Skip to main content

Part of the book series: Lecture Notes in Physics ((LNP,volume 862))

Abstract

The one-dimensional pure Ising model in a transverse field is the simplest solvable model that shows a quantum phase transition. The properties of this model are investigated in detail in Chap. 2. First, the quantum critical point is identified by employing duality argument. The Hamiltonian of the transverse Ising chain is then diagonalised and its ground-state properties are investigated by using the Jordan-Wigner transformation, which maps the system to a free fermion system. Next, approximate methods such as the exact diagonalisation method with finite size scaling, which can be also applied to interacting fermion systems, and the real-space renormalisation group method are introduced. The finite temperature property and an experimental study of the transverse Ising chain are mentioned in the last part of this chapter. Details of the Jordan-Wiger transformation, diagonalisation of a general Hamiltonian quadratic in fermion operators, and the calculation of correlation functions of the transverse Ising chain are included in appendices.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 79.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 99.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Amico, L., Fazio, R., Osterloh, A., Vedral, V.: Entanglement in many-body systems. Rev. Mod. Phys. 80, 517–576 (2008). [2.2.1]

    Article  MathSciNet  ADS  MATH  Google Scholar 

  2. Calabrese, P., Cardy, J.: Entanglement entropy and quantum field theory. J. Stat. Mech. Theory Exp. 2004(06), P06002 (2004). [2.2.1]

    Article  MathSciNet  Google Scholar 

  3. Christe, P., Henkel, M.: Introduction to Conformal Invariance and Applications to Critical Phenomena. Lecture Notes in Physics Monographs, vol. M 16, pp. 122–136. Springer, Heidelberg (1993). Chapter 10. [2.A.2]

    MATH  Google Scholar 

  4. Coldea, R., Tennant, D.A., Wheeler, E.M., Wawrzynska, E., Prabhakaran, D., Telling, M., Habicht, K., Smeibidl, P., Kiefer, K.: Quantum criticality in an Ising chain: experimental evidence for emergent E8 symmetry. Science 327(5962), 177–180 (2010). [1.3, 2.6]

    Article  ADS  Google Scholar 

  5. Drell, S.D., Weinstein, M., Yankielowicz, S.: Quantum field theories on a lattice: variational methods for arbitrary coupling strengths and the Ising model in a transverse magnetic field. Phys. Rev. D 16, 1769–1781 (1977). [2.4]

    Article  ADS  Google Scholar 

  6. Drzewiński, A., Dekeyser, R.: Renormalization of the anisotropic linear XY model. Phys. Rev. B 51, 15218–15228 (1995). [2.4.1]

    Article  ADS  Google Scholar 

  7. Fisher, M.E., Barber, M.N.: Scaling theory for finite-size effects in the critical region. Phys. Rev. Lett. 28, 1516–1519 (1972). See also Barber, M.N.: In: Domb, C., Lebowitz, J.L. (eds.) Phase Transition and Critical Phenomena, vol. 8, p. 146. Academic Press, San Diego (1983). [2.3.1]

    Article  ADS  Google Scholar 

  8. Fisher, M.E., Hartwig, R.E.: Toeplitz determinants: some applications, theorems, and conjectures. Adv. Chem. Phys. 15, 333–353 (1968). [2.A.3]

    Article  Google Scholar 

  9. Hamer, C.J., Barber, M.N.: Finite-lattice methods in quantum Hamiltonian field theory. i. The Ising model. J. Phys. A, Math. Gen. 14(1), 241 (1981). [1.3, 2.3.1, 4.3]

    Article  MathSciNet  ADS  Google Scholar 

  10. Hirsch, J.E., Mazenko, G.F.: Renormalization-group transformation for quantum lattice systems at zero temperature. Phys. Rev. B 19, 2656–2663 (1979). [2.4.1]

