzSj+1z)\displaystyle\approx J\sum_{j}(S_{j}^{x}S_{j+1}^{x}+\lambda S_{j}^{y}S_{j+1}^{y}+S_{j}^{z}S_{j+1}^{z})
+h0jSjz+h2j(1)jSjx\displaystyle\qquad+h_{0}\sum_{j}S_{j}^{z}+h_{2}\sum_{j}(-1)^{j}S_{j}^{x}
+2Jδ3j(1)j𝑺j𝑺j+1\displaystyle\qquad+\frac{2J\delta_{\perp}}{3}\sum_{j}(-1)^{j}\bm{S}_{j}\cdot\bm{S}_{j+1}
Jδ3j(2SjySj+1ySjxSj+1xSjzSj+1z).\displaystyle\qquad-\frac{J\delta_{\perp}}{3}\sum_{j}(2S_{j}^{y}S_{j+1}^{y}-S_{j}^{x}S_{j+1}^{x}-S_{j}^{z}S_{j+1}^{z}). (58)

Since the last term of Eq. (58) is irrelevant, the bosonized Hamiltonian of Eq. (58) is given by

~exp\displaystyle\tilde{\mathcal{H}}_{\rm exp} =v2𝑑x{(xθ)2+(xϕ)2}+h02π𝑑xxϕ\displaystyle=\frac{v}{2}\int dx\bigl{\{}(\partial_{x}\theta)^{2}+(\partial_{x}\phi)^{2}\bigr{\}}+\frac{h_{0}}{\sqrt{2\pi}}\int dx\,\partial_{x}\phi
+g2𝑑xcos(2πθ)+g4𝑑xcos(2πϕ)\displaystyle\qquad+g_{2}\int dx\,\cos(\sqrt{2\pi}\theta)+g_{4}\int dx\,\cos(\sqrt{2\pi}\phi)
+ga𝑑xsin(8πθ),\displaystyle\qquad+g_{a}\int dx\,\sin(\sqrt{8\pi}\theta), (59)

with gaJ(λ1)g_{a}\propto J(\lambda-1). Except for the last term that comes from the exchange anisotropy, the Hamiltonian (59) is identical to the one (41) investigated in Sec. V.

VI.3 Exchange anisotropy

The bosonized effective Hamiltonian (59) shows that the small exchange anisotropy λ1\lambda\approx 1 gives rise to the sin(8πθ)\sin(\sqrt{8\pi}\theta) interaction. Though this interaction iteself can be marginally relevant at most in the RG sense, it is negligible in the presence of the much more relevant interaction cos(2πθ)\cos(\sqrt{2\pi}\theta).

VI.4 Uniform DM interaction

After all, the uniform DM interaction is the only significant difference in the models (1) and (56). The major effect of the uniform DM interaction is a chiral rotation. Let us resurrect the uniform DM intearction in the rotated effective Hamiltonian (58):

~exp\displaystyle\tilde{\mathcal{H}}_{\rm exp} Jj(SjxSj+1x+λSjySj+1y+SjzSj+1z)\displaystyle\approx J\sum_{j}(S_{j}^{x}S_{j+1}^{x}+\lambda S_{j}^{y}S_{j+1}^{y}+S_{j}^{z}S_{j+1}^{z})
+h0jSjz+h2j(1)jSjx\displaystyle\qquad+h_{0}\sum_{j}S_{j}^{z}+h_{2}\sum_{j}(-1)^{j}S_{j}^{x}
+2Jδ3j(1)j𝑺j𝑺j+1\displaystyle\qquad+\frac{2J\delta_{\perp}}{3}\sum_{j}(-1)^{j}\bm{S}_{j}\cdot\bm{S}_{j+1}
+Duj(SjzSj+1xSjxSj+1z).\displaystyle\qquad+D_{u}\sum_{j}(S_{j}^{z}S_{j+1}^{x}-S_{j}^{x}S_{j+1}^{z}). (60)

Here, the irrelevant term is already dropped. The uniform DM interaction itself is bosonized as Gangadharaiah et al. (2008)

Duj(SjzSj+1xSjxSj+1z)\displaystyle D_{u}\sum_{j}(S_{j}^{z}S_{j+1}^{x}-S_{j}^{x}S_{j+1}^{z}) γDu𝑑x(JRyJLy),\displaystyle\approx\gamma D_{u}\int dx\,(J_{R}^{y}-J_{L}^{y}), (61)

with a nonuniversal constant γ>0\gamma>0. I employed the non-Abelian bosonization language (see Sec. V.2). The right hand side of Eq. (61) resembles the Zeeman energy,

h0jSjz\displaystyle h_{0}\sum_{j}S_{j}^{z} =h0𝑑x(JRz+JLz).\displaystyle=h_{0}\int dx\,(J_{R}^{z}+J_{L}^{z}). (62)

While the Zeeman energy (62) is non-chiral (i.e. symmetric in the permutation of RLR\leftrightarrow L), but the uniform DM interaction (61) is chiral. As far as only either the RR part or the LL part is concerned, I cannot distinguish the Zeeman energy and the uniform DM interaction.

A chiral rotation can combine the uniform DM interaction (61) and the Zeeman energy (62Garate and Affleck (2010); Jin and Starykh (2017); Chan et al. (2017).

𝑱ν\displaystyle\bm{J}_{\nu} =(θν)𝑴ν,\displaystyle=\mathcal{R}(\theta_{\nu})\bm{M}_{\nu}, (63)

for ν=R,L\nu=R,L. Here, the rotation R(θν)R(\theta_{\nu}) is defined as

R(θν)\displaystyle R(\theta_{\nu}) =(1000cosθνsinθν0sinθνcosθν).\displaystyle=\begin{pmatrix}1&0&0\\ 0&\cos\theta_{\nu}&\sin\theta_{\nu}\\ 0&-\sin\theta_{\nu}&\cos\theta_{\nu}\end{pmatrix}. (64)

I assume that h0h_{0} and γDu\gamma D_{u} are both positive. Then, the rotation leads to

h0(JRz+JLz)+γDu(JRyJLy)\displaystyle h_{0}(J_{R}^{z}+J_{L}^{z})+\gamma D_{u}(J_{R}^{y}-J_{L}^{y}) =tϕ(MRz+MLz),\displaystyle=t_{\phi}(M_{R}^{z}+M_{L}^{z}), (65)

with

tϕ=h02+(γDu)2,\displaystyle t_{\phi}=\sqrt{{h_{0}}^{2}+(\gamma D_{u})^{2}}, (66)

if θR\theta_{R} and θL\theta_{L} take the following values,

θR\displaystyle\theta_{R} =tan1(γDuh0),\displaystyle=\tan^{-1}\biggl{(}\frac{\gamma D_{u}}{h_{0}}\biggr{)}, (67)
θL\displaystyle\theta_{L} =θR.\displaystyle=-\theta_{R}. (68)

The chiral rotation R(θν)R(\theta_{\nu}) transforms gg into

g=eiσxθL/2geiσxθR/2.\displaystyle g^{\prime}=e^{i\sigma^{x}\theta_{L}/2}ge^{-i\sigma^{x}\theta_{R}/2}. (69)

Effects of this chiral rotation on the low-energy Hamiltonian are explained in Appendix. B. Here, I show results only. The chirally rotated Hamiltonian then becomes

~exp\displaystyle\tilde{\mathcal{H}}_{\rm exp}
=v2𝑑x{(xΘ)2+(xΦ)2}+tϕ2π𝑑xxΦ\displaystyle=\frac{v}{2}\int dx\bigl{\{}(\partial_{x}\Theta)^{2}+(\partial_{x}\Phi)^{2}\bigr{\}}+\frac{t_{\phi}}{\sqrt{2\pi}}\int dx\,\partial_{x}\Phi
+g2𝑑x{cos(2πΘ)cosθR+cos(2πΦ)sinθR}\displaystyle\quad+g^{\prime}_{2}\int dx\,\{\cos(\sqrt{2\pi}\Theta)\cos\theta_{R}+\cos(\sqrt{2\pi}\Phi)\sin\theta_{R}\}
+g4𝑑x{cos(2πΦ)cosθRcos(2πΘ)sinθR},\displaystyle\quad+g^{\prime}_{4}\int dx\,\{\cos(\sqrt{2\pi}\Phi)\cos\theta_{R}-\cos(\sqrt{2\pi}\Theta)\sin\theta_{R}\}, (70)

where g2h2g^{\prime}_{2}\propto h_{2} and g4Jδg^{\prime}_{4}\propto J\delta_{\perp}. Note that the chiral rotation (69) mixes the Néel and dimer orders. Though the right hand side of Eq. (70) is complex, its basic structure is the same as that of Eq. (41) in a sense that coupling constants g2g^{\prime}_{2} and g4g^{\prime}_{4} in the former follow the same RG equations as those for g2g_{2} and g4g_{4} in the latter. Following the argument in Sec. V.2, I obtain the Néel and dimer orders in the ground state (see Appendix B for details):

Nx\displaystyle N_{x} (G/J)1/3sinα,\displaystyle\propto(G/J)^{1/3}\sin\alpha^{\prime}, (71)
D\displaystyle D_{\perp} (G/J)1/3cosα,\displaystyle\propto(G/J)^{1/3}\cos\alpha^{\prime}, (72)

where the coupling constant GG and the angle α\alpha^{\prime} are defined in Eqs. (133) and (134). In analogy with Eq. (54), I obtain

Nx(h0/J)1/3,D(h0/J)4/3,\displaystyle N_{x}\propto(h_{0}/J)^{1/3},\quad D_{\perp}\propto(h_{0}/J)^{4/3}, (73)

at low fields h0/J1h_{0}/J\ll 1.

