Classical and Quantum Nonlinear Integrable Systems
Classical and Quantum Nonlinear Integrable Systems
Edited by
A Kundu
Preface
PART I
Classical Systems
1 A journey through the Korteweg–de Vries equation
M Lakshmanan
1.1 Introduction
1.2 Nonlinear dispersive waves: Scott Russell phenomenon and
solitary waves
1.2.1 KdV equation and cnoidal waves and the solitary waves
1.3 The Fermi–Pasta–Ulam (FPU) numerical experiments on
anharmonic lattices
1.3.1 The FPU lattice and recurrence phenomenon
1.4 The KdV equation again!
1.4.1 Asymptotic analysis and the KdV equation
1.5 Numerical experiments of Zabusky and Kruskal: the birth of
solitons
1.5.1 Periodic boundary conditions
1.5.2 Initial condition with just two solitary waves
1.6 Hirota’s bilinearization method: explicit soliton solutions
1.6.1 One-soliton solution
1.6.2 Two-soliton solution
1.6.3 N-soliton solutions
1.6.4 Asymptotic analysis
1.7 The Miura transformation and linearization of KdV: the Lax pair
1.7.1 The Miura transformation
1.7.2 Galilean invariance and the Schrödinger eigenvalue
problem
1.7.3 Linearization of the KdV equation
1.7.4 Lax pair
1.8 Lax pair and the method of inverse scattering
1.8.1 The IST method for the KdV equation
1.9 Explicit soliton solutions
1.9.1 One-soliton solution (N = 1)
Anjan Kundu
April 2003
CLASSICAL SYSTEMS
1.1 Introduction
All around us, Nature is abundant with innumerable phenomena which can be
described by dispersive wave propagation: hydrodynamic waves, acoustic waves,
electromagnetic waves including optical waves, plasma waves, waves on strings
and rods, etc are just some examples. Such dispersive waves are all characterized
by the appropriate dispersion relations. If the dispersion relation is independent
of the amplitude of the underlying waves so that the frequency is a function of
wavenumber alone, ω = ω(k), we have linear dispersive systems described by
linear partial differential equations. In this case, the system can admit wavepackets
or wavegroup solutions which are linear superpositions of a large number of
elementary waves. Since the group velocity vg , in general, differs from the wave
velocity vp (except for the dispersionless case ω(k) = ck, c = constant), the waves
of the group disperse and die down over distance.
However, in nature, not all waves are so gentle that they disperse and
diminish over distance—there can be permanent waves where the dispersion
relation is amplitude-dependent, ω = ω(k, A), where A is the amplitude of the
wave, corresponding to nonlinear dispersive wave propagation. In this case,
for example, there may be a possibility for solitary waves which can travel
without change of speed and shape over long distances to form. In fact, such a
phenomenon was actually observed as far back as in 1834 by the British naval
architect John Scott Russell in the Union Canal connecting the Scottish cities of
Edinburgh and Glasgow and reported in the scientific literature in 1844 [1].
The Korteweg–de Vries (KdV) equation appeared as a sound basic
formulation of the Scott Russell phenomenon in the year 1895 when the Dutch
Like a tsunami wave, this rolling pile of water, a solitary wave, also somehow
maintained its shape and speed which was much larger than conventional linear
dispersive waves. Scott Russell immediately realized that their distinct feature is
their longevity and that they have so much staying power that he could use them to
pump water uphill, which is not possible ordinarily. Scott Russell also performed
some laboratory experiments generating solitary waves by dropping a weight at
one end of a water channel. He was able to deduce empirically that the volume of
water in the wave is equal to the volume of water displaced and further that the
speed, c, of the solitary wave is obtained from the relation
c2 = g(h + a) (1.1)
where a is the amplitude of the wave, h is the undisturbed depth of water and g is
the acceleration due to gravity. A consequence of (1.1) is that taller waves travel
faster!
To put Russell’s formula on a firmer footing, both Boussinesq and Lord
Rayleigh [6] assumed that a solitary wave has a length scale much greater than
the depth of the water. They deduced from the equations of motion for an inviscid
and incompressible fluid, Russell’s formula, (1.1), for c. In fact, they also showed
that the wave profile is given by
1.2.1 KdV equation and cnoidal waves and the solitary waves
The ultimate explanation of the Scott Russell phenomenon was provided by two
Dutch physicists Korteweg and de Vries in 1895 [2]. Starting from the basic
principles of hydrodynamics and considering unidirectional wave propagation in a
long but shallow channel, they deduced the celebrated wave equation responsible
for the phenomenon, which now goes by their names. The KdV equation is a
simple nonlinear dispersive wave equation (for details of the actual derivation
see, for example, [6, 12]). In its modern version, it reads as
where the arbitrary parameters α1 , α2 , α3 and c are related to the three integration
constants of equation (1.5) and are also interrelated as
c α3 − α2
(α1 + α2 + α3 ) = m2 = . (1.6b)
4 α3 − α1
Equation (1.6) represents, in fact, the so-called cnoidal wave for obvious reasons.
Special cases
(i) m ≈ 0: harmonic wave. When m ≈ 0, (1.6) leads to elementary progressing
harmonic wave solutions. This can be verified by taking the limit m → 0
(corresponding to the linearized version of (1.3)) in (1.6).
(ii) m = 1: solitary wave. When m = 1, we can write
√
f = α2 + (α3 − α2 ) sech2 [ α3 − α1 (x − ct)]. (1.7)
´ ¼µ
1
0
-4 -2 0 2 4
Figure 1.1. Solitary wave solution (1.9) of the KdV equation (1.3). Here c = 4.
d2 yi
m = f (yi+1 − yi ) − f (yi − yi−1 ) i = 1, 2, . . . , N − 1 (1.10)
dt 2
with y0 = 0 and yN = 0, where FPU assumed the following specific forms for
f (y):
their great surprise, FPU found that no equipartition of energy occurred. When
the energy was assigned to the lowest mode, as time went on only the first few
modes were excited and even this energy returned to the lowest mode after a
characteristic time called the recurrence time.
Figure 1.3 contains the of FPU’s results for the case N = 32 with α = 0.25
in case (a). Starting with an initial shape at t = 0 in the form of a half of a sine
wave given by yj = sin(j π/32), so that only the fundamental harmonic mode was
excited with an initial amplitude a1 = 4 and energy E1 = 0.077 . . . , the figure
depicts the evolution of the first four normal mode energies, Ek , k = 1, 2, 3, 4.
During the time interval 0 ≤ t ≤ 160 in figure 1.3, where t is measured in periods
of the fundamental mode, modes 2, 3 and 4 etc, sequentially begin to absorb
energy from the initially dominant first mode, as one would expect from a standard
analysis. After this, the pattern of energy sharing undergoes a dramatic change.
Energy is now exchanged primarily only among modes 1 through 6 with all the
higher modes getting very little energy. In fact, the motion is almost periodic, with
a recurrence period (the so-called FPU recurrence) at about t = 157 fundamental
periods. The energy in the fundamental mode returns to within 3% of its value at
t = 0.
The unexpected recurrence phenomenon in the FPU experiments stimulated
a great variety of research into the following domains:
Now considering unidirectional waves (moving to the right), one can make
a change of variables,
ξ = x − ct τ = a 2 ct y = v/a. (1.16)
which is nothing but the KdV equation (1.3) with (α = 6), for a suitable
choice of α but which has now occurred in an entirely new context!
Similarly for α = 0, β = 0, we have
3
ut + βu2 ux + uxxx = 0 (1.21)
2
which we may call as the modified KdV equation or, briefly, the MKdV
equation.
(iv) Scale change: Note that the KdV equation can always be written in the form
under a suitable scale change. Or, in other words, making a change of scales
of the variables t, x and u and redefining the variables, (1.22) can always be
written in the form (1.20). We will use this freedom to choose the coefficients
faster (taller) waves catching up and overtaking the slower (short) waves
(figure 1.4). These nonlinear waves interact strongly and then continue
thereafter almost as if there had been no interaction at all.
(4) Each of the solitary wave pulses moves uniformly at a rate which is linearly
proportional to its amplitude. Thus, the solitons spread apart. Because of
the periodic boundary condition, two or more solitons eventually overlap
spatially and interact nonlinearly (figure 1.4). Shortly after the interaction,
they reappear virtually unaffected in size and shape.
(5) There exists a period TR , the so-called recurrence time at which all the
solitons arrive almost in the same phase and almost reconstruct the initial
state through nonlinear interactions, thereby explaining the FPU recurrence
phenomenon qualitatively.
(6) The persistence of the solitary waves led Zabusky and Kruskal to coin the
name soliton (after names such as the photon, phonon, etc) to emphasize the
particle-like character of these waves which seem to retain their identities in
a collision.
t 2 ∆ / k1
2 ∆ / k2
u2
u1
solution regains its initial shape and, hence, the two solitary waves also regain
their velocities. The only effect of the interaction is a phase shift, that is the centre
of each wave is at a different position than where it would have been if each one
of them were travelling alone (figure 1.6). Again, because of the analogy with
the elastic collision property of particles, Zabusky and Kruskal referred to these
solitary waves as solitons.
Thus, the major new concepts that emerge from the Zabusky–Kruskal
experiments are:
(1) when the nonlinearity suitably balances the linear dispersion as in the KdV
equation, solitary waves can arise;
(2) these solitary waves in appropriate nonlinear systems can interact elastically
like particles without changing their shapes or velocities;
(3) the solitons can constitute the general solution of the initial value problem of
a class of nonlinear dispersive wave equations like the KdV equation.
Naturally, the next obvious question to arise is whether exact analytical forms
of the soliton solutions of the KdV equation, beyond the solitary wave solution,
can be obtained which can correspond to all these numerical results. In fact,
Substituting this into the transformation (1.24), we finally obtain the one-soliton
solution
k2 1 (0)
u(x, t) = 1 sech2 (k1 x − k13 t + η1 ) (1.32)
2 2
which√is the same as the solitary wave solution (1.12), with the identification
k1 = c .
The solution (1.36) when plotted has exactly the same form as in figure 1.5,
thereby showing the soliton nature of the solitary wave as discussed in the
previous section.
and then solve successively for f (2) , . . . , f (N) and finally obtain F and u.
Explicit expressions can be written down with some effort, which we desist from
doing so here due to its somewhat complicated nature. For more details, see, for
example, [9].
u
= + λ. (1.51)
ψ
Equation (1.51) can be re-expressed as
ψx
x
+ (λ − u
)ψ = 0. (1.52)
Omitting the primes for convenience hereafter, we finally obtain the time-
independent Schrödinger-type linear eigenvalue problem
ψxx + (λ − u)ψ = 0 (1.53)
in which the unknown function u(x, t) of the KdV equation appears as a
‘potential’, while (1.53) defining the transformation function ψ(x, t) itself is
linear.
Lψ = λψ (1.55a)
∂2
L=− + u(x, t) (1.55b)
∂x 2
and λ = λ(t) is the eigenvalue at time t. Let the eigenfunction ψ(x, t) evolve as
ψt = Bψ (1.56a)
where, in the case of the KdV equation, the second linear differential operator is
∂3 ∂ ∂
B = −4 + 3 u + u . (1.56b)
∂x 3 ∂x ∂x
Then with the requirement that the eigenvalue λ does not change with time, that
is
λ(t) = λ(0) = constant (1.57)
the compatibility of (1.55b) and (1.56b) leads to the Lax equation or Lax
condition:
Lt = [B, L] = (BL − LB). (1.58)
For the specific forms of L and B given by (1.55b) and (1.56b), the Lax equation
is, indeed, equivalent to the KdV equation (1.45), provided λ is unchanged in
λ = −κn2 n = 1, 2, . . . , N (1.60)
Thus, from the given potential (initial data) u(x, 0) with the boundary
conditions u → 0 as x → ±∞, one can carry out a direct scattering analysis
of (1.59) to obtain the scattering data at t = 0:
S(0) = {κn , Cn (0), R(k, 0), T (k, 0), n = 1, 2, . . . , N, −∞ < k < ∞}.