    Article  ADS  Google Scholar 

  11. Holzhey, C., Larsen, F., Wilczek, F.: Geometric and renormalized entropy in conformal field theory. Nucl. Phys. B 424(3), 443–467 (1994). [2.2.1]

    Article  MathSciNet  ADS  MATH  Google Scholar 

  12. Hu, B.: The classical Ising model: a quantum renormalization group approach. Phys. Lett. A 71(1), 83–86 (1979). [2.4.1, 4.3]

    Article  ADS  Google Scholar 

  13. Hu, B.: Introduction to real-space renormalization-group methods in critical and chaotic phenomena. Phys. Rep. 91(5), 233–295 (1982). [2.4.1]

    Article  MathSciNet  ADS  Google Scholar 

  14. Its, A.R., Jin, B.Q., Korepin, V.E.: Entanglement in the XY spin chain. J. Phys. A, Math. Gen. 38(13), 2975 (2005). [2.2.1]

    Article  MathSciNet  ADS  MATH  Google Scholar 

  15. Jullien, R., Fields, J.N., Doniach, S.: Zero-temperature real-space renormalization-group method for a Kondo-lattice model Hamiltonian. Phys. Rev. B 16, 4889–4900 (1977). [2.4.1]

    Article  ADS  Google Scholar 

  16. Jullien, R., Pfeuty, P., Fields, J.N., Doniach, S.: Zero-temperature renormalization method for quantum systems. i. Ising model in a transverse field in one dimension. Phys. Rev. B 18, 3568–3578 (1978). [2.4, 3.5]

    Article  ADS  Google Scholar 

  17. Jullien, R., Pfeuty, P., Fields, J.N., Penson, K.A.: In: Brukhardt, T.W., van Leeuween, J.M.J. (eds.) Real Space Renormalisation, p. 119. Springer, Berlin (1982). [2.4]

    Google Scholar 

  18. Katsura, S.: Statistical mechanics of the anisotropic linear Heisenberg model. Phys. Rev. 127, 1508–1518 (1962). [1.1, 1.3, 2.1.21, 10.1.2]

    Article  ADS  MATH  Google Scholar 

  19. Kogut, J.B.: An introduction to lattice gauge theory and spin systems. Rev. Mod. Phys. 51, 659–713 (1979). [1.1, 2.1.1, 3.A.2]

    Article  MathSciNet  ADS  Google Scholar 

  20. Kramers, H.A., Wannier, G.H.: Statistics of the two-dimensional ferromagnet. Part i. Phys. Rev. 60, 252–262 (1941). [2.1.1, 10.1.2]

    Article  MathSciNet  ADS  Google Scholar 

  21. Lieb, E., Schultz, T., Mattis, D.: Two soluble models of an antiferromagnetic chain. Ann. Phys. 16(3), 407–466 (1961). [2.2, 2.A.2, 10.1.2]

    Article  MathSciNet  ADS  MATH  Google Scholar 

  22. McCoy, B.M.: Spin correlation functions of the XY model. Phys. Rev. 173, 531–541 (1968). [2.2.1, 2.A.3]

    Article  MathSciNet  ADS  Google Scholar 

  23. McCoy, B.M.: Advanced Statistical Mechanics. Oxford University Press, Oxford (2010). [2.A.3]

    MATH  Google Scholar 

  24. Mussardo, G.: Statistical Field Theory. Oxford University Press, Oxford (2010). [2.6]

    MATH  Google Scholar 

  25. Nielsen, M.A., Chuang, I.L.: Quantum Computation and Quantum Information. Cambridge University Press, Cambridge (2000). [2.2.1, 8.1, 8.8]

    MATH  Google Scholar 

  26. Osterloh, A., Amico, L., Falci, G., Fazio, R.: Scaling of entanglement close to a quantum phase transition. Nature 416, 608–610 (2002). [2.2.1]

    Article  ADS  Google Scholar 

  27. Peschel, I.: On the entanglement entropy for an XY spin chain. J. Stat. Mech. Theory Exp. 2004(12), P12005 (2004). [2.2.1]