In short, the uniform DM interaction causes the chiral rotation that mixes the Néel and the transverse dimer orders if the following condition is met.

(h0/J)2/3tϕ/J.\displaystyle(h_{0}/J)^{2/3}\gg t_{\phi}/J. (74)

When Du=0D_{u}=0, the condition (74) is trivially satisfied for small h0/Jh_{0}/J. However, the inequality (74) can be violated at extremely small magnetic fields h0/Du1h_{0}/D_{u}\ll 1.

Let me comment on effects of the uniform DM interaction on electron spin resonance (ESR). In one-dimensional quantum spin systems, the uniform DM interaction splits the ESR peak that corresponds to the Zeeman energy [Eqs. (61) and (62)] Povarov et al. (2011); Furuya (2017). In some cases, the DM interaction changes selection rules of ESR and yields an additional resonance that occurs at a frequency away from the Zeeman energy Oshikawa and Affleck (2002); Furuya and Oshikawa (2012); Ozerov et al. (2015); Zhao et al. (2003); Lou et al. (2005).

The experiment Liu et al. (2019) on [Cu(pym)(H2O)4]SiF6H2O\mathrm{[Cu(pym)(H_{2}O)_{4}]SiF_{6}\cdot H_{2}O} found that ESR peaks of this compound exhibit unconventional power-law dependence on the magnetic field. A part of this unconventional behavior is attributed to the chiral rotation and the complex dependence of the coupling constant on the magnetic field. A derivation of ESR selection rules is described in Appendix. B. Here, I simply summarize the result. Elementary excitations of the sine-Gordon theory are a soliton, an antisoliton, and their bound states, breathers. Let us represent these excitation gaps by MM, where MM can be the soliton mass or the breather mass. ESR in the model (60) occurs when the frequency ω\omega of the applied microwave satisfies

ω=M,\displaystyle\omega=M, (75)

or

ω=tϕ2+M2.\displaystyle\omega=\sqrt{{t_{\phi}}^{2}+M^{2}}. (76)

These resonance frequencies are close to neither the Zeeman energy nor the typical gap, ω(h0/J)2/3\omega\propto(h_{0}/J)^{2/3}, in quantum spin chains with the twofold staggered field Dender et al. (1997); Zvyagin et al. (2004); Umegaki et al. (2009). In particular, the latter resonance frequency (76) approaches ωγDu\omega\to\gamma D_{u} in the h00h_{0}\to 0 limit Povarov et al. (2011); Furuya (2017).

VI.5 (Ba/Sr)Co2V2O8

I can find other quantum spin chain compounds with the fourfold screw symmetry such as BaCo2V2O8\mathrm{BaCo_{2}V_{2}O_{8}} Faure et al. (2018); Kimura et al. (2007) and SrCo2V2O8\mathrm{SrCo_{2}V_{2}O_{8}} Wang et al. (2015); Bera et al. (2017). Unlike [Cu(pym)(H2O)4]SiF6H2O\mathrm{[Cu(pym)(H_{2}O)_{4}]SiF_{6}\cdot H_{2}O}, these compounds have Ising-like exchange interactions. As I already showed, the SU(2) symmetry of the exchange interaction is essential for the coexistence of the Néel and the transverse dimer orders. The strong enough Ising anisotropy ruins the coexistence and thus makes the fourfold screw field insignificant. Thus far, most experimental results on these compounds are well understood with models without the fourfold screw field Faure et al. (2018); Kimura et al. (2007); Wang et al. (2015); Bera et al. (2017) though some ESR peaks can be attributed to the presence of the fourfold screw field Kimura et al. (2007).

VII Summary

I discussed the novel type of field-induced gap phenomena, field-induced dimer orders in quantum spin chains. The fourfold screw field with the bond-centered inversion symmetry introduces perturbatively the effective bond alternation to the spin chain. In analogy with the twofold staggered field, the fourfold screw field, which breaks the one-site translation symmetry, gives rise to the excitation gap from the ground state to the excited states.

In the first part of the paper, I applied the fourfold screw field h4h_{4} solely to quantum spin chains. The field-induced excitation gap by h4h_{4} turned out to show a distinctive power law from that by the twofold staggered field h2h_{2}. The gap is proportional to (h4/J)4/3(h_{4}/J)^{4/3} for the fourfold screw field instead of (h2/J)2/3(h_{2}/J)^{2/3} for the twofold staggered field h2h_{2} Oshikawa and Affleck (1997); Affleck and Oshikawa (1999). The power law was predicted from the quantum field theory and consistent with the numerical results (Fig. 2). The field theory also gave the explanation on the power law of the transverse dimer order (7), though it failed for the longitudinal one (8) somehow.

Next, I applied the uniform field, the twofold staggered field, and the fourfold screw field simultaneously to HAFM chains. The SU(2) symmetry of the exchange interaction turned out to make the coexistence of the Néel and dimer order possible in the ground state. The coexistence of these orders are nontrivial and already interesting Garate and Affleck (2010); Jin and Starykh (2017); Chan et al. (2017). More interestingly, the dimer order grows in association with the uniform magnetic field (Figs. 6, 7, and 8). The coexistent growth of the Néel and the dimer orders were numerically found and supported by the effective field theory.

Last but not least, I discussed the relevance of my model to experimental studies, in particular, Ref. Liu et al. (2019) on [Cu(pym)(H2O)4]SiF6H2O\mathrm{[Cu(pym)(H_{2}O)_{4}]SiF_{6}\cdot H_{2}O}. There are three differences in the model for [Cu(pym)(H2O)4]SiF6H2O\mathrm{[Cu(pym)(H_{2}O)_{4}]SiF_{6}\cdot H_{2}O} and the model Hamiltonian (1), or equivalently Eq. (40), that I dealt with in this paper. They are field directions, the weak exchange anisotropy, and the uniform DM interaction. In Sec. VI, I discussed that all the three differences do not interfere with the field-induced growth of the Néel and the dimer orders. However, the uniform DM interaction may cause nontrivial effects on dynamics of the spin chain such as ESR. In the presence of the uniform DM interaction, increase of the magnetic field rotates chirally the spin chain. This chiral rotation affects selection rules of the electron spin resonance. It will be interesting to test experimentally the coexistence of the Néel and the dimer orders in spin-chain compounds with the fourfold screw symmetry.

Acknowledgments

I thank Akira Furusaki, Yusuke Horinouchi, and Tsutomu Momoi for useful discussions.

Appendix A Derivation of effective Hamiltonian

This section is devoted to derivation of the low-energy effective Hamiltonian discussed in Sec. III.2 as generically as possible.

A.1 Framework

I consider a Hamiltonian 0\mathcal{H}_{0} whose eigenstates are exactly known.

0|ϕn\displaystyle\mathcal{H}_{0}\ket{\phi_{n}} =En|ϕn,(n=0,1,2,).\displaystyle=E_{n}\ket{\phi_{n}},\quad(n=0,1,2,\cdots). (77)

I can assume EnEmE_{n}\leq E_{m} for nmn\leq m without loss of generality. Adding a perturbation λV\lambda V, I modify the Hamiltonian to

\displaystyle\mathcal{H} =0+λV,\displaystyle=\mathcal{H}_{0}+\lambda V, (78)

where λ\lambda is a small parameter that controls the perturbation expansion. At low energies, effects of the perturbation can be taken into account as a form of the effective Hamiltonian eff\mathcal{H}_{\rm eff}. The effective Hamiltonian can be easily obtained up to the second order.