(1.62)
A typical example to illustrate this is the model potential u(x) =
−A sech2 αx [12].
Since the scattering data is intimately associated with the asymptotic (x → ±∞)
behaviour of the eigenfunction, where the potential u(x, t) → 0, it is enough if
we confine the analysis to this region. Thus as x → ±∞, (1.63) can be written as
∂ 3ψ
ψt = −4 x → ±∞. (1.64)
∂x 3
(a) Scattering states. Without loss of generality, we can write the asymptotic
form of the general solution to (1.64) as
x→−∞
ψ(x, t) −→ a+ (k, t) eikx + a− (k, t) e−ikx (1.65a)
x→+∞ −ikx
−→ b+ (k, t) e ikx
+ b− (k, t) e . (1.65b)
da± db±
= ±4ik 3 a± = ±4ik 3 b± . (1.66)
dt dt
On integration, we get
Comparing (1.66) and (1.68) and making use of (1.67), it is straightforward to see
that
a− (k, t)
= R(k, 0) e−8ik t
3
R(k, t) = (1.69a)
a+ (k, t)
b+ (k, t)
T (k, t) = = T (k, 0) (1.69b)
a+ (k, t)
Substituting (1.72) into (1.71), we see that the normalization constants Cn (t)
evolve as
dCn
= 4κn3Cn (1.73)
dt
so that
3
Cn (t) = Cn (0) e4κn t . (1.74)
Thus, at an arbitrary future instant of time ‘t’, the scattering data S(t)
corresponding to the potential u(x, t) evolves from Sn (0) of the initial data
u(x, 0):
3
S(t) = {κn (t) = κn (0), Cn (t) = Cn (0) e4κn t , n = 0, 1, 2, . . . , N,
R(k, t) = R(k, 0) e−8ik t , −∞ < k < ∞}.
3
(1.75)
K(x, y, t) + F (x + y, t)
∞
+ F (y + z, t)K(x, z, t) dz = 0 y>x (1.76a)
x
N ∞
1
F (x + y, t) = Cn2 (t) e−κn (x+y) + R(k, t) eik(x+y) dk. (1.76b)
n=1
2π −∞
Note that in this equation the time variable ‘t’ enters only as a parameter and
that all the information about the scattering data are contained in the function
F (x + y, t). Solving (1.76), we finally obtain the potential
d
u(x, t) = −2 K(x, x + 0, t). (1.77)
dx
For details, see, for example, [9, 12]. Thus, the initial value problem of the KdV
equation stands solved.
R(k, t) = 0 (1.78)
∂K
= −κK. (1.82)
∂y
On solving, we have
K(x, y, t) = e−κy h(x, t) (1.83)
where the function h(x, t) is to be determined. Substituting (1.83) back into (1.81)
and simplifying, we can find that
3 t −κx
−C02 e8κ
h(x, t) = 3 t −2κx
. (1.84)
[1 + (C02 /2κ) e8κ ]
So the corresponding solution to the KdV equation can be obtained from (1.77)
as
d
u(x, t) = −2 K(x, x + 0, t)
dx
= −2∂K(x, y, t)/∂x|y=x − 2∂K(x, y, t)/∂y|y=x
e−2κ(x−4κ
2 t )−2δ
1
= −2κ 2 δ= log(2κ/C02 )
[1 + e−2κ(x−4κ t )−2δ ]2
2
2
= −2κ sech2 [κ(x − 4κ 2t) + δ].
2
(1.86)
Expression (1.86) is, indeed, the one-soliton solution (1.32) obtained by the Hirota
method. (Note the scale change and negative sign in (1.86), due to the difference
in the coefficients in front of the nonlinear term in (1.45) and (1.3), and also a
redefinition of the parameters.)
where C10 and C20 are the normalization constants corresponding to the two
bound states. Let
Using (1.88) in (1.87) and equating the coefficients of e−κ1 y and e−κ2 y to zero
separately, we can obtain two algebraic equations for h1 and h2 . Solving them,
one obtains h1 (x, t) and h2 (x, t). Using them in (1.88), we obtain from (1.77)
that
d −κ1 y
u(x, t) = −2 [e h1 (x, t) + e−κ2 y h2 (x, t)]
dx
κ 2 cosech2 γ2 + κ12 sech2 γ1
= −2(κ22 − κ12 ) 2 (1.89a)
(κ2 coth γ2 − κ1 tanh γ1 )2
where
2
1 Ci0 (κ2 − κ1 )
γ i = κi x − 4κi3t − δi δi = log i = 1, 2. (1.89b)
2 2κi (κ2 + κ1 )
One can easily check that the form (1.89) is, indeed, the two-soliton solution of
the KdV equation discussed in section 1.6, with appropriate scale change and
redefinition of parameters.
N
N
F (x + y, t) = Cn2 (t) e−κn (x+y) = Cn e−κn x Cn e−κn y
n=1 n=1
= gn (x, t)gn (y, t) gn (x, t) = Cn (t) e−κn x . (1.90)
Then defining
N
K(x, y) = ωn (x)gn (y) (1.91)
n=1
N ∞
ωm (x) + gm (x) + ωn (x) gm (z)gn (z) dz = 0. (1.92)
n=1 x
d d2
u(x, t) = −2 K(x, x + 0, t) = −2 2 log |P| (1.96)
dx dx
as the required N-soliton solution of the KdV equation. It can also be obtained by
the Hirota method as described in section 1.6. One may note the similarity in the
form of (1.96) and the bilinearizing transformation (1.24).
L = [ 12 ψx ψt − ψx3 − 12 ψxx
2
]. (1.102)
Defining
u = ψx (1.104)
(1.103) is seen to reduce to the KdV equation (1.45). Thus (1.102) may be
considered as the Lagrangian of the KdV equation through the potential field
function ψ(x, t).
Defining now the canonically conjugate momentum
∂L ψx
π= = (1.105)
∂ψt 2
Then with the substitution π = ψx /2, one can easily check that the evolution
equation for ψx or π is identical from both (1.108a) and (1.108b). It also coincides
with the KdV ψ field equation (1.103) as it should be. One can, thus, give both a
Lagrangian and Hamiltonian description for the KdV equation and conclude that
it is a Hamiltonian continuous system in the dynamical sense.
One can also give an alternative Hamiltonian description, by writing (1.106)
into terms of the KdV field function u = ψx :
∞
1 2 3
1 2 3
H = ux + u H= ux + u dx. (1.109)
2 −∞ 2
Then writing the Hamiltonian equation of motion for a single field in the form
(for further details see, for example, [9])
∂ δH
ut = (1.110)
∂x δu
we obtain the KdV equation, after using the definition for the functional
derivative.
We know that any transformation from one set of canonical variables (p, q) to a
new set (P , Q) is canonical provided the Poisson brackets of the new set satisfy
the relations
{P , P } = 0 {Q, Q} = 0 {P , Q} = δ(x − x
). (1.112)
Or in other words, if such a transformation exists, then the relation (1.112) ensures
that P and Q are, indeed, canonical variables.
It so happens that for the KdV equation one can find a suitable canonical
transformation from the continuous field variable u(x) to a new set of canonical
N ∞
= − 32
H
5/2
Pj +8 k 3 P (k) dk. (1.113)
5 j =1 −∞
u = ω + ωx + 2 ω2 (1.119)
where is a small parameter and substituting it into the KdV equation, one can
show that u is a solution of the KdV equation provided
∞
ω(x, t; ) = n ωn (x, t) (1.121)
n=0
and substituting it into (1.120), one can equate each power of separately to zero.
Then one obtains the following conservation laws:
Substituting (1.123) into the conservation laws (1.122), one can obtain the
previous conservation laws (1.114) again, as well as further conservation laws
which are infinite in number.
Finally, one can also check that the infinite number of integrals of motion
arising from this are functionally independent and involutive as the Poisson
brackets among them vanish:
∞
δIn ∂ δIm
{In , Im } = dx = 0. (1.124)
−∞ δu(x) ∂x δu(x)
This is yet another property indicative of the complete integrability of the KdV
equation. One can also proceed further and show that the existence of these infinite
number of conserved quantities is intimately related to the existence of an infinite
number of generalized symmetries, the so-called Lie–Bäcklund symmetries. For
more details, see, for example, [16].
Equation (1.125) is often called the potential KdV equation. If ω and ω are any
two solutions of the potential KdV equation (1.125), then the auto-Bäcklund
transformation of it is
ωx + ωx + 2κ 2 + 12 (ω − ω)2 = 0 (1.126a)
ωt + ωt − 3(ωx − ω x )(ωx + ωx ) + ωxxx − ωxxx = 0 (1.126b)
where κ is a real parameter. Then the two equations are compatible with (1.125).
which is nothing but the one-soliton solution of the KdV equation. One can then
use the one-soliton solution in (1.126) as the new ‘seed’ solution and obtain two-
soliton solution and the process can be continued to obtain higher-order solitons.
For more details, we refer to [9, 17, 18].
such that RXi = Xi+1 , which is a new generator of symmetries. Associated with
each of these symmetries, one can identify an independent conserved quantity
and thereby clarify the origin of the existence of the infinite number of involutive
integrals of motion. The existence of Lax pair and Bäcklund transformations
can also be related to the existence of recursion operator and Lie–Bäcklund
symmetries, thereby giving a group theoretical interpretation of the complete
integrability of the KdV equation.
1.15 Conclusion
We have come a long way from Scott Russell’s observation of solitary waves
in 1834 to the modern-day methods of identification of completely integrable
Acknowledgments
I wish to thank Mr T Kanna for help in preparing this article and the Department
of Science and Technology, Government of India for support.
References
[1] Scott Russell J 1844 Report on Waves British Association Reports
[2] Korteweg D J and de Vries G 1895 Phil. Mag. 39 422
[3] Kruskal M D and Zabusky N J Progress on the Fermi–Pasta–Ulam nonlinear
string problem Princeton Plasma Physics Laboratory Annual Report MATT-Q-
21, Princeton, pp 301–8
[4] Zabusky N J and Kruskal M D 1965 Phys. Rev. Lett. 15 240
Belgium
Let us recall that a singularity is said to be movable (as opposed to fixed) if its
location depends on the initial conditions, and critical if multi-valuedness takes
place around it. Other definitions of the PP excluding, for instance, the essential
singularities or replacing ‘movable critical singularities’ by ‘movable singularities
other than poles’ or ‘its general solution’ by ‘all its solutions’ are incorrect. Two
examples taken from Chazy [1] explain why this is so. The first example is the
celebrated Chazy class III equation
− 2uu
+ 3u
2 = 0 (2.1)
(u
− 2u
u
)2 + 4u
2 (u
− u
2 − 1) = 0 (2.2)
ec1 x+c2 c2 − 4
u= + 1 x + c3 (2.3)
c1 4c1
but which also admits a singular solution (envelope solution) with a movable
critical singularity,
u = C2 − log cos(x − C1 ). (2.4)
For more details, see the arguments of Painlevé [2, section 2.6] and Chazy [2,
section 5.1].
The PP is invariant under an arbitrary homography on the dependent variable
and an arbitrary change of the independent variable (homographic group)
Every linear ODE possesses the PP since its general solution depends
linearly on the movable constants so, in order to define new functions, one must
turn to nonlinear ODEs in a systematic way: first-order algebraic equations, then
second-order, etc. The current achievements are the following.