    Article  MathSciNet  Google Scholar 

  28. Pfeuty, P.: The one-dimensional Ising model with a transverse field. Ann. Phys. 57(1), 79–90 (1970). [1.1, 1.3, 2.2, 2.2.1, 2.A.3, 4.3, 5.2, 10.1.2]

    Article  ADS  Google Scholar 

  29. Sachdev, S.: Universal, finite-temperature, crossover functions of the quantum transition in the Ising chain in a transverse field. Nucl. Phys. B 464(3), 576–595 (1996). [2.A.3]

    Article  MathSciNet  ADS  MATH  Google Scholar 

  30. Sachdev, S.: Quantum Phase Transitions. Cambridge University Press, Cambridge (1999). [1.1, 2.A.3]

    Google Scholar 

  31. Shankar, R.: In: Pati, J., Shafi, Q., Lu, Yu (eds.) Current Topics in Condensed Matter and Particle Physics. World Scientific, Singapore (1993). [2.2.1]

    Google Scholar 

  32. Stella, A.L., Vanderzande, C., Dekeyser, R.: Unified renormalization-group approach to the thermodynamic and ground-state properties of quantum lattice systems. Phys. Rev. B 27, 1812–1831 (1983). [2.4.1]

    Article  MathSciNet  ADS  Google Scholar 

  33. Syljuåsen, O.F.: Entanglement and spontaneous symmetry breaking in quantum spin models. Phys. Rev. A 68, 060301 (2003). [2.2.1]

    Article  MathSciNet  ADS  Google Scholar 

  34. Szegö, G.: On certain Hermitian forms associated with the Fourier series of a positive function. Comm. Sém. Math. Univ. Lund (Medd. Lunds Univ. Mat. Sem.) 1952(Tome Supplementaire), 228–238 (1952). [2.A.3]

    Google Scholar 

  35. Vidal, G., Latorre, J.I., Rico, E., Kitaev, A.: Entanglement in quantum critical phenomena. Phys. Rev. Lett. 90, 227902 (2003). [2.2.1]

    Article  ADS  Google Scholar 

  36. Wootters, W.K.: Entanglement of formation of an arbitrary state of two qubits. Phys. Rev. Lett. 80, 2245–2248 (1998). [2.2.1]

    Article  ADS  Google Scholar 

  37. Zamolodchikov, A.B.: Integrals of motion and s-matrix of the (scaled) t=t c Ising model with magnetic field. Int. J. Mod. Phys. A 4, 4235 (1989). [2.6]

    Article  MathSciNet  ADS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Appendix 2.A

Appendix 2.A

2.1.1 2.A.1 Jordan-Wigner Fermions

To check that the operators c i and \(c_{i}^{\dag}\) satisfy fermionic anti-commutation relations

$$ \bigl[c_{i}, c_{i}^{\dag}\bigr]_{+} = \delta_{ik}, \qquad [c_{i},c_{k}]_{+} = 0 \quad \mbox{and} \quad \bigl[c_{i}^{\dag},c_{k}^{\dag} \bigr]_{+} = 0, $$
(2.A.1)

one uses the simple relations obeyed by Pauli spin operators

$$ S_{i}^{-}S_{i}^{+} = \frac{1}{2}\bigl(1-S_{i}^{z}\bigr); \quad S_{i}^{+}S_{i}^{-} = \frac{1}{2}\bigl(1+S_{i}^{z}\bigr); \quad \exp \biggl( \frac{i\pi}{2}S_{i}^{z} \biggr) = iS_{i}^{z}. $$
(2.A.2)

So that, the Jordan-Wigner transformation (2.2.2a), (2.2.2b) can be rewritten as

(2.A.3a)
(2.A.3b)

Using (2.A.3a) and (2.A.3b), one can immediately check that

$$ c_{i}c_{i}^{\dag} + c_{i}^{\dag}c_{i} = S_{i}^{-}S_{i}^{+} + S_i^+ S_i^- = 1. $$
(2.A.4)