To derive the low-energy effective Hamiltonian eff\mathcal{H}_{\rm eff}, I focus on the NN low-energy eigenstates |ϕn\ket{\phi_{n}} with n=n0,n0+1,n0+2,,n0+N1n=n_{0},n_{0}+1,n_{0}+2,\cdots,n_{0}+N-1 for n00n_{0}\geq 0. One can take n0=0n_{0}=0 or n0>0n_{0}>0. The latter case is useful for later application to the free spinless fermion chain. When applying to the free spinless fermion chain, I assume |EnϵF|<W|E_{n}-\epsilon_{F}|<W for n=n0,n0+1,,n0+Nn=n_{0},n_{0}+1,\cdots,n_{0}+N, where ϵF\epsilon_{F} is the Fermi energy and W>0W>0 is an energy cutoff.

Each eigenstate defines a projection operator Pn=|ϕnϕn|P_{n}=\ket{\phi_{n}}\bra{\phi_{n}} into that eigenstate. PnP_{n} satisfies Pn0=0Pn=EnPnP_{n}\mathcal{H}_{0}=\mathcal{H}_{0}P_{n}=E_{n}P_{n}. An operator PP,

P\displaystyle P =n=n0n0+N1Pn,\displaystyle=\sum_{n=n_{0}}^{n_{0}+N-1}P_{n}, (79)

then projects an arbitrary state into the subspace spanned by the NN eigenstates. Q=1PQ=1-P projects any state into the supplementary space.

The key idea is to perform a canonical transformation of the Schrieffer-Wolff type on the Hamiltonian Schrieffer and Wolff (1966); Bravyi et al. (2011),

\displaystyle\mathcal{H}^{\prime} =eηeη,\displaystyle=e^{\eta}\mathcal{H}e^{-\eta}, (80)

where η\eta is anti-Hermitian so that eηe^{\eta} is unitary. The Schrieffer-Wolff formulation is useful in quantum spin systems Slagle and Kim (2017). Here, I briefly review the derivation of the effective Hamiltonian based on the Schrieffer-Wolff formulation to make the paper self-contained.

Two Hamiltonians \mathcal{H} and \mathcal{H}^{\prime} have one-to-one corresponding lists of eigenstates with exactly the same eigenenergies. An appropriate choice of η\eta simplifies the transformed Hamiltonian \mathcal{H}^{\prime}. I expand eηe^{\eta} and determine η\eta.

\displaystyle\mathcal{H}^{\prime} =+[η,]+12[η,[η,]]+\displaystyle=\mathcal{H}+[\eta,\mathcal{H}]+\frac{1}{2}[\eta,[\eta,\mathcal{H}]]+\cdots
=0+λ([η1,0]+V)\displaystyle=\mathcal{H}_{0}+\lambda\Bigl{(}[\eta_{1},\mathcal{H}_{0}]+V\biggr{)}
+λ2([η2,0]+[η1,V]+12[η1,[η1,0]])+.\displaystyle\quad+\lambda^{2}\biggl{(}[\eta_{2},\mathcal{H}_{0}]+[\eta_{1},V]+\frac{1}{2}[\eta_{1},[\eta_{1},\mathcal{H}_{0}]]\biggr{)}+\cdots. (81)

In the last line, I expanded η\eta around λ=0\lambda=0:

η\displaystyle\eta =p=0λpp!ηp.\displaystyle=\sum_{p=0}^{\infty}\frac{\lambda^{p}}{p!}\eta_{p}. (82)

ηp\eta_{p} is determined so that Slagle and Kim (2017)

[P,]\displaystyle[P,\mathcal{H}^{\prime}] =0.\displaystyle=0. (83)

I solve Eq. (83) at each order of λ\lambda. At the first order, Eq. (83) leads to

[P,[η1,0]]=[P,V].\displaystyle[P,[\eta_{1},\mathcal{H}_{0}]]=-[P,V]. (84)

The anti-Hermitian η1\eta_{1} that satisfies Eq. (84) is given by Slagle and Kim (2017)

η1\displaystyle\eta_{1} =n=n0n0+N1(PnV1En0QQ1En0VPn).\displaystyle=\sum_{n=n_{0}}^{n_{0}+N-1}\biggl{(}P_{n}V\frac{1}{E_{n}-\mathcal{H}_{0}}Q-Q\frac{1}{E_{n}-\mathcal{H}_{0}}VP_{n}\biggr{)}. (85)

The second order of Eq. (83),

[η2,0]=X2,\displaystyle[\eta_{2},\mathcal{H}_{0}]=-X_{2}, (86)

with X2X_{2} being

X2\displaystyle X_{2} =[η1,V]+12[η1,[η1,V]],\displaystyle=[\eta_{1},V]+\frac{1}{2}[\eta_{1},[\eta_{1},V]], (87)

is similar to the first-order equation (84). The solution is immediately obtained.

η2\displaystyle\eta_{2} =n=n0n0+N1(PnX21En0QQ1En0X2Pn).\displaystyle=\sum_{n=n_{0}}^{n_{0}+N-1}\biggl{(}P_{n}X_{2}\frac{1}{E_{n}-\mathcal{H}_{0}}Q-Q\frac{1}{E_{n}-\mathcal{H}_{0}}X_{2}P_{n}\biggr{)}. (88)

I am now ready to write down the low-energy effective Hamiltonian,

eff\displaystyle\mathcal{H}_{\rm eff} =PP=n=0λnneff,\displaystyle=P\mathcal{H}^{\prime}P=\sum_{n=0}^{\infty}\lambda^{n}\mathcal{H}_{n}^{\rm eff}, (89)

up to the second order of λ\lambda. First three terms neff\mathcal{H}_{n}^{\rm eff} for n=0,1,2n=0,1,2 are shown below.

0eff\displaystyle\mathcal{H}_{0}^{\rm eff} =P0P,\displaystyle=P\mathcal{H}_{0}P, (90)
1eff\displaystyle\mathcal{H}_{1}^{\rm eff} =P([η1,0]+V)P\displaystyle=P\Bigl{(}[\eta_{1},\mathcal{H}_{0}]+V\Bigr{)}P
=PVP,\displaystyle=PVP, (91)

where P[ηn,0]P=0P[\eta_{n},\mathcal{H}_{0}]P=0 holds true for n=1,2n=1,2. The second-order term 2eff\mathcal{H}_{2}^{\rm eff} is given by

2eff\displaystyle\mathcal{H}_{2}^{\rm eff} =P([η2,0]+X2)P\displaystyle=P\Bigl{(}[\eta_{2},\mathcal{H}_{0}]+X_{2}\Bigr{)}P
=PX2P\displaystyle=PX_{2}P
=P([η1,V]+12[η1,[η1,0]])P.\displaystyle=P\biggl{(}[\eta_{1},V]+\frac{1}{2}[\eta_{1},[\eta_{1},\mathcal{H}_{0}]]\biggr{)}P. (92)

One can simplify the last line:

2eff\displaystyle\mathcal{H}_{2}^{\rm eff} =12n,mPn(V1En0QV+V1Em0QV)Pm.\displaystyle=\frac{1}{2}\sum_{n,m}P_{n}\biggl{(}V\frac{1}{E_{n}-\mathcal{H}_{0}}QV+V\frac{1}{E_{m}-\mathcal{H}_{0}}QV\biggr{)}P_{m}. (93)

This leads to the following effective Hamiltonian of the second order of λ\lambda:

eff\displaystyle\mathcal{H}_{\rm eff} =P(0+λV)P+λ22n,m=0n0+N1Pn(V1En0QV\displaystyle=P(\mathcal{H}_{0}+\lambda V)P+\frac{\lambda^{2}}{2}\sum_{n,m=0}^{n_{0}+N-1}P_{n}\biggl{(}V\frac{1}{E_{n}-\mathcal{H}_{0}}QV
+V1Em0QV)Pm.\displaystyle\quad+V\frac{1}{E_{m}-\mathcal{H}_{0}}QV\biggr{)}P_{m}. (94)