First-order algebraic ODEs (polynomial in u, u
, analytic in x) define only
one function, the Weierstrass elliptic function ℘, new in the sense that its ODE
u
2 − 4u3 + g2 u + g3 = 0 (g2 , g3 ) arbitrary complex constants (2.6)
is not reducible to a linear ODE. Its only singularities are movable double poles.
Second-order algebraic ODEs (polynomial in u, u
, u
, analytic in x) define
six functions, the Painlevé functions Pn, n = 1, . . . , 6, new because they are
not reducible to either a linear ODE or a first-order ODE. This question of
irreducibility, the subject of a long dispute between Painlevé and Joseph Liouville,
has been rigorously settled only recently [3]. The canonical representatives of
P1 : u
= 6u2 + x
P2 : u
= 2u3 + xu + α
u
2 u
αu2 + γ u3 β δ
P3 : u
= − + 2
+ +
u x 4x 4x 4u
u
2 3 β
P4 : u
= + u3 + 4xu2 + 2x 2u − 2αu +
2u 2 u
2
1 1
2 u (u − 1) β u u(u + 1)
P5 : u = + u − + αu + +γ +δ
2u u − 1 x x 2 u x u−1
1 1 1 1 1 1 1
P6 : u
= + + u
2 − + + u
u = u1 + u2 u
1 = 6u21 + x u
2 = 6u22 + x. (2.7)
This ODE (easy to write by the elimination of u1 , u2 ) has a general solution which
depends transcendentally on the four constants of integration and it is reducible.
The master equation P6 was first written by Picard in 1889 in a particular
case, in a very elegant way. Let ϕ be the elliptic function defined by
ϕ
dz
ϕ : y → ϕ(y, x) y= √ (2.8)
∞ z(z − 1)(z − x)
and let ω1 (x), ω2 (x) be its two half-periods. Then the function
2 d2 ψ A B C D 3
− = 2 + + + +
ψ dt 2 t (t − 1) 2 (t − x) 2 t (t − 1) 4(t − u)2
a b
+ + (2.10)
t (t − 1)(t − x) t (t − 1)(t − u)
(A, B, C, D denote constants and a, b parameters). The requirement that the
monodromy matrix (which transforms two independent solutions ψ1 , ψ2 when
t goes around a singularity) be independent of the non-apparent singularity x
results in the condition that u, as a function of x, satisfies P6.
A useful by-product of this search for new functions is the construction
of several exhaustive lists (classifications) of second [11–15], third [1, 4, 6],
fourth [4,5] or higher-order [16] ODEs, whose general solution is explicitly given
because they have the PP. Accordingly, if one has an ODE in such an already
well studied class (e.g. second-order second-degree binomial-type ODEs [14]
u
2 = F (u
, u, x) with F rational in u
and u, analytic in x), and which is
suspected to have the PP (for instance, because one has been unable to detect any
movable critical singularity, see section 2.3), then two cases are possible: either
there exists a transformation (2.5) mapping it to a listed equation, in which case
the ODE has the PP and is explicitly integrated; or such a transformation does not
exist and the ODE does not have the PP.
Definition 2.2 [17, vol III ch XII, 18]. A Bäcklund transformation (BT) between
two given PDEs,
Fj (u, x, t, U, X, T ) = 0 j = 1, 2 (2.12)
u
2 g2 u
2
u
− 6u 2
− −
=0 (2.13)
4u − g2 u − g3
3 2 λ 4u3 − g2 u − g3
which has the PP iff 2πiλ is a period of the elliptic function ℘, a transcendental
condition on (λ, g2 , g3 ) impossible to obtain in a finite number of algebraic steps
ϕ → a
d
− b
c
= 0 (2.18)
c
ϕ + d
p = −1 q = −3 − 2 − γ u0 + 2d −2 u20 = 0. (2.25)
and finally to enforce the necessary condition that, for each family, these two
indices be distinct integers [33, 34]. Considering each family separately would
produce a countable number of solutions, which is incorrect. Considering the two
families simultaneously, the diophantine condition that the two values i1 , i2 of
the Fuchs index 4 + γ u0 be integer has a finite number of solutions, namely [35,
appendix I]
∀i ∈ Z P (i) = 0 : Qi = 0 (2.34)
holds true. At index i = 4, the two conditions, one for each sign of d [36],
are not identically satisfied, so the PDE fails the test. This ends the test.
If, instead of the PDE (2.23), one considers its reduction u(x, t) = U (ξ ), ξ =
x − ct to an ODE, then C = constant = c, and the two conditions Q4 = 0 select
the seven values c = 0 and c2 = (s1 − 3ek )2 (bd)−2 , k = 1, 2, 3. For all these
values, the necessary conditions are then sufficient since the general solution U (ξ )
is single-valued (equation number 8 in Gambier list [11] reproduced in [37]).
+ 3uu
− 4u
2 = 0 (2.36)
which admits the family
p = −2, u0 = −60 Fuchs indices (−3, −2, −1, 20). (2.37)
The series (2.16) depends on two, not four, arbitrary constants, so two are missing
and may contain multi-valuedness. In such cases, one must perform a perturbation
in order to represent the general solution and to test the missing part of the solution
for multi-valuedness.
This perturbation [34] is close to the identity (for brevity, we skip the t
variable)
+∞
+∞
x unchanged u= εn u(n) : E = εn E (n) = 0 (2.38)
n=0 n=0
where, as in the Painlevé α-method, the small parameter ε is not in the original
equation.
Then, the single equation (2.14) is equivalent to the infinite sequence
n=0 E (0) ≡ E(x, u(0) ) = 0 (2.39)
∀n ≥ 1 E (n) ≡ E
(x, u(0) )u(n) + R (n) (x, u(0) , . . . , u(n−1) ) = 0 (2.40)
with R (1) identically zero. From a basic theorem of Poincaré [2, theorem II,
section 5.3], necessary conditions for the PP are:
• the general solution u(0) of (2.39) has no movable critical points;
• the general solution u(1) of (2.40) has no movable critical points; and
• for every n ≥ 2, there exists a particular solution of (2.40) without movable
critical points.
Order zero is just the original equation (2.14) for the unknown u(0) , so one
takes for u(0) the already computed (particular) Laurent series (2.16).
Order n = 1 is identical to the linearized equation
E (1) ≡ E
(x, u(0) )u(1) = 0 (2.41)
and one must check the existence of N independent solutions u(1) locally single-
valued near χ = 0, where N is the order of (2.14).
The two main implementations of this perturbation are the Fuchsian
perturbative method [34] and the non-Fuchsian perturbative method [38]. In this
example (2.36), both methods indeed detect multi-valuedness, at perturbation
order n = 7 for the first one and n = 1 for the second one (details later).
+∞
uj χ j +p
(1)
u(1) = (2.42)
j =ρ
+ 4uu
+ 2u3 = 0 (2.43)
possesses the single family
(0) (0) (0)
p = −1 E0 = u0 (u0 − 1)2 = 0 indices (−1, 0) (2.44)
(0)
with the puzzling fact that u0 should be both equal to 1, according to the equation
(0)
E0 = 0, and arbitrary, according to the index 0. The necessity of performing a
perturbation arises from the multiple root of the equation for u(0)
0 , responsible for
the insufficient number of arbitrary parameters in the zeroth-order series u(0) . The
application of the method provides
E (x, u ) = ∂x2 + 4χ −1 ∂x + 2χ −2
(0)
(2.46)
−1
u(1) = u(1)
0 χ u(1)
0 arbitrary (2.47)
(0) (1)2 (1) (1)
(2)
E = E (x, u
(0) (2)
)u + 6u u + 4u u
(1)2
= χ −2 (χ 2 u(2))
+ 2u0 χ −3 = 0 (2.48)
(1)2
u(2) = −2u0 χ −1 (log χ − 1). (2.49)
The movable logarithmic branch point is, therefore, detected in a systematic way
at order n = 2 and index i = 0. This result was, of course, found long ago by the
α-method [39, section 13, p 221].
Equation (2.36) possesses the two families:
+ 3uu
− 4u
2
(0)
p = −2
(2.50)
2
p = −3 u(0)
0 arbitrary, indices (−1, 0), Ê = 3uu − 4u . (2.51)
= 0. (2.56)
Then the local study of χ = ∞ is unnecessary, since one recognizes the Bessel
equation. The two other solutions in global form are:
c = 0 : v1
v2 = χ 17/2
N23 ( 12c/χ) (2.58)
r = α log Y + R (2.74)
Y −1 = B(χ −1 + A) (2.75)
A and B are two adjustable fields and the gradient of χ is (2.22). The left-hand
side (lhs) of the PDE is then
6
p-mKdV(r) ≡ Ej (S, C, A, B, R)Y j −3 (2.76)
j =0
∀j Ej (S, C, A, B, R) = 0. (2.77)
and the equivalence of the cross-derivative condition (Yx )t = (Yt )x to the mKdV
equation (2.68) for W proves that one has obtained a Darboux involution and a
Lax pair, with the correspondence y = BY.
The auto-BT of mKdV is obtained by the elimination of Y , i.e. by the
substitution
log BY = α −1 (w − W ) dx (2.79)
bC − S − 6λ2 = 0 (2.80)
(log A)
or, equivalently,
(log ω1 )
A = χ −1 + a2 χ + O(χ 3 ) χ = τ − τ2
B = χ −1 + b2 χ + O(χ 3 ) (2.83)
−1
C =χ + c2 χ + O(χ ) 3
has the Fuchs indices −1, −1, −1, 2, 2, 2 and the Gambier test detects no
logarithms at the triple index 2. The Fuchsian perturbative method
N
2+N−n
A = χ −1 εn aj(n) χ j χ = τ − τ2 and cyclically (2.84)
n=0 j =−n
then detects movable logarithms at (n, j ) = (3, −1) and (5, −1) [57] and the
enforcement of these no-log conditions generates the three solutions:
k1 k22 sinh k1 (τ − τ1 )
A= B =C = . (2.88)
sinh k1 (τ − τ1 ) k1 sinh2 k2 (τ − τ2 )
The second constraint (2.86) amounts to suppressing the triple Fuchs index 2,
thus defining a three-dimensional subsystem with a triple Fuchs index −1. One
can, indeed, check that the perturbed Laurent series (2.84) is identical to that of
the Darboux–Halphen system [59, 60]
ω1
= ω2 ω3 − ω1 ω2 − ω1 ω3 and cyclically (2.89)
the Laurent series (2.83) is identical to that of the three-dimensional Euler system
(1750) [61], describing the motion of a rigid body around its centre of mass
ω1
= ω2 ω3 and cyclically (2.91)
ωj = (log(℘ (τ − τ0 , g2 , g3 ) − ej ))
j = 1, 2, 3, (τ0 , g2 , g3 ) arbitrary
(2.92)
℘
2 = 4(℘ − e1 )(℘ − e2 )(℘ − e3 ) = 4℘ 3 − g2 ℘ − g3 . (2.93)
First integrals in the class P (x, y, z) eλt , with P polynomial and λ constant,
should not be searched for with the assumption P the most general polynomial
in three variables. Indeed, P must have no movable singularities. The movable
singularities of (x, y, z) are
1
= 1 + αx + α 2 (x 2 + y + z) + α 3 (x 3 + xy + xz)
(1 − αx)(1 − α 2 y)(1 − α 2 z)
+ α 4 (x 4 + x 2 y + x 2 z + yz + z2 + y 2 ) + · · ·
(2.96)
Six polynomial first integrals are known [65] with a singularity degree at most
equal to four and these are the only ones in the polynomial class [66].
k k
χ −1 = tanh (ξ − ξ0 ) ξ = x − ct k 2 = −2S c=C (2.99)
2 2
the singular part operator D has constant coefficients; therefore, the solutions r
in (2.71) are solitary waves r = f (ξ ), in which f is a polynomial in sech kξ and
1 π 1 π
tanh z − = −2i sech 2z + i tanh z + = 2 tanh 2z + i .
tanh z 2 tanh z 2
(2.100)
In the simpler case of a one-family PDE, the (degenerate) Darboux
involution is
u = D log ψ + U ∂x log ψ = χ −1 (2.101)
and this class of solitary waves r = f (ξ ) reduces to the class of polynomials in
tanh(k/2)ξ . In the example of the chaotic Kuramoto–Sivashinsky (KS) equation
15(16µν − b 2 )
D = 60ν∂x3 + 15b∂x2 + ∂x (2.103)
76ν
k
u = D log cosh (ξ − ξ0 ) + c (c, ξ0 ) arbitrary (2.104)
2
in which b2 /(µν) only takes the values 0, 144/47, 256/73, 16, and k is not
arbitrary. In the quite simple form (2.104), much more elegant than a third-degree
polynomial in tanh, the only nonlinear item is the logarithm, D being linear and
cosh solution of a linear system. This displays the enormous advantage of taking
into account the singularity structure when searching for such solitary waves.