To prove \([c_{i},c_{k}^{\dag} ]_{+} = 0\), for ki, without loss of generality we assume k<i. Now

$$ c_{k}c_{i}^{\dag} + c_{i}^{\dag}c_{k} = S_{k}^{-} \prod_{j=k}^{i-1} \bigl[-S_{j}^{z}\bigr]S_{i}^{+}+ S_{i}^{+} \prod_{j=k}^{i-1} \bigl[-S_{j}^{z}\bigr]S_{k}^{-}. $$
(2.A.5)

If one now uses the fact

$$S_{m}^{-}S_{m}^{z} = -S_{m}^{z}S_{m}^{-} $$

the right hand side of (2.A.5) vanishes identically. Similarly, one can derive the other anti-commutation relations.

To express the spin Hamiltonian in terms of the fermion operators, we use the relation, which can be checked using (2.A.3a) and (2.A.3b). Now from (2.A.2), we get

$$ S_{i}^{z}= 2S_{i}^{+}S_{i}^{-}-1=2c_{i}^{\dag}c_{i}-1, $$
(2.A.6)

which can be easily derived using (2.A.3a), (2.A.3b). The coupling term

$$ S_{i}^{x}S_{i+1}^{x}= \bigl[S_{i}^{+}+S_{i}^{-}\bigr] \bigl[S_{i+1}^{+}+S_{i+1}^{-}\bigr] $$
(2.A.7)

is now to be rewritten in terms of fermions. Consider the product

$$ c_{i}^{\dag} c_{i+1} = S_{i}^{+} \bigl[-S_{i}^{z}\bigr]S_{i+1}^{-} \quad \bigl(\mbox{from (2.A.3a), (2.A.3b)}\bigr). $$
(2.A.8)

But \(S_{i}^{+}S_{i}^{z}=-S_{i}^{+}\), so we get

$$ c_{i}^{\dag}c_{i+1} = S_{i}^{+}S_{i+1}^{-}. $$
(2.A.9a)

Similarly one can derive the following relations

(2.A.9b)
(2.A.9c)
(2.A.9d)

The coupling term can be written in terms of fermion operators which are also only coupled to nearest neighbours. This result is a consequence of the Ising character \(S_{i}^{2}= + 1\) of the original degrees of freedom. Collecting all the terms in (2.A.5) and (2.A.9a), (2.A.9b), (2.A.9c), (2.A.9d), one can arrive at the Hamiltonian (2.2.6). One should note here that if the original spin Hamiltonian incorporates next-nearest neighbour interaction, the resulting fermion Hamiltonian includes four-fermion term.

2.1.2 2.A.2 To Diagonalise a General Hamiltonian Quadratic in Fermions

We wish to diagonalise a general quadratic Hamiltonian of the form

$$ H = \sum_{ij} c_{i}^{\dag} A_{ij} c_{j} + \frac{1}{2} \sum _{ij} c_{i}^{\dag} B_{ij} c_{j}^{\dag}+ h.c. $$
(2.A.10)

where c i and \(c_{i}^{\dag}\) are fermion annihilation and creation operators respectively. For system size N, A and B are both N×N matrices. Hermiticity of H demands A to be a Hermitian matrix and anti-commutation of fermion operators demands B to be antisymmetric. Both can be chosen to be real.