A.2 Application to spinless fermion chains

Here I apply the generic formalism of the effective Hamiltonian to the spinless fermion chain (15). The low-energy region is defined as Eq. (17). The operator PP is redefined as a projection operator onto the subspace (17) of the reciprocal space [Eq. (17)]. PP acts on ckc_{k} as follows. PckP=ckPc_{k}P=c_{k} for kRk\in R and PckP=0Pc_{k}P=0 otherwise. PP acts on ckc_{k}^{\dagger} in the same manner. Q=1PQ=1-P acts on ckc_{k} and ckc_{k}^{\dagger} as PckQ=QckP=PckQ=QckP=0Pc_{k}Q=Qc_{k}P=Pc_{k}^{\dagger}Q=Qc_{k}^{\dagger}P=0. In applying the generic Schrieffer-Wolff formulation to the XY chain (15), I regard 0\mathcal{H}_{0} and VV of Eq. (78) as

0\displaystyle\mathcal{H}_{0} =kϵ(k)ckck,\displaystyle=\sum_{k}\epsilon(k)c_{k}^{\dagger}c_{k}, (95)
λV\displaystyle\lambda V =h42k(eπi/4ckck+π2+eπi/4ck+π2ck).\displaystyle=-\frac{h_{4}}{\sqrt{2}}\sum_{k}\bigl{(}e^{-\pi i/4}c_{k}^{\dagger}c_{k+\frac{\pi}{2}}+e^{\pi i/4}c_{k+\frac{\pi}{2}}^{\dagger}c_{k}\bigr{)}. (96)

The effective Hamiltonian (94) is then given by

eff\displaystyle\mathcal{H}_{\rm eff} =kRϵ(k)ckck+V,\displaystyle=\sum_{k\in R}\epsilon(k)c_{k}^{\dagger}c_{k}+V^{\prime}, (97)

where V=h422effV^{\prime}=h_{4}^{2}\mathcal{H}_{2}^{\rm eff} is the second-order term. Note that the first-order term 1eff\mathcal{H}_{1}^{\rm eff} vanishes trivially because Pck+π2ckP=0Pc_{k+\frac{\pi}{2}}^{\dagger}c_{k}P=0 for any kRk\in R or kRk\not\in R thanks to the assumption Λ1\Lambda\ll 1. The second-order correction VV^{\prime} is calculated as follows.

V\displaystyle V^{\prime} =h424k,k[eπi/2Pck+π2ckQck+π2ckP(1ϵ(k)ϵ(k+π2)+1ϵ(k+π2)ϵ(k))\displaystyle=\frac{h_{4}^{2}}{4}\sum_{k,k^{\prime}}\biggl{[}e^{-\pi i/2}Pc_{k+\frac{\pi}{2}}^{\dagger}c_{k}Qc_{k^{\prime}+\frac{\pi}{2}}^{\dagger}c_{k^{\prime}}P\biggl{(}\frac{1}{\epsilon(k^{\prime})-\epsilon(k^{\prime}+\frac{\pi}{2})}+\frac{1}{\epsilon(k+\frac{\pi}{2})-\epsilon(k)}\biggr{)}
+Pck+π2ckQckck+π2P(1ϵ(k+π2)ϵ(k)+1ϵ(k+π2)ϵ(k))\displaystyle\qquad+Pc_{k+\frac{\pi}{2}}^{\dagger}c_{k}Qc_{k^{\prime}}^{\dagger}c_{k^{\prime}+\frac{\pi}{2}}P\biggl{(}\frac{1}{\epsilon(k^{\prime}+\frac{\pi}{2})-\epsilon(k^{\prime})}+\frac{1}{\epsilon(k+\frac{\pi}{2})-\epsilon(k)}\biggr{)}
+Pckck+π2Qck+π2ckP(1ϵ(k)ϵ(k+π2)+1ϵ(k)ϵ(k+π2))\displaystyle\qquad+Pc_{k}^{\dagger}c_{k+\frac{\pi}{2}}Qc_{k^{\prime}+\frac{\pi}{2}}^{\dagger}c_{k^{\prime}}P\biggl{(}\frac{1}{\epsilon(k^{\prime})-\epsilon(k^{\prime}+\frac{\pi}{2})}+\frac{1}{\epsilon(k)-\epsilon(k+\frac{\pi}{2})}\biggr{)}
+eπi/2Pckck+π2Qckck+π2P(1ϵ(k+π2)ϵ(k)+1ϵ(k)ϵ(k+π2))]\displaystyle\qquad+e^{\pi i/2}Pc_{k}^{\dagger}c_{k+\frac{\pi}{2}}Qc_{k^{\prime}}^{\dagger}c_{k^{\prime}+\frac{\pi}{2}}P\biggl{(}\frac{1}{\epsilon(k^{\prime}+\frac{\pi}{2})-\epsilon(k^{\prime})}+\frac{1}{\epsilon(k)-\epsilon(k+\frac{\pi}{2})}\biggr{)}\biggr{]} (98)

. One can simplify these projections. Since the Fermi surface is located at k=±π/2modπk=\pm\pi/2\mod\pi, the projection Pck+πckPPc_{k^{\prime}+\pi}^{\dagger}c_{k^{\prime}}P gives back ck+πckc_{k^{\prime}+\pi}^{\dagger}c_{k^{\prime}} itself for kRk^{\prime}\in R and zero otherwise. In the end, I obtain

V\displaystyle V^{\prime} =ih424kR(ck+πckckck+π)(1ϵ(k)ϵ(k+π2)+1ϵ(k+π)ϵ(k+π2))\displaystyle=-i\frac{h_{4}^{2}}{4}\sum_{k\in R}(c_{k+\pi}^{\dagger}c_{k}-c_{k}^{\dagger}c_{k+\pi})\biggl{(}\frac{1}{\epsilon(k)-\epsilon(k+\frac{\pi}{2})}+\frac{1}{\epsilon(k+\pi)-\epsilon(k+\frac{\pi}{2})}\biggr{)}
+h422kRckck(1ϵ(k)ϵ(kπ2)+1ϵ(k)ϵ(k+π2)).\displaystyle\qquad+\frac{h_{4}^{2}}{2}\sum_{k\in R}c_{k}^{\dagger}c_{k}\biggl{(}\frac{1}{\epsilon(k)-\epsilon(k-\frac{\pi}{2})}+\frac{1}{\epsilon(k)-\epsilon(k+\frac{\pi}{2})}\biggr{)}. (99)

The first line of Eq. (99) is the the bond alternation for k±kFk\approx\pm k_{F} Giamarchi (2004) and the second line is a small correction to the Zeeman energy.

Appendix B Electron spin resonance

Here, I describe how the uniform DM interaction affects the ESR spectrum. In this Appendix, I start with the spin chain model (60) with λ=1\lambda=1. Namely, I consider the spin chain with the following Hamiltonian:

~exp\displaystyle\tilde{\mathcal{H}}_{\rm exp} =Jj𝑺j𝑺j+1+h0jSjz\displaystyle=J\sum_{j}\bm{S}_{j}\cdot\bm{S}_{j+1}+h_{0}\sum_{j}S_{j}^{z}
+h2j(1)jSjx+2Jδ3j(1)j𝑺j𝑺j+1\displaystyle\qquad+h_{2}\sum_{j}(-1)^{j}S_{j}^{x}+\frac{2J\delta_{\perp}}{3}\sum_{j}(-1)^{j}\bm{S}_{j}\cdot\bm{S}_{j+1}
+Duj(SjzSj+1xSjxSj+1z).\displaystyle\qquad+D_{u}\sum_{j}(S_{j}^{z}S_{j+1}^{x}-S_{j}^{x}S_{j+1}^{z}). (100)

The ESR spectrum is obtained from the q=0q=0 part of the dynamical correlation function SaSb(q,ω)\braket{S^{a}S^{b}}(q,\omega) for a,b=x,y,za,b=x,y,z. I can obtain selection rules of the ESR spectrum by relating the q=0q=0 part of the spin, Sq=0aS^{a}_{q=0}, where

Sqa:=jeiqjSja,\displaystyle S^{a}_{q}:=\sum_{j}e^{iqj}S_{j}^{a}, (101)

to the boson fields of the effective field theory.