The correct method to obtain all the trigonometric solitary waves of
autonomous PDEs and their elliptic generalization has been recently built [69].
= A2 (U, x)U
2 + A1 (U, x)U
+ A0 (U, x, A, B, ,
). (2.105)
in which j, k = ∞, 0, 1, x.
The well-known confluence from P6 down to P2 then allows us to recover
[73] all the first-degree birational transformations of the five Painlevé equations
(P1 admits no such transformation because it does not depend on any parameter),
thus providing a unified picture of these transformations.
p2
H (q, p, t) = − 2q 3 − tq (2.115)
2
p2
H (q, p, t) = + q5 (2.116)
2
in which the first system is Painlevé-integrable and not Liouville-integrable
and vice versa for the second system. However, given a Liouville-integrable
Hamiltonian system which, in addition, passes the Painlevé test, one must try
to prove its Painlevé integrability by explicitly integrating.
In these three cases, the general solution q1 (hence q22 ) is, indeed, single-valued
and expressed with genus two hyperelliptic functions. This was proven by Drach
in 1919 for the second case, associated to KdV5 and, only recently [74], in the
two other cases. This proof completes the result of [75], who found the separating
variables (a global object) by just considering the Laurent series (a local object),
following a powerful method due to van Moerbeke and Vanhaecke [76].
algebraic in the values of the field variable, with coefficients analytic in x and the
stepsize h or q. As compared to the continuous case, the main missing item is an
undisputed definition for the discrete Painlevé property. The currently proposed
definitions are:
but none is satisfactory. Indeed, the first one says nothing about discrete equations
without continuum limit, and the second one excludes the continuous P6 equation.
is a second-degree polynomial in u
, the coefficient of u
2 is the
sum of, at most, four simple poles in u, etc), directly inherited from the property
of the elliptic equations isolated by Briot and Bouquet. In the discrete counterpart,
the main feature is the existence of an addition formula for the elliptic function ℘
of Weierstrass. As remarked earlier by Baxter and Potts (see references in [83]),
this formula defines an exact discretization of (2.6). Then, all the autonomous
discrete second-order first-degree equations with the (undefined!) discrete PP
have a precise form resulting from the most general discrete differentiation
of the addition formula and the non-autonomous ones simply inherit variable
coefficients as in the continuous case. Of course, the second-order higher-degree
(mostly multi-component) equations are much richer, see details in the review
[84].
Another open question concerns the continuum limit of the contiguity
relation of the ODEs which admit such a relation. The contiguity relation of
the (linear) hypergeometric equation has a continuum limit which is not the
hypergeometric equation but a confluent one. In contrast, the contiguity relation
of the (linearizable) Ermakov equation has a continuum limit which is again an
Ermakov equation [85]. One could argue that the latter depends on a function and
the former only on a finite number of constants. Nevertheless, this could leave
the hope to upgrade from P5 to P6 the highest continuum limit for the contiguity
relation of P6 [73, 86, 87].
2.7 Conclusion
References
[1] Chazy J 1911 Acta Math. 34 317–85
[2] Conte R 1999 The Painlevé Property, One Century Later (CRM Series in
Mathematical Physics) ed R Conte (New York: Springer) pp 77–180 solv-
int/9710020
[3] Umemura H 1990 Nagoya Math. J. 119 1–80
[4] Bureau F J 1964 Ann. Mat. Pura Appl. LXVI 1–116
[5] Cosgrove C M 2000 Stud. Appl. Math. 104 1–65
(http://www.maths.usyd.edu.au:8000/res/Nonlinear/Cos/1998-22.html)
[6] Cosgrove C M 2000 Stud. Appl. Math. 104 171–228
(http://www.maths.usyd.edu.au:8000/res/Nonlinear/Cos/1998-23.html)
[7] Garnier R 1912 Ann. Éc. Norm. 29 1–126
[8] Kudryashov N A and Soukharev M B 1998 Phys. Lett. A 237 206–16
[9] Painlevé P 1906 C. R. Acad. Sci. Paris 143 1111–17
[10] Fuchs R 1905 C. R. Acad. Sci. Paris 141 555–8
[11] Gambier B 1910 Acta. Math. 33 1–55
[12] Bureau F J 1972 Ann. Mat. Pura Appl. XCI 163–281
[13] Cosgrove C M 1993 Stud. Appl. Math. 90 119–87
[14] Cosgrove C M and Scoufis G 1993 Stud. Appl. Math. 88 25–87
[15] Cosgrove C M 1997 Stud. Appl. Math. 98 355–433
[16] Exton H 1971 Rend. Mat. 4 385–448
[17] Darboux G 1894 Leçons sur la théorie générale des surfaces et les applications
géométriques du calcul infinitésimal (4 vol) (Paris: Gauthier-Villars) Reprinted
1972 Théorie générale des surfaces (New York: Chelsea) Reprinted 1993 (Paris:
Gabay)
[18] Matveev V B and Salle M A 1991 Darboux Transformations and Solitons (Springer
Series in Nonlinear Dynamics) (Berlin: Springer)
[19] Ablowitz M J, Kaup D J, Newell A C and Segur H 1974 Stud. Appl. Math. 53 249–315
[20] Cosgrove C M 1993 Stud. Appl. Math. 89 1–61
[21] Cosgrove C M 1993 Stud. Appl. Math. 89 95–151
[22] Mikhailov A V, Shabat A B and Sokolov V V 1991 What is Integrability?
ed V E Zakharov (Berlin: Springer) pp 115–84
[23] Clarkson P A, Fokas A S and Ablowitz M J 1989 SIAM J. Appl. Math. 49 1188–209
[24] Camassa R and Holm D D 1993 Phys. Rev. Lett. 71 1661–4
[25] Degasperis A, Holm D D and Hone A N W 2002 Theor. Math. Phys. 133 1461–72,
nlin.SI/0205023
[26] Ablowitz M J, Ramani A and Segur H 1980 J. Math. Phys. 21 715–21, 1006–15
[27] Mason L J and Woodhouse N M J 1993 Nonlinearity 6 569–81
Kowalevski S 1889 Acta. Math. 12 177–232
Picard E 1893 Acta. Math. 17 297–300
[28] Weiss J, Tabor M and Carnevale G 1983 J. Math. Phys. 24 522–6
Discrete integrability
K M Tamizhmani† and A Ramani† , B Grammaticos‡ and
T Tamizhmani‡
† CPT, Ecole Polytechnique, CNRS, Palaiseau, France
‡ GMPIB, Université Paris VII, France
For the past few centuries, physicists and mathematicians have been familiar
with differential systems. The fundamental equations used in the modelling of
natural phenomena are cast in differential form based on the underlying (often
tacit) assumption that spacetime is continuous. The power of this differential
description lies in the richness of existing tools (although this is an a posteriori
statement) and the success of this approach is undeniable [1]. Integrable systems
hold a privileged position among the differential family. They are, according
to Calogero [2], both ‘universal’ and ‘widely applicable’. Given this situation
why should a physicist be interested in discrete difference systems? Well, for
one, difference systems are ubiquitous in physical modelling. As soon as one
attempts a numerical simulation of a differential system, one has to transcribe it
into an algorithm which is invariably cast in discrete form. It is not exaggerated to
state that our knowledge of the physical universe based on numerical simulations
relies on discrete equations. Still, while numerical algorithms are often considered
as inaccurate approximations of a continuous (differential) ‘reality’, there exist
domains where discrete systems arise naturally. For instance, when some physical
quantity depends on a particular parameter, there exist cases where one can
establish recursion relations thus reducing the computation of this quantity
to that of some basic ‘seed’ one and the solution of the recursion. Solvable
recursion relations are one example of integrable discrete systems. In recent
years, the domain of discrete integrability has undergone a real revolution. As
a consequence discrete analogues have been proposed for most well-known
Mapping (3.2) is presumably integrable and it turns out that indeed it is. As we
have shown in [9], it does possess a Lax pair. Moreover, it is the contiguity relation
of the solutions of the one-parameter PIII equation [10]. Its continuous limit is PI ,
so (3.2) can be considered as its discrete analogue.
/2 + · · · ) + δ(w1 + δw1
+
δ 2 w1
w1
= −w12 + λ1 w1 (3.4)
which is a Riccati equation and has movable poles as its only singularities. Then,
at order δ 3 , we have an equation for w2 :
λ2
λ
w2
= −w2 (λ1 − 2w1 ) − w13 + w12 + λ2 − 1 w1
1
(3.5)
2 2
and similarly at higher orders. Notice that (3.5) is linear for w2 . The same applies
to all subsequent equations. Indeed, at order δ n+1 , we find a differential equation
for the new quantity wn in terms of the ws that have been obtained before. Since
this equation is linear, it cannot have movable singularities when considered as
an equation for wn , everything else being supposed known. However, when we
consider the whole cascade of equations, the subsequent objects will, in general,
have singularities whenever the earlier ones are singular and these singularities
are movable in terms of the whole cascade. Moreover, they are not poles. Already
equation (3.5) shows that, in the neighbourhood of a pole of w1 , where w1 ≈ 1/s
with s = t − t0 (t0 being the location of the movable singularity of w1 ), w2
has logarithmic singularities w2 ≈ −log(s)/s 2 . This singularity is a critical one
which must be considered as movable in terms of the cascade and, therefore, the
perturbative Painlevé property is not satisfied. This is consistent with the fact that
the logistic map is known to be non-integrable.