Equation (2.A.10) is most effectively diagonalised by Bogoliubov transformation employed for fermion operators by Lieb et al. [245]. In particular, the technique is directly applicable for quite general nearest neighbour interactions, including for example quasiperiodic and random ones [77]. One wants to rewrite the Hamiltonian in a diagonal form using variables η q and \(\eta_{q}^{\dag}\) as

$$ H = \sum_{q}\omega_{q} \eta_{q}^{\dag} \eta_{q} + \mbox{const}, $$
(2.A.11)

ω q are one fermion energies. One makes a linear transformation of the form [245]

(2.A.12a)
(2.A.12b)

where g qi and h qi can be chosen to be real. For η q ’s to satisfy fermionic anti-commutation relations we require

(2.A.13a)
(2.A.13b)

If (2.A.11) holds, then we must have

$$ [\eta_{q},H]_{+} -\omega_{q} \eta_{q} =0. $$
(2.A.14)

Using (2.A.12a), (2.A.12b) in (2.A.14) one finds

(2.A.15a)
(2.A.15b)

The above coupled equations can be written in the following form

(2.A.16a)
(2.A.16b)

where the components of the 2N vectors Φ and Ψ are obtained from the matrices g and h as

(2.A.17a)
(2.A.17b)

One fermion energies ω q are obtained from the eigenvalues of a N×N matrix

(2.A.18a)
(2.A.18b)

For ω q ≠0, either (2.A.18a) or (2.A.18b) is solved for Φ q or Ψ q and the other vector is obtained from (2.A.17a) or (2.A.17b). For ω q =0 these vectors are determined by (2.A.18a), (2.A.18b) or more simply by (2.A.17a), (2.A.17b), their relative sign being arbitrary.

The achievement of the method is obvious. The problem of diagonalising a 2N×2N matrix has been reduced to the eigenvalue problem of a N×N matrix [77, 245]. Even when the eigenvalues of M=(A+B)(AB)=(AB)(A+B)T cannot be found analytically, numerical studies can be made for large systems, for example, in one dimensional nearest-neighbour model with random transverse field (see Sect. 5.3). The price to pay is that one only obtains the one particle energies directly. Multiparticle states have to be built up by linear superposition of the one particle ones. Since A is symmetric and B is antisymmetric both (AB)(A+B) and (A+B)(AB) are symmetric and at least positive semi-definite. Thus all the λ q ’s are real and all the Φ q ’s and Ψ q ’s can be chosen to be real as well as orthogonal. If Φ q ’s are normalised, Ψ q ’s are automatically normalised. To evaluate the constant in (2.A.11) we use the trace invariance of the Hamiltonian under the canonical transformation to the variables η. From (2.A.10)

$$\mathrm{Tr} H = 2^{N-1} \sum_{i} A_{ii}. $$

Again

$$\mathrm{Tr} H = 2^{N-1}\sum_{q} \omega_{q}+2^{N} \times\mbox{const}, $$

from (2.A.11). The constant is thus

$$ \frac{1}{2} \biggl[\sum_{i}A_{ii}- \sum_{q}\omega_{q}\biggr]. $$
(2.A.19)

The complete diagonalised form of the Hamiltonian (2.A.10) is therefore written as

$$ H = \sum_{q}\omega_{q} \eta_{q}^{\dag} \eta_{q} + \frac{1}{2} \biggl(\sum_{i}A_{ii}-\sum _{q}\omega_{q}\biggr). $$
(2.A.20)

We illustrate the above diagonalisation procedure using the example of the one dimensional anisotropic XY chain in a transverse field described by the Hamiltonian (see Sect. 10.1.2)

$$ H = -\frac{J}{2} \sum_{i} \bigl[ (1+ \gamma)S_{i}^{x}S_{i+1}^{x} + (1-\gamma) S_{i}^{y} S_{i+1}^{y} \bigr] - \varGamma\sum_{i}S_{i}^{z} $$
(2.A.21)

where γ is the measure of anisotropy. γ=0 corresponds to the isotropic XY chain in a transverse field, γ=1 corresponds to the transverse Ising chain. The Hamiltonian (2.A.21) is written in terms of Jordan-Wigner fermions as