Let us bosonize the spin chain by using the non-Abelian bosonization formula Gogolin et al. (2004); Affleck and Haldane (1987),

𝑺j\displaystyle\bm{S}_{j} =𝑱R+𝑱Lib02(1)jtr(g𝝈),\displaystyle=\bm{J}_{R}+\bm{J}_{L}-\frac{ib_{0}}{2}(-1)^{j}\operatorname{tr}(g\bm{\sigma}), (102)

where 𝑱R\bm{J}_{R}, 𝑱L\bm{J}_{L}, and gg are defined as

JRz\displaystyle J_{R}^{z} =i2π¯φR,\displaystyle=-\frac{i}{\sqrt{2\pi}}\bar{\partial}\varphi_{R}, (103)
JLz\displaystyle J_{L}^{z} =i2πφL,\displaystyle=\frac{i}{\sqrt{2\pi}}\partial\varphi_{L}, (104)
JR±\displaystyle J_{R}^{\pm} =12πe±i8πφR,\displaystyle=\frac{1}{2\pi}e^{\pm i\sqrt{8\pi}\varphi_{R}}, (105)
JL±\displaystyle J_{L}^{\pm} =12πei8πφL,\displaystyle=\frac{1}{2\pi}e^{\mp i\sqrt{8\pi}\varphi_{L}}, (106)
g\displaystyle g =(ei2πϕiei2πθiei2πθei2πϕ).\displaystyle=\begin{pmatrix}e^{i\sqrt{2\pi}\phi}&ie^{-i\sqrt{2\pi}\theta}\\ ie^{i\sqrt{2\pi}\theta}&e^{-i\sqrt{2\pi}\phi}\end{pmatrix}. (107)

Here φ\varphi and φ\varphi at a position xx and a time tt are related to φR\varphi_{R} and φL\varphi_{L} through

ϕ(x,t)\displaystyle\phi(x,t) =φR(xvt)+φL(x+vt),\displaystyle=\varphi_{R}(x-vt)+\varphi_{L}(x+vt), (108)
θ(x,t)\displaystyle\theta(x,t) =φR(xvt)φL(x+vt).\displaystyle=\varphi_{R}(x-vt)-\varphi_{L}(x+vt). (109)

The derivatives \partial and ¯\bar{\partial} are abbreviations of the following derivatives.

\displaystyle\partial =i2(x+v1t),\displaystyle=\frac{-i}{2}(\partial_{x}+v^{-1}\partial_{t}), (110)
¯\displaystyle\bar{\partial} =i2(xv1t).\displaystyle=\frac{i}{2}(\partial_{x}-v^{-1}\partial_{t}). (111)

Boson fields ϕ\phi and θ\theta are subject to equal-time commutation relations,

[ϕ(x),θ(y)]=iY(yx),\displaystyle[\phi(x),\theta(y)]=iY(y-x), (112)

with a step function,

Y(yx)\displaystyle Y(y-x) ={1,(y>x),0,(y<x),1/2,(y=x).\displaystyle=\left\{\begin{array}[]{ccc}1,&&(y>x),\\ 0,&&(y<x),\\ 1/2,&&(y=x).\end{array}\right. (116)

SjaS_{j}^{a} for a=x,y,za=x,y,z are thus bosonized as Mudry et al. (2019)

Sjx\displaystyle S_{j}^{x} =b0cos(2πθ)+ib1sin(2πθ)sin(2πϕ),\displaystyle=b_{0}\cos(\sqrt{2\pi}\theta)+ib_{1}\sin(\sqrt{2\pi}\theta)\sin(\sqrt{2\pi}\phi), (117)
Sjy\displaystyle S_{j}^{y} =b0sin(2πθ)+ib1cos(2πθ)sin(2πϕ),\displaystyle=b_{0}\sin(\sqrt{2\pi}\theta)+ib_{1}\cos(\sqrt{2\pi}\theta)\sin(\sqrt{2\pi}\phi), (118)
Sjz\displaystyle S_{j}^{z} =12πxϕ+b0sin(2πϕ).\displaystyle=\frac{1}{\sqrt{2\pi}}\partial_{x}\phi+b_{0}\sin(\sqrt{2\pi}\phi). (119)

In the non-Abelian bosonization laugage, the Hamiltonian (100) is expressed as

~exp\displaystyle\tilde{\mathcal{H}}_{\rm exp} =2πv3𝑑x(𝑱R𝑱R+𝑱L𝑱L)\displaystyle=\frac{2\pi v}{3}\int dx\,(\bm{J}_{R}\cdot\bm{J}_{R}+\bm{J}_{L}\cdot\bm{J}_{L})
+h0𝑑x(JRz+JLz)+γDu𝑑x(JRyJLy)\displaystyle\qquad+h_{0}\int dx\,(J_{R}^{z}+J_{L}^{z})+\gamma D_{u}\int dx(J_{R}^{y}-J_{L}^{y})
ib0h22𝑑xtr(gσx)+dxyJδ3𝑑xtr(g).\displaystyle\qquad-\frac{ib_{0}h_{2}}{2}\int dx\,\operatorname{tr}(g\sigma^{x})+\frac{d_{xy}J\delta_{\perp}}{3}\int dx\,\operatorname{tr}(g). (120)

In order to combine terms on the second line, I perform the chiral rotation (63). The chiral rotation turns the Hamiltonian into

~exp\displaystyle\tilde{\mathcal{H}}_{\rm exp} =2πv3𝑑x(𝑴R𝑴R+𝑴L𝑴L)\displaystyle=\frac{2\pi v}{3}\int dx\,(\bm{M}_{R}\cdot\bm{M}_{R}+\bm{M}_{L}\cdot\bm{M}_{L})
+tϕ𝑑x(MRz+MLz)\displaystyle\qquad+t_{\phi}\int dx\,(M_{R}^{z}+M_{L}^{z})
ib0h22𝑑x[tr(gσx)cosθR+itr(g)sinθR]\displaystyle\qquad-\frac{ib_{0}h_{2}}{2}\int dx\,[\operatorname{tr}(g^{\prime}\sigma^{x})\cos\theta_{R}+i\operatorname{tr}(g^{\prime})\sin\theta_{R}]
+dxyJδ3𝑑x[tr(g)cosθR+itr(gσx)sinθR].\displaystyle\qquad+\frac{d_{xy}J\delta_{\perp}}{3}\int dx\,[\operatorname{tr}(g^{\prime})\cos\theta_{R}+i\operatorname{tr}(g^{\prime}\sigma^{x})\sin\theta_{R}]. (121)

𝑴R\bm{M}_{R}, 𝑴L\bm{M}_{L}, and gg^{\prime} are related to U(1) bosons Φ=φR+φL\Phi=\varphi^{\prime}_{R}+\varphi^{\prime}_{L} and Θ=φRφL\Theta=\varphi^{\prime}_{R}-\varphi^{\prime}_{L} in analogy with 𝑱R\bm{J}_{R}, 𝑱L\bm{J}_{L}, and gg. I can eliminate the Zeeman energy tϕ(MRz+MRz)t_{\phi}(M_{R}^{z}+M_{R}^{z}) by shifting

φR\displaystyle\varphi^{\prime}_{R} φRtϕv8πx,\displaystyle\to\varphi^{\prime}_{R}-\frac{t_{\phi}}{v\sqrt{8\pi}}x, (122)
φL\displaystyle\varphi^{\prime}_{L} φLtϕv8πx.\displaystyle\to\varphi^{\prime}_{L}-\frac{t_{\phi}}{v\sqrt{8\pi}}x. (123)

This shift affect 𝑴R\bm{M}_{R}, 𝑴L\bm{M}_{L}, and gg^{\prime} as follows.

MRz\displaystyle M_{R}^{z} =tϕ4πvi2π¯φR,\displaystyle=-\frac{t_{\phi}}{4\pi v}-\frac{i}{\sqrt{2\pi}}\bar{\partial}\varphi^{\prime}_{R}, (124)
MLz\displaystyle M_{L}^{z} =tϕ4πv+i2πφL,\displaystyle=-\frac{t_{\phi}}{4\pi v}+\frac{i}{\sqrt{2\pi}}\partial\varphi^{\prime}_{L}, (125)
MR±\displaystyle M_{R}^{\pm} =eitϕx/ve±i8πφR,\displaystyle=e^{\mp it_{\phi}x/v}e^{\pm i\sqrt{8\pi}\varphi^{\prime}_{R}}, (126)
ML±\displaystyle M_{L}^{\pm} =e±itϕx/vei8πφL,\displaystyle=e^{\pm it_{\phi}x/v}e^{\mp i\sqrt{8\pi}\varphi^{\prime}_{L}}, (127)
g\displaystyle g^{\prime} =(eitϕx/vei2πΦiei2πΘiei2πΘeitϕx/vei2πΦ).\displaystyle=\begin{pmatrix}e^{-it_{\phi}x/v}e^{i\sqrt{2\pi}\Phi}&ie^{-i\sqrt{2\pi}\Theta}\\ ie^{i\sqrt{2\pi}\Theta}&e^{it_{\phi}x/v}e^{-i\sqrt{2\pi}\Phi}\end{pmatrix}. (128)