Although the perturbative Painlevé approach is powerful enough, it is not
without drawbacks. The main critique is based on the fact that not all discrete
systems possess non-trivial continuous limits. In this case, if one does not
have a valid starting point, the whole approach collapses. Moreover, the direct
discretization of Conte and Musette consists in discretizing a given continuous
equation by introducing some freedom and using the perturbative Painlevé
approach in order to pinpoint the integrable subcases. However, this method is
only as good as one’s imagination and if the proposed discretization is not rich
enough, one may miss very interesting cases. In particular, the cases where,
starting from a single equation, the singularity confinement leads to terms of the
form (−1)n , suggesting that the natural form of the mapping is that of a system,
are almost impossible to guess in the direct discretization approach. In [13], we
have presented another example of a mapping:
xn+1 − xn + xn+1
2
+ xn2 + axn+1 xn = 0 (3.6)
1
xn+1 + xn−1 = xn + (3.7)
xn2
a 1
xn+1 + xn−1 = + 2. (3.8)
xn xn
a 1
xn+1 + xn−1 = + 2 (3.10)
xn xn
r 2 + a1 qr − pq 2 qQ4
x2 = x3 =
q2 r(r 2 + a1 qr − pq 2 )2
(r 2 + a1 qr − pq 2 )Q7 qQ4 Q12
x4 = x5 =
qQ24 r(r + a1 qr − pq 2 )Q27
2
Using these two identities, we can easily prove that the characteristic function of a
homographic transformation of f (with constant coefficients) is equal to T (r; f )
up to a bounded quantity. From a theorem due to Valiron [24], we have
P (f )
T r; sup(p, q)T (r; f ) (3.13)
Q(f )
where P and Q are polynomials in f with constant coefficients, of degrees p and
q respectively, provided the rational expression P /Q is irreducible.
Let us also give some useful classical inequalities:
This relation (which is valid for r large enough for any given ) makes it possible
to have access to the characteristic, and thus the order, of the solution of some
difference equations.
As an application of the Nevanlinna approach, we shall examine a mapping
which is related to the q-PIII family:
P (xn )
xn+1 xn−1 = . (3.18)
Q(xn )
As we have shown in [25], all the solutions of (3.18) are of infinite order (except
a finite number of constant solutions) if the maximum of the degrees of P , Q
exceeds 2. The main ingredient in the proof of this result is the inequality
T (r; xn+1 xn−1 ) T (r; xn+1 ) + T (r; xn−1 ), leading to
w
T (r + 1; x) T (r; x). (3.20)
2(1 + )
Now if w > 2, for r large enough one can always choose small enough so
that λ ≡ w/2(1 + ) becomes strictly greater than unity. The precise meaning of
(3.20) is that, for r large enough, we have
T (r + 1; x) ≥ λT (r; x) − C (3.21)
ηxn2 + ζ xn + µ
xn+1 xn−1 = . (3.23)
αxn2 + βxn + γ
m+1 1 1
Xn+1 = Xnm + m − m+1 . (3.24)
Xn+1 Xn
1 7 19 31 41 51 · · ·
1 5 13 19 25 31 · · ·
1 3 5 7 9 11 · · ·
m 1 1 1 1 1 1 ···
−−→
n
At this point, we must indicate how the analytical expression for the degree can
be obtained. First we compute several points on the lattice which allow us to
have a good guess at how the degree behaves. In the particular case of a two-
dimensional discrete equation relating four points on an elementary square like
(3.24) and with the present choice of initial conditions (and given our experience
on one-dimensional mappings), we can reasonably surmise that the dominant
behaviour of the degree will be of the form dnm ∝ mn. Moreover, the sub-dominant
m+1 znm
xn+1 = xnm + . (3.25)
xnm+1 − xn+1
m
(The name ‘potential’ is given here in analogy to the continuous case: the
dependent variable x of equation (3.25) is related to the dependent variable X
of equation (3.24) through xnm+1 − xn+1 m = X m and (3.24) is recovered exactly
n
if zn = 1.) The de-autonomization we are referring to consists in finding an
m
.. .. .. .. .. .. .
. . . . . . ..
1 4 7 10 13 16 · · ·
1 3 5 7 9 11 · · ·
1 2 3 4 5 6 ···
m 1 1 1 1 1 1 ···
−−→
n
.. .. .. .. .. .. .
. . . . . . ..
1 4 10 20 35 56 · · ·
1 3 6 10 15 21 · · ·
1 2 3 4 5 6 ···
m 1 1 1 1 1 1 ···
−−→
n
suffices to reduce the degrees of all higher xs to those of the autonomous case. The
solution of (3.26) is znm = f (n) + g(m) where f , g are two arbitrary functions.
This form of znm is precisely the one obtained in the analysis of convergence
acceleration algorithms [31] using singularity confinement. The integrability of
the non-autonomous form of (3.25) (and its relation to cylindrical KdV) has been
discussed by Nagai and Satsuma [32] in the framework of the bilinear formalism.
zn
δ-PI xn+1 + xn−1 = −xn + +1
xn
zn xn + a
δ-PII xn+1 + xn−1 =
1 − xn2
(xn − aqn )(xn − bqn )
q-PIII xn+1 xn−1 =
(1 − cxn )(1 − dxn )
(xn2 − a 2 )(xn2 − b 2 )
δ-PIV (xn+1 + xn )(xn + xn−1 ) =
(xn − zn )2 − c2
(xn − a)(xn − 1/a)(xn − b)(xn − 1/b)
q-PV (xn+1 xn − 1)(xn xn−1 − 1) =
(1 − xn qn /c)(1 − xn qn /d)
δ-PV / δ-PIV / δ-PII / δ-PI .
Eδ8 / Eδ / Eδ / Dc / Ac / (2A1 )c / Ac
7 6 4 N3N NN 1
NN NN
NN NN
NN NN
NN NN
NN NNN
N& c &
A
2
/ Ac
1
where a, b, c, d are four constants. This is a discrete form of PVI derived first by
Jimbo and Sakai [34]. In the second group, Dc4 , we have [37]
where a, b, c and d are four constants. In the same space, one also has, in a
different direction [38],
(ym − zm )2 − a
xm+1 xm = 2 −b
ym
ζm − c ζm + c
ym + ym−1 = +
1 + dxm 1 + xm /d
i.e. in the form of a homographic mapping, since the latter is the discrete analogue
of the Riccati equation. The coefficients α, β, . . . , θ appearing in (3.33) depend,
in general, on the dependent variable n. The existence of a solution in the form of
(3.33) is possible only when some special relation exists between the parameters
of the discrete Painlevé equation. We shall refer to this relation as ‘linearizability
constraint’. Eliminating yn between (3.33a) and (3.33b), one can obtain a Riccati
relation between xn+1 and xn of the form
an xn + bn
xn+1 = . (3.34)
cn x n + d n
The linearization we referred to earlier is one obtained through a Cole–Hopf
transformation x = u/v. Substituting into equation (3.34), we obtain a linear
equation for v:
We give here the linearization of the two examples we have already presented. For
q
the D5 , asymmetric q-PIII , the linearizability condition has already been obtained
by Jimbo and Sakai [34]. The constraint reads
cdµ = ab (3.36)
√
where µ = λ. The homographic system
yn − aqn
xn+1 = (3.37a)
d(yn − c)
xn − bρn
yn = (3.37b)
c(xn − d)
leads, with x = u/v, to the linear equation
(yn − zn )2 − p2
xn+1 xn = (3.39a)
yn2 − a 2
ζn − r ζn + r
yn + yn−1 = + (3.39b)
1 − bxn 1 − xn /b
a + p + r + δ/2 = 0. (3.40)
with a a constant, where zn and ζn are defined from a single arbitrary function g
of n through zn = gn+1 + gn−1 , ζn = gn+1 + gn , can be solved through the linear
equation
An xn+1 + Bn (xn − a) + An+1 xn−1
=K (3.52)
zn xn+1 + (zn+1 + zn )(xn − a) + zn+1 xn−1
where An = gn2 (gn+1 + gn−1 ) and Bn = −(gn+1 + gn )gn+2 gn−1 − (gn+2 +
gn−1 )gn+1 gn . This mapping, while linearizable, is generically non-confining
unless g is a constant.
However, singularity confinement still plays an important role. As a matter
of fact, while a generic linearizable mapping has linear growth a confining
linearizable mapping has zero degree growth. The simplest example of this is
projective mapping but more complicated examples do exist.
pqr − q
r − qr
u2 =
qr
q 2 (pqr − q
r − qr
+ r
2 − rr
) + r 2 (p
q 2 + qq
− q
2 )
u3 =
(pqr − q
r − qr
)qr
and so on. Since p and t are of degree 0 and q and r of degree 1, we find that
the homogeneity degrees of the numerator and denominator of u2 and u3 are,
respectively, d2 = 2 and d3 = 4. Computing the degree of the successive iterates,
we find dn = 0, 1, 2, 4, 7, 11, 16, 22, . . . , i.e. given by dn = (n2 − n + 2)/2 for
n > 0. The fact that the degree growth is polynomial is not astonishing given that
the Kac–Moerbeke system is integrable. The second system we shall examine is
the semi-discrete mKdV equation [47]:
u
n
un+1 = un−1 + . (3.54)
u2n − 1
αu
n + βu2n + γ un + δ
un+1 = un−1 + (3.55)
κu
n + ζ u2n + ηun + θ
vn
Other equivalent expressions exist for this limit and the notation that is often
used is the truncated power function (x)+ ≡ max(0, x). It is easy to show that
lim→0+ log(ex/ + ey/ ) = max(x, y) and the extension to n terms in the
argument of the logarithm is straightforward.
Two remarks are in order at this point. First, since the function (x)+ takes
only integer values when the argument is integer, the ultra-discrete equations
can describe generalized cellular automata, provided one restricts the initial
conditions to integer values. This approach has already been used in order to
introduce cellular automata (and generalized cellular automata) related to many
interesting evolution equations [50]. Second, the necessary condition for the
t
+ log(1 + δ 2 (eun−1 − 1)) (3.63)
d2 rn
= ern+1 − 2 ern + ern−1 . (3.64)
dt 2
For the ultra-discrete limit, one introduces w through δ = e−L/2 , wnt = utn − L
and takes the limit → 0. Thus, the ultra-discrete limit of (3.63) simply becomes
Equation (3.65) is the cellular automaton analogue of the Toda system (3.64).
Let us now restrict ourselves to a simple periodic case with period two, i.e.
rn+2 = rn and similarly wn+2 = wn . Calling r0 = x and r1 = y, we have from
(3.64) the equation ẍ = 2 ey − 2 ex and ÿ = 2 ex − 2 ey , resulting in ẍ + ÿ = 0.
Thus, x + y = µt + ν and we obtain, after some elementary manipulations,
Equation (3.66) is a special form of the Painlevé PIII equation. Indeed, putting
v = ex−µt /2, we find that
v̇ 2
v̈ = + eµt /2 (a − 2v 2 ). (3.67)
v
The same periodic reduction can be performed on the ultra-discrete Toda equation
(3.65). We introduce w0t = Xt , w1t = Y t and have, in perfect analogy to the
continuous case, Xt +1 − 2Xt + Xt −1 = 2(Y t )+ − 2(Xt )+ and Y t +1 − 2Y t +
Y t −1 = 2(Xt )+ − 2(Y t )+ . Again,
2t (Xt + Y t ) = 0 and we can take Xt + Y t =
mt + p (where m, t, p take integer values). We thus find that X obeys the ultra-
discrete equation:
This is the ultra-discrete analogue of the special form (3.67) of the Painlevé PIII
equation.
In order to construct the ultra-discrete analogues of the Painlevé equations,
we must start with the discrete form that allows the ultra-discrete limit to be taken.
Ultra-discrete forms have been derived for all Painlevé equations [52]. Moreover,
we have shown that their properties are perfectly parallel with those of their
discrete and continuous analogues (degeneration through coalescence, existence
of special solutions, auto-Bäcklund and Schlesinger transformations).
Acknowledgments
This review was made possible thanks to invitations from Paris VII University for
T Tamizhmani and from Ecole Polytechnique for KM Tamizhmani.
References
[1] Grammaticos B and Ramani A 1997 Integrability of Nonlinear Systems (Lect.