$$ H = -2 \Biggl( \sum_{i}\biggl(c_{i}^{\dag}c_{i}- \frac{1}{2}\biggr) + \frac{1}{2} \overline{\lambda} \sum _{i=1}^{N} \bigl(c_{i}^{\dag} c_{i+1}+ c_{i}c_{i+1}^{\dag} +\gamma c_{i}^{\dag}c_{i}^{\dag} +\gamma c_{i}c_{i+1}\bigr) \Biggr), $$
(2.A.22)

where \(\overline{\lambda}= J/\varGamma\). Equation (2.A.22) can be put in the general form

$$ H = \biggl( \sum_{ij} c_{i}^{\dag} A_{ij} c_{j} + \frac{1}{2} \sum _{ij} \bigl(c_{i}^{\dag} B_{ij} c_{j}^{\dag}+h.c. \bigr) \biggr) + N, $$
(2.A.23)

where

(2.A.24)

The eigenvalue problem (2.A.18a), (2.A.18b) is solved considering an ansatz wave function exp(i qR j ) and the excitation spectrum is given as

$$ \omega_{q} = \sqrt{ (\gamma\overline{\lambda})^{2} \sin^{2}q + (1-\overline{\lambda}\cos q)^{2}}, $$
(2.A.25)

so that the diagonalised form of the Hamiltonian (with ∑ i A ii =−N) is given by

$$ H = 2\sum_{q} \sqrt{ (\gamma\overline{\lambda} \sin q)^{2}+ (1-\overline{\lambda} \cos q)^{2} } \eta_{q}^{\dag} \eta_{q} - \sum _{q}\omega_{q} $$
(2.A.26)

which reduces to the transverse Ising Hamiltonian (2.2.18) for γ=1.

2.1.3 2.A.3 Calculation of Correlation Functions

Longitudinal spin-spin correlation function is defined as

(2.A.27)

In terms of Jordan-Wigner fermions (with j>i)

$$ C_{ij}^{x} = \langle\psi_{0}| \bigl(c_{i}^{\dag}+c_{i}\bigr) \exp \Biggl( -i \pi \sum_{i=0}^{j-1} c_{l}^{\dag} c_{l} \Biggr) \bigl(c_{j}^{\dag} +c_{j}\bigr)| \psi_{0} \rangle $$
(2.A.28)

where the averages are calculated over the ground state. One can now verify using a representation in which \(c_{l}^{\dag}c_{l}\) is diagonal, that

Defining \(A_{l} = c_{l}^{\dag}+c_{l}\) and \(B_{l} = c_{l}^{\dag}-c_{l}\) and noting that \(A_{l}^{2}= 1\), we have

(2.A.29)

The complicated expression can be simplified using Wick’s theorem, and following relations

$$ \langle \psi_{0}|A_{i}A_{j}| \psi_{0} \rangle = \langle\psi_{0}| \bigl( \delta_{ij}-c_{j}^{\dag}c_{i} + c_{i}^{\dag}c_{j}\bigr)|\psi_{0} \rangle= \delta_{ij}, $$
(2.A.30)

and

(2.A.31)

Only nonzero contractions are 〈A j B i 〉 and 〈B i A j 〉, since 〈A i A j 〉 and 〈B i B j 〉 never occur. Defining 〈B i A i+r 〉=−〈A i+r B i 〉=G ii+r =G r =G r , the correlation function is given by a determinant

(2.A.32)

which is of size r. Similarly one can evaluate the transverse correlation function defined as

(2.A.33)

One can now check that,

(2.A.34)

so that

$$ C_{i\,i+r}^{z} = -G_{i\,i+r}G_{i+r\,i}. $$
(2.A.35)

To evaluate G ii+r we consider the inverse Fourier-Bogoliubov transformation (see (2.2.7a), (2.2.7b) and (2.2.13)) given by

$$ c_i^{\dag} = \sqrt{\frac{1}{N}}\sum _q e^{i q R_i} \bigl(u_q \eta_q^{\dag} + i v_q \eta_{-q} \bigr) $$
(2.A.36)

so that we get

$$ c_i^{\dag} + c_i = \sqrt{\frac{1}{N}} \sum_q \bigl(e^{i q R_i}u_q \eta_q^{\dag} + e^{-i q R_i}u_q \eta_q -i e^{-iq R_i}v_q\eta_{-q}^{\dag} + i e^{i q R_i}v_q\eta_{-q} \bigr) $$
(2.A.37a)

and

(2.A.37b)