The shift introduces incommensurate oscillations to the Hamiltonian (121). Here, I assume an inequality,

vΔvtϕ\displaystyle\frac{v}{\Delta}\ll\frac{v}{t_{\phi}} (129)

This condition guarantees that the incommensurate oscillation is negligible (cf. Secs. V.1 and V.2). At low fields h0/J1h_{0}/J\ll 1, this inequality reads

(h0/J)2/3tϕ/J.\displaystyle(h_{0}/J)^{2/3}\gg t_{\phi}/J. (130)

Under this condition (130), I can safely discard the incommensurate oscillation with the wave number tϕ/vt_{\phi}/v in the Hamiltonian. Note that tϕt_{\phi} must be kept in the relations between operators and quantum fields. The Hamiltonian is thus given by

~exp\displaystyle\tilde{\mathcal{H}}_{\rm exp} 2πv3𝑑x(𝑴R𝑴R+𝑴L𝑴L)\displaystyle\approx\frac{2\pi v}{3}\int dx\,(\bm{M}_{R}\cdot\bm{M}_{R}+\bm{M}_{L}\cdot\bm{M}_{L})
ib0h22𝑑x[tr(gσx)cosθR+itr(g)sinθR]\displaystyle\qquad-\frac{ib_{0}h_{2}}{2}\int dx\,[\operatorname{tr}(g^{\prime}\sigma^{x})\cos\theta_{R}+i\operatorname{tr}(g^{\prime})\sin\theta_{R}]
+dxyJδ3𝑑x[tr(g)cosθR+itr(gσx)sinθR].\displaystyle\qquad+\frac{d_{xy}J\delta_{\perp}}{3}\int dx\,[\operatorname{tr}(g^{\prime})\cos\theta_{R}+i\operatorname{tr}(g^{\prime}\sigma^{x})\sin\theta_{R}]. (131)

Here, as I did in Sec. V.2, I rotate the system by π\pi around the yy axis: (σx,σy,σz)(σz,σy,σx)(\sigma^{x},\,\sigma^{y},\,\sigma^{z})\to(\sigma^{z},\,\sigma^{y},\,-\sigma^{x}). The π\pi-rotated Hamiltonian finally becomes simple.

~exp\displaystyle\tilde{\mathcal{H}}_{\rm exp} =v2𝑑x[v2(tϕ)2+(xϕ)2]\displaystyle=\frac{v}{2}\int dx\,[v^{-2}(\partial_{t}\phi)^{2}+(\partial_{x}\phi)^{2}]
+G𝑑xcos(2πΦ+θRα),\displaystyle\qquad+G\int dx\,\cos(\sqrt{2\pi}\Phi+\theta_{R}-\alpha^{\prime}), (132)

with

G\displaystyle G =(b0h2)2+(2dxyJδ/3)2,\displaystyle=\sqrt{(b_{0}h_{2})^{2}+(2d_{xy}J\delta_{\perp}/3)^{2}}, (133)
α\displaystyle\alpha^{\prime} =tan1(3b0h22dxyJδ).\displaystyle=\tan^{-1}\biggl{(}\frac{3b_{0}h_{2}}{2d_{xy}J\delta_{\perp}}\biggr{)}. (134)

Let us relate the spin 𝑺j\bm{S}_{j} in the original coordinate frame to the Φ\Phi and Θ\Theta fields in Eq. (132), recalling all the chiral and nonchiral rotations performed.

Sjx\displaystyle S_{j}^{x} =MRx+MLxib02[tr(gσx)cosθR+itr(g)sinθR]\displaystyle=M_{R}^{x}+M_{L}^{x}-\frac{ib_{0}}{2}[\operatorname{tr}(g^{\prime}\sigma^{x})\cos\theta_{R}+i\operatorname{tr}(g^{\prime})\sin\theta_{R}]
=12π(xΦ)cos(tϕx/v)+ib1cos(2πΘ)sin(2πΦ)sin(tϕx/v)+b0(1)jsin(2πΦ+θR),\displaystyle=\frac{1}{\sqrt{2\pi}}(\partial_{x}\Phi)\cos(t_{\phi}x/v)+ib_{1}\cos(\sqrt{2\pi}\Theta)\sin(\sqrt{2\pi}\Phi)\sin(t_{\phi}x/v)+b_{0}(-1)^{j}\sin(\sqrt{2\pi}\Phi+\theta_{R}), (135)
Sjy\displaystyle S_{j}^{y} =(MRy+MLy)cosθR(MRzMLz)sinθRib02(1)jtr(gσy)\displaystyle=(M_{R}^{y}+M_{L}^{y})\cos\theta_{R}-(M_{R}^{z}-M_{L}^{z})\sin\theta_{R}-\frac{ib_{0}}{2}(-1)^{j}\operatorname{tr}(g^{\prime}\sigma^{y})
=(ib1cos(2πΘ)sin(2πΦ)cos(tϕx/v)12π(xΘ)sin(tϕx/v))cosθR\displaystyle=\biggl{(}ib_{1}\cos(\sqrt{2\pi}\Theta)\sin(\sqrt{2\pi}\Phi)\cos(t_{\phi}x/v)-\frac{1}{\sqrt{2\pi}}(\partial_{x}\Theta)\sin(t_{\phi}x/v)\biggr{)}\cos\theta_{R}
ib1cos(2πΘ)cos(2πΦ)sinθR+b0(1)jsin(2πΘ),\displaystyle\qquad-ib_{1}\cos(\sqrt{2\pi}\Theta)\cos(\sqrt{2\pi}\Phi)\sin\theta_{R}+b_{0}(-1)^{j}\sin(\sqrt{2\pi}\Theta), (136)
Sjz\displaystyle S_{j}^{z} =(MRyMLy)sinθR+(MRz+MLz)cosθRib02(1)jtr(gσz)\displaystyle=(M_{R}^{y}-M_{L}^{y})\sin\theta_{R}+(M_{R}^{z}+M_{L}^{z})\cos\theta_{R}-\frac{ib_{0}}{2}(-1)^{j}\operatorname{tr}(g^{\prime}\sigma^{z})
=tϕ2πvcosθR+(ib1sin(2πΘ)cos(2πΦ)cos(tϕx/v)+12π(xΦ)cos(tϕx/v))sinθR\displaystyle=-\frac{t_{\phi}}{2\pi v}\cos\theta_{R}+\biggl{(}ib_{1}\sin(\sqrt{2\pi}\Theta)\cos(\sqrt{2\pi}\Phi)\cos(t_{\phi}x/v)+\frac{1}{\sqrt{2\pi}}(\partial_{x}\Phi)\cos(t_{\phi}x/v)\biggr{)}\sin\theta_{R}
ib1sin(2πΘ)sin(2πΦ)+b0(1)jsin(2πΦtϕx/v).\displaystyle\qquad-ib_{1}\sin(\sqrt{2\pi}\Theta)\sin(\sqrt{2\pi}\Phi)+b_{0}(-1)^{j}\sin(\sqrt{2\pi}\Phi-t_{\phi}x/v). (137)

Note that the π\pi rotation was performed on the rightmost hand sides of Eqs. (135), (136), and (137). The transverse dimer order is expressed as

(1)j(SjxSj+1x+SjySj+1z)\displaystyle(-1)^{j}(S_{j}^{x}S_{j+1}^{x}+S_{j}^{y}S_{j+1}^{z}) =dxy2(tr(g)cosθR+itr(gσx)sinθR)\displaystyle=\frac{d_{xy}}{2}(\operatorname{tr}(g^{\prime})\cos\theta_{R}+i\operatorname{tr}(g\sigma^{x})\sin\theta_{R})
=dxycos(2πΦ+θR).\displaystyle=d_{xy}\cos(\sqrt{2\pi}\Phi+\theta_{R}). (138)

I can confirm that the spin chain model (100) has the Néel and dimer orders.

Nx\displaystyle N_{x} (G/J)2/3sinα,\displaystyle\propto(G/J)^{2/3}\sin\alpha^{\prime}, (139)
D\displaystyle D_{\perp} (G/J)2/3cosα,\displaystyle\propto(G/J)^{2/3}\cos\alpha^{\prime}, (140)

in analogy with Eqs. (52) and (53).