Notes Phys. vol 495) ed Y Kosmann-Schwarzbach, B Grammaticos and K M
Tamizhmani (Berlin: Springer) p 30
[2] Calogero F 1990 What is Integrability? ed V Zakharov (Berlin: Springer) p 1
[3] Painlevé P 1902 Acta Math. 25 1
[4] Ramani A, Grammaticos B and Bountis T 1989 Phys. Rep. 180 159
4.1 Introduction
There exists a large class of nonlinear evolution PDEs in one space variable
which can be treated analytically. Such equations are called integrable and
the method for solving the initial value problem on the infinite line for such
equations is called the inverse scattering method or inverse spectral method.
The most well-known integrable equations are the nonlinear Schrödinger, the
Korteweg–deVries and the sine-Gordon equations. The inverse scattering method
is based on a certain mathematical problem in the theory of functions of
one complex variable called the Riemann–Hilbert (RH) problem (see [1] for
an introduction). Some integrable evolution equations in one space dimension
possess particular solutions, which are localized in space and which retain their
shape upon interaction with any other localized disturbance. Such solutions are
called solitons; they are important not because they are exact solutions but because
they characterize the long-time behaviour of the solution. Indeed, it can be shown
that the large-time asymptotics of the solution of integrable evolution equations
in one space dimension is dominated by solitons [2]. Solitons appear in a large
number of physical circumstances, including fluid mechanics, nonlinear optics,
plasma physics, quantum field theory, relativity, elasticity, biological models,
nonlinear networks, etc [3]. This is a consequence of the fact that a soliton is
the realization of a certain physical coherence which is natural to a variety of
nonlinear phenomena.
4.1.2.1 Lumps
The KPI equation is
∂x [qt + 6qqx + qxxx ] = 3qyy . (4.1)
The 1-lump solution of this equation is given by
1
q(x, y, t) = 2∂x ln |L(x, y, t)| + 2
2 2
L = x − 2λy + 12λ2 t + a
4λI
(4.2)
λ = λR + iλI λI > 0
where λ and a are complex constants. Several types of multi-lump solutions are
given in [9].
The focusing DSII equation is
iqt + qzz + qz̄z̄ − 2q(∂z̄−1|q|2z + ∂z−1 |q|2z̄ ) = 0 (4.3)
4.1.2.2 Dromions
The DSI equation is
iqt + (∂x2 + ∂y2 )q + qu = 0
(4.6)
uxy = 2(∂x2 + ∂y2 )|q|2.
The 1-dromion solution of this equation is given by
ρ eX−Ȳ
q(x, y, t) =
α eX+X̄ + β e−Y −Ȳ + γ eX+X̄−Y −Ȳ + δ (4.7)
X = px + ip t 2
Y = qy + iq t 2
|ρ| = 4(Re p)(Re q)(αβ − γ δ)
2
0.75
10
0.5
5
0.25
0
-10
0
-5
0 -5
-10
10
8
|U|2
0
10
5 10
5
0
0
Ŧ5
Ŧ5
Ŧ10 Ŧ10
Y
X
Figure 4.2. The interaction of a 1-lump and a 1-line soliton for KPI.
Thus,
∞ ∞
1 e−ip1 x−ip2 y
G(x, y, k) = dp1 dp2 .
(2π)2 −∞ −∞ p2 − p1 (p1 − 2(kR + ikI ))
Equations (4.13) follow from the identity
∞
dp2 e−ip2 y
= −2πi e−i(a+ib)y (H (y) − H (b)) a, b ∈ R
−∞ p2 − (a + ib)
ψ ± = eikx−ik + G± ±
2y
k qψ
2y
(4.16)
ψ L = eikx−ik + GL qψ L .
where
∞ ∞
i
T ± (k, m) = − dη e−imξ +im η q(ξ, η)ψ ± (ξ, η, k).
2
dξ (4.18)±
2π −∞ −∞
Indeed,
ψ + − ψ L = G+ + + + +
k ψ − G ψ = (Gk − G )ψ + G (ψ − ψ )
L L L L L
or ∞
(ψ + − ψ L ) = dm eimx−im y T + (k, m) + GL (ψ + − ψ L ).
2
This equation and the definition of ψ L immediately imply (4.17a); similarly for
equation (4.17b).
Equations (4.17) together with the analyticity properties of ψ ± yield the
following linear integral equation for µL :
k
− −ikx+ik 2 y
dm eimx−im y T − (k, m)µL (x, y, m)
2
µ (x, y, k) = 1 + P e
L
−∞
∞
+ −ikx+ik 2 y
dm eimx−im y T + (k, m)µL (x, y, m)
2
−P e
k
(4.19)
where ∞
1 f (l) dl
P ± f (k) = k ∈ R. (4.20)
2iπ −∞ l − (k ± i0)
Indeed, multiplying equations (4.17) by e = exp(−ikx + ik 2 y) and using the fact
that P ∓ (µ± − 1) = 0, we find
∞
− −
P (eψ − 1) + P e
L
dm = 0
k
k
P + (eψ L −) + P + e dm = 0
−∞
∞ k
where k dm and −∞ dm denote the right-hand sides of equations (4.17a) and
(4.17b) respectively. Subtracting these equations and using
(P + − P − )(eψ L − 1) = eψ L − 1
q = −2i∂x (µL
1 (x, y)). (4.21)
where S denotes the space of Schwartz functions and q̂0 (ξ, y) denotes the Fourier
transform of q0 (x, y) in the x variable.
Given q0 (x, y), define ψ ± (x, y, k) by equations (4.14)± where q is replaced
by q0 . Given ψ ± (x, y, k), define T ± (k, m) by equation (4.18)± where q is
replaced by q0 . Given T ± , define µL (x, y, t, k) by equation (4.19) where T ±
are replaced by T ± (k, m) exp[i(k 3 + m3 )t]. Given µL (x, y, t, k) define q by
Then q(x, y, t) satisfies the KPI equation (4.1) with q(x, y, 0) = q0 (x, y).
Remark 4.1 (1) The KP equation without the zero mass assumption (4.22) is
studied in [14, 15].
(2) If the small norm assumption (4.23) is violated then equations (4.13)±
can have homogeneous solutions. These homogeneous solutions give rise to
lumps: lumps were formally incorporated into the inverse scattering scheme in [5].
µz̄ − kµ = q (4.24)
Indeed,
lim (z ek̄z−k z̄ µ(x, y, k)) = α(kR , kI ). (4.29)
z→∞
Equation (4.27) implies
∂µ
= ek z̄−k̄z α (4.30a)
∂ k̄
as well as
1
µ=O as k → ∞. (4.30b)
k
These equations define a ∂¯ problem for the function µ. The unique solution of
equations (4.30) is given by
1 e2i(lIx−lR y) α(lR , lI )
µ(x, y, k) = dlR dlI l = lR + ilI . (4.31)
π R2 k−l
Given α, equation (4.31) yields µ, which then implies q through equation (4.24):
1
q(x, y, t) = − dkR dkI e2i(kI x−kR y) α(kR , kI ). (4.32)
π R2
Remark 4.2 If the assumption (4.34) is violated, then equations (4.35) can
have homogeneous solutions. These solutions were formally incorporated in the
inverse scattering scheme in [18] and [19].
∂µ1 ∂µ2
e = e−i(kz+k̄z̄)−i(k
2 +k̄ 2 )t
= T eµ̄2 = T eµ̄1 .
∂ k̄ ∂ k̄
These equations simply
∂(µ1 ± µ2 )
= T e(µ1 ± µ2 ). (4.39)
∂ k̄
Thus, the functions µ1 ± µ2 are generalized analytic functions [21]; therefore, the
solution of equations (4.39) exists without the need for small norm assumption
on T .
4.4 Summary
The dbar method is an effective tool for solving the Cauchy problem on the
plane for two-dimensional integrable nonlinear PDEs. However, up to now it has
not been possible to extend this method to evolution PDEs in higher than two
dimensions. In fact, even the question of the existence of integrable evolution
equations in three or higher dimensions remains open.
Acknowledgment
This work was partially supported by the EPSRC.
References
[1] Deift P 1999 Orthogonal Polynomials and Random Matrices. A Riemann–Hilbert
Approach (Courant Institute Lecture Notes) (New York: New York University)
[2] Deift P and Zhou X 1993 Ann. Math. 137 245
[3] Crighton D 1995 Applications of KdV in KdV’95 ed M Hazewinkel, H Capel and
E de Jager (Amsterdam: Kluwer) pp 2977–84
[4] Calogero F 1993 Important Developments in Soliton Theory ed A S Fokas and
V Zakharov (Berlin: Springer)
[5] Fokas A S and Ablowitz M J 1983 Stud. Appl. Math. 69 211–28
[6] Beals R and Coifman R 1985 Proc. Symp. Pure Math. 43 45
[7] Ablowitz M J, BarYaakov D and Fokas A S 1983 Stud. Appl. Math. 69 135–42
[8] Fokas A S and Santini P M 1990 Physica D 44 99–130
[9] Ablowitz M J and Villaroel J 1997 Phys. Rev. Lett. 78 570–3
5.1 Introduction
We introduce the six-vertex model defined on a two-dimensional square lattice.
We describe the model in detail, since it gives an important prototype of many
solvable lattice models defined on two-dimensional lattices [1]. The transfer
matrix of the six-vertex model generalizes the XXZ quantum spin chain which
plays a central role among integrable quantum spin chains [2,3]. The eight-vertex
model, which generalizes the six-vertex model directly, may be considered as
the most important exactly solvable model in statistical mechanics [4]. Moreover,
many mathematical theories such as the algebraic Bethe ansatz [7] and quantum
groups [5, 6] are closely related to the six-vertex model. Starting from the
six-vertex model, one may have a wide viewpoint on various physical and
mathematical topics related to solvable models. There are quite a large number
of topics related to exactly solvable models in physics and mathematics [1, 8–29].
We explain in section 5.2 some features of the six-vertex model defined on
a square lattice. We introduce the Boltzmann weights and the transfer matrix
for the six-vertex model. We review a method for diagonalizing the transfer
matrix, which is called the coordinate Bethe ansatz, and we give the expressions
of the free energy per site in the ferroelectric, the antiferroelectric and the
disordered phases, respectively [30, 31]. The disordered phase is gapless, while
the ferroelectric and the antiferroelectric phases have gaps [32, 33]. We derive a
critical singularity appearing at the phase transition from the antiferroelectric to
the disordered phase. We review the calculation of the singular part of the free
energy through analytic continuation, as shown in [1]. The critical singularity
is very weak and has an essential singularity similar to the Kosterlitz–Thouless
(KT) transition. We have, thus, derived the KT-like singularity through exact
α + β = γ + δ. (5.1)
α δ
-
6 ? ?
Figure 5.2. Vertex configurations satisfying the ice rule. They have the Boltzmann
weights w(α, β|γ , δ) as follows: (1) w(1, 1|1, 1); (2) w(2, 2|2, 2); (3) w(1, 2|2, 1); (4)
w(2, 1|1, 2); (5) w(1, 2|1, 2); (6) w(2, 1|2, 1). Configurations (1) and (2) are for the
weight a, (3) and (4) for b and (5) and (6) for c.
For an illustration, let us consider the case when α, β, γ and δ are given by +1.
Then, α and β give +2 to the vertex, while γ and δ remove +2 from it, so that the
net charge around the vertex is kept neutral: α + β − γ − δ = 0.
There are only six configurations satisfying the condition. The other
configurations that do not satisfy the condition are not allowed in thermal
equilibrium. We call this the six-vertex model. Condition (5.1) is sometimes
called the ice rule, since ice as a crystal consists of water molecules connected
by hydrogen bonding.
We denote by 1 and 2 the values of polarization 1 and −1, respectively. The
symbols 1 and 2 are useful for matrix notation. Let p denote the notation of ±1
and k 1 and 2. Then, they are related by the relation: k = 1 + (1 − p)/2.