Note that u q and v q are given by (2.2.17). In the ground state, G r =G ii+r =〈ψ 0|B i A i+r |ψ 0〉. At a finite temperature T, G r =G ii+r =〈B i A i+r β , where 〈B i A j β denotes an average over the canonical ensemble at temperature k B T=1/β. Thus,

(2.A.38)

The average fermion occupation at temperature T

$$\bigl\langle\eta_{q}^{\dag} \eta_{q} \bigr \rangle_{\beta} = \bigl\langle\eta_{-q}^{\dag} \eta_{-q} \bigr\rangle_{\beta} = \bigl(\exp (\beta\cdot2 \omega_{q})+1\bigr)^{-1} $$

so that,

(2.A.39)

For ground state tanh(βω q /2)=1, we have

$$ G_{r} = \int_{-\pi}^{\pi} \frac{dq}{2\pi} e^{i q r} \biggl(\frac{1 + \overline{\lambda}e^{iq}}{1 + \overline{\lambda}e^{-iq}} \biggr)^{1/2}. $$
(2.A.40)

One can now evaluate the following values G r for some special values of \(\overline{\lambda}\), e.g.,

etc.

In order to evaluate the determinant (2.A.32), let us define

(2.A.41)

In terms of this, (2.A.32) is rewritten as

$$ C_{i\,i+r}^{x} = \det \left ( \begin{array}{l@{\quad}l@{\quad}l@{\quad}l} D_{0} & D_{-1} & \cdots& D_{-r+1} \\ D_{1} & D_{0} & \cdots& D_{-r+2} \\ \cdots& \cdots& \cdots& \cdots\\ D_{r-1} & D_{r-2} & \cdots& D_{0} \end{array} \right ). $$
(2.A.42)

This Toeplitz determinant can be evaluated by using the Szegö’s theorem [393] (see also Ref. [144]), which stated as follows. We refer to Ref. [268] for the proof.

Szegö’s Theorem

Assume that \(\hat{D}(e^{iq})\) and \(\ln\hat{D}(e^{iq})\) are continuous cyclic functions of q for q∈[−π,π]. Let

$$D_p = \int_{-\pi}^{\pi} \frac{dq}{2\pi} \hat{D}\bigl(e^{iq}\bigr) e^{-i q p}, \qquad d_p = \int_{-\pi}^{\pi} \frac{dq}{2\pi} \ln\hat{D}\bigl(e^{iq}\bigr) e^{-i q p}. $$

The Toeplitz determinant

$$\varDelta_r = \det \left ( \begin{array}{l@{\quad}l@{\quad}l@{\quad}l} D_{0} & D_{-1} & \cdots& D_{-r+1} \\ D_{1} & D_{0} & \cdots& D_{-r+2} \\ \cdots& \cdots& \cdots& \cdots\\ D_{r-1} & D_{r-2} & \cdots& D_{0} \end{array} \right ) $$

is given by

$$ \lim_{r\to\infty} \frac{\varDelta_r}{e^{r d_0}} = \exp \Biggl[\sum _{p=1}^{\infty} p d_{-p} d_p \Biggr] , $$
(2.A.43)

whenever the sum in (2.A.43) converges.