A list of low-energy excitations created by operators on the rightmost hand sides of Eqs (135), (136), and (137) is available Lukyanov (1997); Lukyanov and Zamolodchikov (2001); Babujian and Karowski (2002). Vertex operator operators eiqxe±i2πΘe^{iqx}e^{\pm i\sqrt{2\pi}\Theta} create solitons and antisolitons with an excitation energy,

Es(q)\displaystyle E_{s}(q) =(vq)2+Δs2,\displaystyle=\sqrt{(vq)^{2}+{\Delta_{s}}^{2}}, (141)
Δs\displaystyle\Delta_{s} =2vπΓ(1/6)Γ(2/3)(π2vΓ(3/4)Γ(1/4)G)2/3.\displaystyle=\frac{2v}{\sqrt{\pi}}\frac{\Gamma(1/6)}{\Gamma(2/3)}\biggl{(}\frac{\pi}{2v}\frac{\Gamma(3/4)}{\Gamma(1/4)}G\biggr{)}^{2/3}. (142)

Other vertex operators eiqxe±i2πΦe^{iqx}e^{\pm i\sqrt{2\pi}\Phi} create breathers, bound states of a soliton and an antisoliton, with an excitation energy,

En(q)\displaystyle E_{n}(q) =(vq)2+Δn2,\displaystyle=\sqrt{(vq)^{2}+\Delta_{n}^{2}}, (143)
Δn\displaystyle\Delta_{n} =2Δssin(nπξ2).\displaystyle=2\Delta_{s}\sin\biggl{(}\frac{n\pi\xi}{2}\biggr{)}. (144)

Here, ξ=1/(8K1)=3\xi=1/(8K-1)=3. The index nn takes values of n=1,2,3n=1,2,3.

I can now predict the ESR frequency caused by Sq=0aS_{q=0}^{a} for a=x,y,za=x,y,z. For example, cos(2πΘ)cos(2πΦ)\cos(\sqrt{2\pi}\Theta)\cos(\sqrt{2\pi}\Phi) in Sq=0yS_{q=0}^{y} yields the resonance peak at

ω=ΔsG2/3.\displaystyle\omega=\Delta_{s}\propto G^{2/3}. (145)

Though in the h0/J0h_{0}/J\to 0 limit, this resonance frequency follows a simple power law ω(h0/J)2/3\omega\propto(h_{0}/J)^{2/3}, it will be a complicated function of h0h_{0} in general. Another interesting term is sin(2πΘ)cos(2πΦ)cos(tϕx/v)\sin(\sqrt{2\pi}\Theta)\cos(\sqrt{2\pi}\Phi)\cos(t_{\phi}x/v) in Sq=0zS_{q=0}^{z}. This term yields resonance peaks at

ω\displaystyle\omega =tϕ2+M2,\displaystyle=\sqrt{{t_{\phi}}^{2}+M^{2}}, (146)

where MM can be Δs\Delta_{s} or Δn\Delta_{n} for n=1,2,3n=1,2,3.

The selection rule is also affected by the canonical transformation (80). Precisely speaking, the left hand sides of Eqs. (135), (136), and (137) should be denoted as S~ja\tilde{S}_{j}^{a} for a=x,y,za=x,y,z. Here, S~ja\tilde{S}_{j}^{a} is defined as

S~ja\displaystyle\tilde{S}_{j}^{a} =eηSjaeη\displaystyle=e^{\eta}S_{j}^{a}e^{-\eta}
=Sja+[η1,Sja]+[η2,Sja]+12[η1,[η1,Sja]]+.\displaystyle=S_{j}^{a}+[\eta_{1},S_{j}^{a}]+[\eta_{2},S_{j}^{a}]+\frac{1}{2}[\eta_{1},[\eta_{1},S_{j}^{a}]]+\cdots. (147)

η1\eta_{1} and η2\eta_{2} create particle-hole excitations with q=±π/2q=\pm\pi/2 and π\pi, respectively, when applied to the TL-liquid ground state. Such a mixing of different wave numbers will allow ESR to detect q=±π/2q=\pm\pi/2 and q=πq=\pi excitations. Excitations with q=πq=\pi can be read from the staggered terms of Eqs. (135), (136), and (137), which are similar to those with q=0q=0.