Under the ice rule, there are only six configurations allowed round a vertex. Here,
it is assumed that the energy of a configuration violating the ice rule should be
infinite. We denote by j the energy of the j th vertex configuration shown in
figure 5.2.
Under no external field, the Boltzmann weights must be invariant when
reversing all the polarizations simultaneously. Thus, we have 1 = 2 , 3 = 4
and 5 = 6 , when there is no external field. We denote the Boltzmann weights
as follows:
w(1, 1|1, 1) = w(2, 2|2, 2) = w1 = a
w(1, 2|2, 1) = w(2, 1|1, 2) = w2 = b (5.3)
w(1, 2|1, 2) = w(2, 1|2, 1) = w3 = c.
The Boltzmann weights of the zero-field six-vertex model have essentially
only two parameters. For instance, we may choose a/c and b/c. Note that the
probability for the vertex configuration of a is given by a/(a + b + c), which
does not change by replacing a, b and c with ρa, ρb and ρc.
Let us consider π/2 rotation of the square lattice. If we rotate vertex
configuration (1) of figure 5.2 by the angle π/2 in the counterclockwise direction,
then it becomes vertex configuration (4). Under the π/2 rotation, the weight a is
exchanged with the weight b, while the weight c does not change.
The partition function of the square lattice can be formulated as the trace of
the products of the transfer matrices. Let us define the transfer matrix τ of the
six-vertex model. The matrix elements of the transfer matrix τ acting on N lattice
b1 b2 bN
a ,...,a
Figure 5.3. Matrix elements of the transfer matrix τb 1,...,b N .
1 N
ZNN
= Tr(τ N ) = ,...,aN = 1 + 2 + · · · + 2N .
(τ N )aa11 ,...,a N (5.6)
a1 ,...,aN
, b
, c
). We denote by τ
and τ
, b
, c
),
respectively. If the three sets of Boltzmann weights satisfy the Yang–Baxter
equation
w(α, γ |a1 , a2 )w
(β, b3 |γ , a3 )w
(b1 , b2 |α, β)
α,β,γ
= w
(β, α|a2 , a3 )w
(b1 , γ |a1 , β)w(b2 , b3 |γ , α) (5.8)
α,β,γ
a 2 + b2 − c2
= . (5.9)
2ab
For the zero-field six-vertex model, we can show that if the two sets of Boltzmann
weights have the same value for the parameter
, then their transfer matrices
commute. We shall explicitly discuss in section 5.4 that it is indeed derived from
the Yang–Baxter equations (5.8).
Here, g(x1 , . . . , xn ) denotes the matrix element of vector g for the entry
(x1 , . . . , xn ). In the coordinate Bethe ansatz, we assume the following form for
the matrix element of the possible eigenvector g:
g(x1 , . . . , xn ) = AP exp(kP 1 x1 + · · · + kP n xn ). (5.11)
P ∈Sn
n
exp(iNkj ) = (−1)n−1 exp(−i(kj , k )) for j = 1, . . . , n (5.12)
=1
1 − 2
eip + ei(p+q)
exp(−i(p, q)) = . (5.13)
1 − 2
eiq + ei(p+q)
For the solutions kj to the Bethe ansatz equations, the eigenvalue of the transfer
matrix is given by
where zj = exp(ikj ) for j = 1, . . . , n and the functions L(z) and M(z) are
defined by
ab + (c2 − b 2 )z a 2 − c2 − abz
L(z) = M(z) = . (5.15)
a(a − bz) b(a − bz)
When we discuss the spectrum of an integrable model through the coordinate
Bethe ansatz, we often assume that all the eigenvectors of the transfer matrix
are characterized by the Bethe ansatz wavefunction (5.11). However, it is not
certain whether the assumption is valid or not. Thus, we have to check it by
other methods. In fact, there are several numerical studies on the validity of the
completeness of the Bethe ansatz for some integrable models.
However, there is no doubt about the mathematical structure of the Bethe
ansatz wavefunction. We can derive the expression (5.11) by the algebraic Bethe
ansatz through the ‘two-site’ model [16,79]. (There is an instructive note in [111].)
It was shown that the matrix elements of the product of the B operators acting on
the vacuum are given by the Bethe ansatz wavefunction (5.11) with the kj being
generic.
The Yang–Baxter relation leads not only to the integrability of the six-vertex
model but also to the systematic construction of the eigenvectors. In fact, we shall
see in section 5.4 that the algebraic Bethe ansatz is solely based on the Yang–
Baxter equation.
N
x−1
N
τyx g(y) = a x−y−1bN−x+y−1 c2 g(y) + a N+x−y−1 by−x−1c2 g(y)
y=1 y=1 y=x+1
a N−1 bN−x c2 z
= (a N L(z) + bN M(z))g(x) + (1 − zN ). (5.16)
a − bz
II
III
1
@
@
@
IV @
I
@
@ a/c
0 1
Figure 5.4. Phase diagram of the six-vertex model: regimes I and II are ferroelectric,
regime III is disordered and regime IV is antiferroelectric.
where ρ is the normalization factor. Let us define the rapidity α for the
wavenumber k as
eλ − e−iα sin 12 (α − iλ)
exp(ik) = = − . (5.19)
eλ−iα − 1 sin 12 (α + iλ)
(2) Disordered phase. For −1 <
< 1, we define a positive real parameter µ
by
= − cos µ (0 < µ < π) (5.21)
and we parameterize the Boltzmann weights as follows:
µ−w µ+w
a = ρ sin b = ρ sin c = ρ sin µ (−µ < w < µ).
2 2
(5.22)
−i . (5.36)
m=1
(m − 1/2) sinh((m − 1/2)π 2 /λ)
The real part of the analytic continuation corresponds to the expression for the
antiferroelectric free energy. Therefore, we obtain the singular part of the free
energy
∞
(−1)m cos((m − 1/2)πv/λ) exp(−(m − 1/2)π 2 /λ)
fsing = ikB T . (5.37)
m=1
(m − 1/2) sinh((m − 1/2)π 2 /λ)
We define the reduced temperature t by
t = (a + b − c)/c. (5.38)
In the low-temperature phase (T < Tc ) and near Tc , t is given by
t ≈ − 18 (λ2 − v 2 ). (5.39)
Approximately, t is given by t ≈ −λ2 /8. Near Tc , we have
πv
fsing ≈ −4ikB T e−π /λ cos
2
. (5.40)
2λ
Thus, we have
constant
fsing ∝ exp − √ . (5.41)
−t
Near Tc , the free energy has an essential singularity.
The singularity of the free energy is very close to that of the Kosterlitz–
Thouless transition. In fact, calculating
exactly,
√ we can show that the correlation
length ξ diverges at Tc as ξ ∝ exp constant/ −t , when T approaches Tc in the
antiferroelectric phase.
The integral equation (5.45) can be solved by changing the variable k to rapidity
α and then by taking the Fourier transform for the half-filling case M/N = 1/2.
When M/N is close to 1/2, the integral equation can be solved by the Wiener–
Hopf method [3].
5.2.7 Low-lying excited spectrum of the transfer matrix and conformal field
theory
In this section, we assume that the low-lying excited spectra of the transfer matrix
of gapless models should be characterized by conformal invariance, if the system
Here c iscalled the central charge. We define the operators Ln by the expansion
T (z) = ∞ n=−∞ Ln z
−n−2 . The operator product expansion (5.47) corresponds to
2πv
Eex − E0 = (h + h̄ + N + N̄)
L (5.54)
2π
Pex − P0 = (h − h̄ + N − N̄ )
L
where h and h̄ are the conformal weights related to the zero modes of the field
and the eigenvalues of N and N̄ are given by non-negative integers.
There is another viewpoint on finite-size scaling. Let us consider the t-axis
as the space axis for an infinitely long quantum spin chain, and the x-axis as the
imaginary time axis. Here we assume that L = vβ = v/T . Thus, our system now
becomes the quantum spin chain in the finite temperature T . Replacing E0 with
vβf , where f denotes the finite-temperature free energy of the spin chain, we
have f = −πcT 2 /6v. Thus, we may calculate the specific heat C by the formula
C = −T ∂ 2 f/∂T 2 and we have
πc
C= T. (5.55)
3v
1 2 2π
H= π + (a−n an + ā−n ān ). (5.58)
2gL 0 2L n=0
Here, the integer m is called the winding number. The mode expansion of ϕ is
given by
4π 2πRm 1
ϕ(x, t) = ϕ0 + π0 t + x+i (an e2πin(x−t )/L − ā−n e2πin(x+t )/L).
L L n=0
n
(5.60)
In terms of the coordinates z = exp(2π(τ − ix)/L) and z̄ = exp(2π(τ + ix)/L),
ϕ(x, t) is given by the sum of the holomorphic and antiholomorphic parts:
ϕ(z, z̄) = φ(z) + φ̄(z̄) . Here, they are given by
ϕ0 1
φ(z) = − ia0 log(z) + i ak z−k
2 k=0
k
1 (5.61)
ϕ0
φ̄(z̄) = − iā0 log(z̄) + i āk z̄−k
2 k=0
k
with a0 = n/R + mR/2 and ā0 = n/R − mR/2. Here we can show that the
operator J (z) = i∂φ(z)/∂z is the U (1) current operator.
Making use of Noether’s theorem, we have
2
1 ∂φ(z) 1
T (z) = − : : T̄ (z̄) = − : (∂¯ φ̄(z̄))2 : . (5.62)
2 ∂z 2
1 ∞
L0 = a02 + a−n an . (5.63)
2 n=1
We discuss the finite-size corrections to the XXZ spin chain. The finite-size
corrections to the ground-state energy are calculated in [42–45] using the Euler–
MacLaurin formula [41]. (For a review, see [16, 46, 47].) The result is
πv 2πv 2 2 (
M)2
Eex = Le∞ − + (
D) ξ + + N + N̄ (5.65)
6L L 4ξ 2
2π
Pex − P0 = 2kF
D + (
D
M + N − N̄ ). (5.66)
L
Here the term e∞ denotes the ground-state energy per site. The Fermi velocity is
obtained by the derivative of the dressed energy [16] with respect to the rapidity
at the Fermi level.
The central charge c is given by 1.
D and
M are integers.
M denotes
the change in the number of down spins and
D the number of particles jumping
over the Fermi sea through the backscattering. We note that the difference of
the conformal weights (5.64) is given by hnm − h̄nm = nm. Thus,
M and
D
correspond to the n and m of the c = 1 CFT, respectively. N and N̄ are derived
from particle–hole excitations near the Fermi surface. The Fermi wavenumber kF
is given by kF = πM/L where M is the number of down spins. If the dispersion
is linear, kF is consistent with the number of particles M.
The parameter ξ is given by the dressed charge, which is defined by an
integral equation. We note that the sum of the conformal weights (5.64) is given
by hn,m + h̄n,m = n2 /R 2 + m2 R 2 /4. Thus, the dressed charge ξ corresponds to
the radius R of the c = 1 CFT, ξ = R/2. Under zero magnetic field, the dressed
charge ξ or the radius R is given by [33]
−1/2
1 1
R= − cos−1 (
) . (5.67)
2 2π
where the symbol "i, j # denotes that sites i and j are nearest neighbours and we
take the sum over all the pairs of adjacent sites on the lattice.
There have been many papers written on the two-dimensional Ising model
[8]. However, it should be noted that correlation functions are calculated exactly
for the two-dimensional Ising model in the scaling limit [50, 51].
We note that exact solutions are discussed for the Ising model defined on
various two-dimensional lattices such as the Kagome lattice (for a review, see
[52]).