Applying this theorem to (2.A.41) and (2.A.42), one finds that the longitudinal correlation function behaves as \(C_{i\,i+r}^{x}\approx A e^{-r/\xi}\) for r≫1 and the correlation length ξ is given by

(2.A.44)

(Note that, when \(\overline{\lambda} < 1\), one needs additional algebras because \(\ln\hat{D}(e^{iq})\) is not cyclic with respect to q. See Ref. [265].) Substituting (2.2.16) for ω q and recovering the lattice constant a and scale of energy Γ, one can write (2.A.44) as

$$ \frac{1}{\xi} = \left\{ \begin{array}{l@{\quad}l} \int_{0}^{\pi/a}\frac{dq}{\pi} \ln\coth\beta\varGamma\sqrt{1 + 2\overline{\lambda}\cos qa + \overline {\lambda}^2} & \mbox{for }\overline{\lambda} > 1 \\ [3pt] \int_{0}^{\pi/a}\frac{dq}{\pi} \ln\coth\beta\varGamma\sqrt{1 + 2\overline{\lambda}\cos qa + \overline {\lambda}^2} - \frac{1}{a}\ln\overline{\lambda} & \mbox{for }\overline{\lambda} < 1 \end{array} \right.. $$
(2.A.45)

Let us now define \(\mathrm{\varDelta} = |1 - \overline{\lambda}|/2\varGamma a^{2}\) and c=2Γa. By making a→0 with fixing β, \(\mathrm{\varDelta}\) and c, and defining a new variable of integral y=βcq, we reach an expression for the continuous limit [338]

$$ \frac{1}{\xi} = \frac{1}{\beta c} f\bigl(\beta\mathrm{ \varDelta}c^2\bigr), $$
(2.A.46)

where

$$ f(x) = \left\{ \begin{array}{l@{\quad}l} \int_{0}^{\infty}\frac{dy}{\pi} \ln\coth\frac{\sqrt{y^2 + x^2}}{2} & \mbox{for }x > 0 \\ [3pt] \int_{0}^{\infty}\frac{dy}{\pi} \ln\coth\frac{\sqrt{y^2 + x^2}}{2} - x & \mbox{for }x < 0 \end{array} \right.. $$
(2.A.47)

One can easily show that the function f(x) is not only continuous but also smooth at x=0. This fact shows that the transverse Ising chain has no singularity at finite temperatures. In special cases, f(x) takes following forms:

$$ f(x) = \left\{ \begin{array}{l@{\quad}l} \sqrt{\frac{2 x}{\pi}}e^{-x} & (x \to\infty) \\ [4pt] \frac{\pi}{4} & (x = 0) \\ [3pt] -x + \sqrt{2|x|}{\pi}e^{-|s|} & (x \to-\infty) \end{array} \right.. $$
(2.A.48)

Substituting this in (2.A.46) and recovering the original parameters, one obtains

$$ \frac{1}{\xi} = \left\{ \begin{array}{l@{\quad}l} \frac{1}{a} \sqrt{\frac{T}{\pi}(\overline{\lambda} - 1)}e^{-2(\overline{\lambda} - 1)/T} & (\overline{\lambda} > 1,\ T \to0) \\[4pt] \frac{\pi}{4}\frac{T}{2a} & (\overline{\lambda} = 1) \\[3pt] \frac{1-\overline{\lambda}}{a} + \frac{1}{a}\sqrt{\frac{T}{\pi}(1-\overline{\lambda})} e^{-2(1 - \overline{\lambda})/T} & (\overline{\lambda} < 1,\ T \to0) \end{array} \right.. $$
(2.A.49)

At the zero temperature, one finds \(\xi= a/(1-\overline{\lambda})\) for \(\overline{\lambda} < 1\), so that the exponent ν turns out to be 1.

The result (2.A.49) implies that the low temperature phase of the transverse Ising chain is divided into three phases by crossover lines \(T\approx|\overline{\lambda} - 1|\). We show a schematic phase diagram in Fig. 2.10.

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Suzuki, S., Inoue, Ji., Chakrabarti, B.K. (2013). Transverse Ising Chain (Pure System). In: Quantum Ising Phases and Transitions in Transverse Ising Models. Lecture Notes in Physics, vol 862. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-33039-1_2

Download citation

Publish with us

Policies and ethics