References

  • Lieb et al. (1961) Elliott Lieb, Theodore Schultz,  and Daniel Mattis, “Two soluble models of an antiferromagnetic chain,” Annals of Physics 16, 407 – 466 (1961).
  • Affleck and Haldane (1987) Ian Affleck and F. D. M. Haldane, “Critical theory of quantum spin chains,” Phys. Rev. B 36, 5291–5300 (1987).
  • Furuya and Oshikawa (2017) Shunsuke C. Furuya and Masaki Oshikawa, “Symmetry Protection of Critical Phases and a Global Anomaly in 1+11+1 Dimensions,” Phys. Rev. Lett. 118, 021601 (2017).
  • Giamarchi (2004) T. Giamarchi, Quantum Physics in One Dimension (Oxford University Press, Oxford, 2004).
  • Metlitski and Thorngren (2018) Max A. Metlitski and Ryan Thorngren, “Intrinsic and emergent anomalies at deconfined critical points,” Phys. Rev. B 98, 085140 (2018).
  • Dender et al. (1997) D. C. Dender, P. R. Hammar, Daniel H. Reich, C. Broholm,  and G. Aeppli, “Direct Observation of Field-Induced Incommensurate Fluctuations in a One-Dimensional S=1/2S=1/2 Antiferromagnet,” Phys. Rev. Lett. 79, 1750–1753 (1997).
  • Zvyagin et al. (2004) S. A. Zvyagin, A. K. Kolezhuk, J. Krzystek,  and R. Feyerherm, “Excitation Hierarchy of the Quantum Sine-Gordon Spin Chain in a Strong Magnetic Field,” Phys. Rev. Lett. 93, 027201 (2004).
  • Umegaki et al. (2009) Izumi Umegaki, Hidekazu Tanaka, Toshio Ono, Hidehiro Uekusa,  and Hiroyuki Nojiri, “Elementary excitations of the S=12S=\frac{1}{2} one-dimensional antiferromagnet KCuGaF6{\text{KCuGaF}}_{6} in a magnetic field and quantum sine-Gordon model,” Phys. Rev. B 79, 184401 (2009).
  • Oshikawa and Affleck (1997) Masaki Oshikawa and Ian Affleck, “Field-Induced Gap in S=1/2\mathit{S}=1/2 Antiferromagnetic Chains,” Phys. Rev. Lett. 79, 2883–2886 (1997).
  • Affleck and Oshikawa (1999) Ian Affleck and Masaki Oshikawa, “Field-induced gap in Cu benzoate and other S=12S=\frac{1}{2} antiferromagnetic chains,” Phys. Rev. B 60, 1038–1056 (1999).
  • Furuya and Oshikawa (2012) Shunsuke C. Furuya and Masaki Oshikawa, “Boundary Resonances in S=1/2S\mathbf{=}1/2 Antiferromagnetic Chains Under a Staggered Field,” Phys. Rev. Lett. 109, 247603 (2012).
  • Furuya et al. (2011) Shunsuke C. Furuya, Masaki Oshikawa,  and Ian Affleck, “Semiclassical approach to electron spin resonance in quantum spin systems,” Phys. Rev. B 83, 224417 (2011).
  • Faure et al. (2018) Quentin Faure, Shintaro Takayoshi, Sylvain Petit, Virginie Simonet, Stéphane Raymond, Louis-Pierre Regnault, Martin Boehm, Jonathan S White, Martin Månsson, Christian Rüegg, et al., “Topological quantum phase transition in the Ising-like antiferromagnetic spin chain BaCo2V2O8,” Nature Physics 14, 716 (2018).
  • Takayoshi et al. (2018) Shintaro Takayoshi, Shunsuke C. Furuya,  and Thierry Giamarchi, “Topological transition between competing orders in quantum spin chains,” Phys. Rev. B 98, 184429 (2018).
  • Kimura et al. (2013) Shojiro Kimura, Koichi Okunishi, Masayuki Hagiwara, Koichi Kindo, Zhangzhen He, Tomoyasu Taniyama, Mitsuru Itoh, Keiichi Koyama,  and Kazuo Watanabe, “Collapse of Magnetic Order of the Quasi One-Dimensional Ising-Like Antiferromagnet BaCo2V2O8 in Transverse Fields,” Journal of the Physical Society of Japan 82, 033706 (2013).
  • Liu et al. (2019) J. Liu, S. Kittaka, R. D. Johnson, T. Lancaster, J. Singleton, T. Sakakibara, Y. Kohama, J. van Tol, A. Ardavan, B. H. Williams, S. J. Blundell, Z. E. Manson, J. L. Manson,  and P. A. Goddard, “Unconventional Field-Induced Spin Gap in an S=1/2S=1/2 Chiral Staggered Chain,” Phys. Rev. Lett. 122, 057207 (2019).
  • (17) ITensor Library (version 2.0.11) http://itensor.org .
  • Schrieffer and Wolff (1966) J. R. Schrieffer and P. A. Wolff, “Relation between the Anderson and Kondo Hamiltonians,” Phys. Rev. 149, 491–492 (1966).
  • Slagle and Kim (2017) Kevin Slagle and Yong Baek Kim, “Fracton topological order from nearest-neighbor two-spin interactions and dualities,” Phys. Rev. B 96, 165106 (2017).
  • Shelton et al. (1996) D. G. Shelton, A. A. Nersesyan,  and A. M. Tsvelik, “Antiferromagnetic spin ladders: Crossover between spin S=1/2 and S=1 chains,” Phys. Rev. B 53, 8521–8532 (1996).
  • Hikihara and Furusaki (2004) T. Hikihara and A. Furusaki, “Correlation amplitudes for the spin-12\frac{1}{2} XXZ\mathrm{XXZ} chain in a magnetic field,” Phys. Rev. B 69, 064427 (2004).
  • Takayoshi and Sato (2010) Shintaro Takayoshi and Masahiro Sato, “Coefficients of bosonized dimer operators in spin-12\frac{1}{2} XXZXXZ chains and their applications,” Phys. Rev. B 82, 214420 (2010).
  • Hikihara et al. (2017) Toshiya Hikihara, Akira Furusaki,  and Sergei Lukyanov, “Dimer correlation amplitudes and dimer excitation gap in spin-12\frac{1}{2} XXZ and Heisenberg chains,” Phys. Rev. B 96, 134429 (2017).
  • Essler and Konik (2005) Fabian H. L. Essler and Robert M. Konik, “Application of massive integrable quantum field theories to problems in condensed matter physics,” in From Fields to Strings: Circumnavigating Theoretical Physics: Ian Kogan Memorial Collection (In 3 Volumes) (World Scientific, 2005) pp. 684–830.
  • Lukyanov and Zamolodchikov (1997) Sergei Lukyanov and Alexander Zamolodchikov, “Exact expectation values of local fields in the quantum sine-gordon model,” Nuclear Physics B 493, 571 – 587 (1997).
  • Zamolodchikov (1995) Al. B. Zamolodchikov, “Mass scale in the sine–gordon model and its reductions,” International Journal of Modern Physics A 10, 1125–1150 (1995).
  • Eßler (1999) Fabian H. L. Eßler, “Sine-gordon low-energy effective theory for copper benzoate,” Phys. Rev. B 59, 14376–14383 (1999).
  • Cardy (1996) John Cardy, Scaling and renormalization in statistical physics, Vol. 5 (Cambridge university press, 1996).
  • Starykh et al. (2005) Oleg A. Starykh, Akira Furusaki,  and Leon Balents, “Anisotropic pyrochlores and the global phase diagram of the checkerboard antiferromagnet,” Phys. Rev. B 72, 094416 (2005).
  • Garate and Affleck (2010) Ion Garate and Ian Affleck, “Interplay between symmetric exchange anisotropy, uniform Dzyaloshinskii-Moriya interaction, and magnetic fields in the phase diagram of quantum magnets and superconductors,” Phys. Rev. B 81, 144419 (2010).
  • Chan et al. (2017) Yang-Hao Chan, Wen Jin, Hong-Chen Jiang,  and Oleg A. Starykh, “Ising order in a magnetized Heisenberg chain subject to a uniform Dzyaloshinskii-Moriya interaction,” Phys. Rev. B 96, 214441 (2017).
  • Jin and Starykh (2017) Wen Jin and Oleg A. Starykh, “Phase diagram of weakly coupled Heisenberg spin chains subject to a uniform Dzyaloshinskii-Moriya interaction,” Phys. Rev. B 95, 214404 (2017).
  • Gogolin et al. (2004) Alexander O Gogolin, Alexander A Nersesyan,  and Alexei M Tsvelik, Bosonization and strongly correlated systems (Cambridge university press, 2004).
  • Gangadharaiah et al. (2008) Suhas Gangadharaiah, Jianmin Sun,  and Oleg A. Starykh, “Spin-orbital effects in magnetized quantum wires and spin chains,” Phys. Rev. B 78, 054436 (2008).
  • Povarov et al. (2011) K. Yu. Povarov, A. I. Smirnov, O. A. Starykh, S. V. Petrov,  and A. Ya. Shapiro, “Modes of magnetic resonance in the spin-liquid phase of cs2cucl4{\mathrm{cs}}_{2}{\mathrm{cucl}}_{4},” Phys. Rev. Lett. 107, 037204 (2011).
  • Furuya (2017) Shunsuke C. Furuya, “Angular dependence of electron spin resonance for detecting the quadrupolar liquid state of frustrated spin chains,” Phys. Rev. B 95, 014416 (2017).
  • Oshikawa and Affleck (2002) Masaki Oshikawa and Ian Affleck, “Electron spin resonance in S=12S=\frac{1}{2} antiferromagnetic chains,” Phys. Rev. B 65, 134410 (2002).
  • Ozerov et al. (2015) M. Ozerov, M. Maksymenko, J. Wosnitza, A. Honecker, C. P. Landee, M. M. Turnbull, S. C. Furuya, T. Giamarchi,  and S. A. Zvyagin, “Electron spin resonance modes in a strong-leg ladder in the tomonaga-luttinger liquid phase,” Phys. Rev. B 92, 241113 (2015).
  • Zhao et al. (2003) J. Z. Zhao, X. Q. Wang, T. Xiang, Z. B. Su,  and L. Yu, “Effects of the Dzyaloshinskii-Moriya Interaction on Low-Energy Magnetic Excitations in Copper Benzoate,” Phys. Rev. Lett. 90, 207204 (2003).
  • Lou et al. (2005) Jizhong Lou, Changfeng Chen, Jize Zhao, Xiaoqun Wang, Tao Xiang, Zhaobin Su,  and Lu Yu, “Midgap States in Antiferromagnetic Heisenberg Chains with a Staggered Field,” Phys. Rev. Lett. 94, 217207 (2005).
  • Kimura et al. (2007) S. Kimura, H. Yashiro, K. Okunishi, M. Hagiwara, Z. He, K. Kindo, T. Taniyama,  and M. Itoh, “Field-Induced Order-Disorder Transition in Antiferromagnetic BaCo2V2O8{\mathrm{BaCo}}_{2}{\mathrm{V}}_{2}{\mathrm{O}}_{8} Driven by a Softening of Spinon Excitation,” Phys. Rev. Lett. 99, 087602 (2007).
  • Wang et al. (2015) Zhe Wang, M. Schmidt, A. K. Bera, A. T. M. N. Islam, B. Lake, A. Loidl,  and J. Deisenhofer, “Spinon confinement in the one-dimensional Ising-like antiferromagnet SrCo2V2O8{\mathrm{SrCo}}_{2}{\mathrm{V}}_{2}{\mathrm{O}}_{8},” Phys. Rev. B 91, 140404 (2015).
  • Bera et al. (2017) A. K. Bera, B. Lake, F. H. L. Essler, L. Vanderstraeten, C. Hubig, U. Schollwöck, A. T. M. N. Islam, A. Schneidewind,  and D. L. Quintero-Castro, “Spinon confinement in a quasi-one-dimensional anisotropic heisenberg magnet,” Phys. Rev. B 96, 054423 (2017).
  • Bravyi et al. (2011) Sergey Bravyi, David P. DiVincenzo,  and Daniel Loss, “Schrieffer–wolff transformation for quantum many-body systems,” Annals of Physics 326, 2793 – 2826 (2011).
  • Mudry et al. (2019) Christopher Mudry, Akira Furusaki, Takahiro Morimoto,  and Toshiya Hikihara, “Quantum phase transitions beyond landau-ginzburg theory in one-dimensional space revisited,” Phys. Rev. B 99, 205153 (2019).
  • Lukyanov (1997) Sergei Lukyanov, “Form Factors of Exponential Fields in the Sine–Gordon Model,” Modern Physics Letters A 12, 2543–2550 (1997).
  • Lukyanov and Zamolodchikov (2001) Sergei Lukyanov and Alexander Zamolodchikov, “Form factors of soliton-creating operators in the sine-Gordon model,” Nuclear Physics B 607, 437 – 455 (2001).
  • Babujian and Karowski (2002) H Babujian and M Karowski, “Sine-Gordon breather form factors and quantum field equations,” Journal of Physics A: Mathematical and General 35, 9081–9104 (2002).