It is known that the Ashkin–Teller model and the eight-vertex model are in the
same universality class [38]. The universality class is described by c = 1 CFT
with the twisted boson [40].
with a constant Rpqr depending on the three rapidity variables. The constraint
that the Boltzmann weight should have periodicity modulo N: Wpq (n + N) =
Wpq (n) gives, for all rapidity pairs p and q,
N
µp ypN − xqN yqN − ypN
= (µp µq )N = . (5.76)
µq yqN − xpN xpN − xqN
k
1 − kypN
µp = = (5.77)
1 − kxpN k
where k 2 + (k
)2 = 1. Thus, the rapidities are placed on a curve of genus g > 1 of
Fermat type.
We make some comments. Multiplying the two equations of (5.76) and
noting that p and q are independent, we can show equation (5.77) and then
equation (5.78). The star–triangle equation is proven illustratively in the appendix
of [60]. We note that the vertex-type formulation of the chiral Potts model
is related to the tetrahedron equation [63]. The integrable chiral Potts model
generalizes the self-dual ZN model given by Fateev and Zamolodchikov [65]. An
elliptic extension of the self-dual ZN model is introduced in [66], and the Yang–
Baxter equation is proven for the model in [67]. Some non-trivial connections
among higher-rank chiral Potts models, elliptic IRF models and Belavin’s ZN
symmetric model are explicitly discussed in [68].
Let us explain the eight-vertex model which was solved by Baxter in 1972 [4].
The Boltzmann weights w(α, β|γ , δ; u) are non-zero if the charge is conserved
modulo 2: α + β = γ + δ (mod) 2. We have the six configurations around a vertex
shown in figure 5.2 and the two in figure 5.5. We assume that there is no external
field. The weights, therefore, become symmetric: 1 = 2 , 3 = 4 , 5 = 6 and
Figure 5.5. Vertex configurations for the eight-vertex model. (7) w(2, 2|1, 1);
(8) w(1, 1|2, 2).
7 = 8 .
w(1, 1|1, 1) = w(2, 2|2, 2) = w1 = a8V
w(1, 2|2, 1) = w(2, 1|1, 2) = w2 = b8V
(5.79)
w(1, 2|1, 2) = w(2, 1|2, 1) = w3 = c8V
w(1, 1|2, 2) = w(2, 2|1, 1) = w4 = d8V.
We give a parametrization of the Boltzmann weights. We define the theta function
∞
θ (z; τ ) = 2p1/4 sin πz (1 − p2n )(1 − p2n exp(2πiz))(1 − p2n exp(−2πiz))
n=1
(5.80)
where the nome p is related to the parameter τ by p = exp(πiτ ) with Im τ >
0. We also define the theta functions θ0 (z) and θ1 (z) satisfying θα (z + 1) =
(−1)α θα (z) and θα (z + τ ) = i e−πi(z+τ/2) θ1−α (z) for α = 0, 1, where we define
θ1 (z) by θ1 (z; τ ) = θ (z; 2τ ). The Boltzmann weights a8V (z), b8V (z), c8V (z) and
d8V (z) are expressed as
θ0 (z)θ0(2η) θ1 (z)θ0 (2η)
a8V(z) = b8V (z) =
θ0 (z − 2η)θ0 (0) θ1 (z − 2η)θ0 (0)
(5.81)
θ0 (z)θ1 (2η) θ1 (z)θ1 (2η)
c8V (z) = − d8V(z) = − .
θ1 (z − 2η)θ0(0) θ0 (z − 2η)θ0 (0)
where the summation of the variable g is taken over all the admissible states.
The Boltzmann weights are given by
θ (2η − z)
w(d + 1, d + 2, d + 1, d; z, w0 ) = w(d, d − 1, d − 2, d − 1; z) =
θ (2η)
w(d − 1, d, d + 1, d; z, w0 ) = w(d + 1, d, d − 1, d; z, w0 )
θ (z)
=
θ (2η)
√
θ (2η(d + 1) + w0 )θ (2η(d − 1) + w0 )
×
θ (2ηd + w0 )
θ (z + 2ηd + w0 )
w(d + 1, d, d + 1, d; z, w0 ) =
θ (2ηd + w0 )
θ (z − 2ηd − w0 )
w(d − 1, d, d − 1, d; z, w0 ) = . (5.83)
θ (2ηd + w0 )
(b1 , b2 |α, β)
α,β,γ
= w
(β, α|a2 , a3 )w
(b1 , γ |a1 , β)w(b2 , b3 |γ , α). (5.85)
α,β,γ
*
HH 6 6
HH γ a3 a1 β
u H
v
H *
YH
H
HH
H
α H
= H
H α
H H
u + vHH u + v HH
β b3 b1 γ H
HH
v HH
u H
b1 b2 b2 b3
Figure 5.6. The Yang–Baxter equations for vertex models. The spectral parameters are
shown by the angles between pairs of straight lines.
= bc
b
+ ca
c
ab
c
= ba
c
+ cc
b
(5.89)
cb
a
= ca
b
+ bc
c
.
A non-trivial solution (a
, b
, c
− .
cb −ca −bc
ab
(5.90)
a 2 + b2 − c2
= . (5.91)
2ab
The condition that the determinant vanishes is given by
=
. (5.92)
(α, β|γ , δ)
as w(α, β|γ , δ; u), w(α, β|γ , δ; u + v) and w(α, β|γ , δ; v), respectively. The
Yang–Baxter equations are depicted in figure 5.7. As a solution, we may have
(a, b, c) = (ρ sinh(u + 2η), ρ sinh u, ρ sinh 2η). Here we set
= cosh(2η).
The transfer matrices τ (u) and τ (v) commute: τ (u)τ (v) = τ (v)τ (u).
also satisfy the Yang–Baxter equations [82, 83]. Here = ±1, and the number κ
is arbitrary.
The gauge transformation is important in the derivation of the Jones
polynomial from the symmetric Boltzmann weights of the six-vertex model under
zero field [82]. It is also quite useful when we discuss the relation of the six-vertex
model to the quantum group, as we shall see in section 5.5 (see also the appendix
in [95]).
Here, the functions a(z), b(z) and c(z) are equivalent to the Boltzmann weights
a, b and c in section 5.2.
We now introduce the L-operators. We write the matrix element for the L-
j
operator with entry (j, k) as (Ln (z))j k or Ln (z)k . The L-operator for the XXZ
spin chain is given by
Ln (z)11 Ln (z)12 sinh(zIn + ησnz ) sinh 2ησn−
Ln (z) = = . (5.97)
Ln (z)2 Ln (z)2 sinh 2ησn+ sinh(zIn − ησnz )
1 2
Here In and σna (n = 1, . . . , L) are acting on the nth vector space Vn . The
L-operator is an operator-valued matrix which acts on the auxiliary vector
space V0 .
The symbols σ ± denote σ + = E12 and σ − = E21 , and σ x , σ y , σ z are the
Pauli matrices.
In terms of the R-matrix and L-operators, the Yang–Baxter equation can be
expressed as
Here the tensor symbol in Ln (z) ⊗ Ln (t) denotes the tensor product of the
auxiliary spaces.
The Yang–Baxter equation (5.98) gives the relation between the two products
of 4 × 4 matrices. For an illustration, we consider the lhs of (5.98):
a ,a
c ,c
[R(z − t) · Ln (z) ⊗ Ln (t)]b11 ,b22 = R(z − t)ac11,c
,a2
2
(Ln (z) ⊗ Ln (t))b11 ,b22
c1 ,c2
= R(z − t)ac11,c
,a2
L (z)cb11 × Ln (t)cb22 . (5.99)
2 n
c1 ,c2
Here the symbol × denotes the product of matrices acting on the nth space Vn .
Expressing the operator products in the nth space Vn , the lhs of (5.98) can be
a ,a
Figure 5.7. (1) Ln (z)ab |αn ,βn ; (2) R(z)b1 ,b2 .
1 2
αn t −η αn
b1 HH 6
6 6
6 * a1
HH z − η
HH
H c1 a1 b1 c1
HH HH
H *
γn H H γn
H
z−t = z−t H
H
HH H
t − η j
H HH
c2 a2 b2 c2 HH
HH z − η
HH
b2 β β j a2
H
n n
Figure 5.8. The Yang–Baxter equation: R(z − t)(L(z) ⊗ L(t)) = (L(t) ⊗ L(z))R(z − t),
equation (5.98).
written as
a ,a
[R(z − t) · Ln (z) ⊗ Ln (t)]b11 ,b22 αn ,βn
c1 ,c2
= R(z − t)ac11,c
,a2
(L n (z) ⊗ L n (t)) b1 ,b2 α
2 n ,βn
c1 ,c2
c c
= R(z − t)ac11,c
,a2
L (z)b11 αn ,γn Ln (t)b22 γn ,βn .
2 n
(5.100)
c1 ,c2 γn
We define the monodromy matrix by the product of the L-operators (see also
figure 5A.1):
T (z) = LN (z) · · · L2 (z)L1 (z). (5.101)
The transfer matrix τ6V (z) of the six-vertex model is given by the trace of T (z):
A(z) B(z)
τ6V (z) = tr T (z) = A(z) + D(z) where T (z) = . (5.102)
C(z) D(z)
a(t − z) c(t − z)
A(z)B(t) = B(t)A(z) − B(z)A(t) (5.103)
b(t − z) b(t − z)
a(z − t) c(z − t)
D(z)B(t) = B(t)D(z) − B(z)D(t). (5.104)
b(z − t) b(z − t)
N
$ %& '
|0# = |↑#1 |↑#2 · · · |↑#N . (5.105)
Then, through the commutation relations such as (5.103) and (5.104) we can
show [7,16] that the vector |M# gives an eigenvector of the transfer matrix τ6V (z)
if rapidities t1 , t2 , . . . , tM satisfy the set of equations
N
M
a(tj − η) c(tk − tj ) b(tj − tk ) a(tj − tk ) b(tk − tj )
=−
b(tj − η) b(tk − tj ) c(tj − tk ) k=1;k=j
b(tj − tk ) a(tk − tj )
for j = 1, . . . , M. (5.108)
These are the Bethe ansatz equations (5.12) with a different parametrization. For a
set of solutions, t1 , t2 , . . . , tM to equations (5.108), the eigenvalue of the transfer
matrix τ6V (z) is given by
M
a(tj − z) M
a(z − tj )
(z; t1 , t2 , . . . , tM ) = a(z − η)N + b(z − η)N
j =1
b(tj − z) j =1
b(z − tj )
M
sinh(tj − z + 2η)
= sinhN (z + η)
j =1
sinh(tj − z)
M
sinh(tj − z − 2η)
+ sinhN (z − η) . (5.109)
j =1
sinh(tj − z)
which is equivalent to R(z). In terms of the Xj (z), the Yang–Baxter equation can
be expressed as
bj bj +1 bj = bj +1 bj bj +1
bi bj = bj bi for |i − j | > 1. (5.116)
Let us assume that the limit limz→∞ Xj (z) exists. Then, equation (5.115)
becomes
This is nothing but the defining relations of the braid group. Thus, the Boltzmann
weights of solvable models expressed in terms of the spectral parameter lead to
representations of the braid group.
We now show that from a given exactly solvable model, one can derive two
different representations of the braid group [82]. Here, the gauge transformation
(5.93) plays a central role. This technical point is quite fundamental when we
make connections between exactly solvable models and quantum groups and the
Temperley–Lieb algebra (for instance, see the appendix in [95]).
We first consider the Boltzmann weights of the zero-field six-vertex model
given by equations (5.96). Taking the infinite limit to them, we have
1 0 0 0
0 0 exp(−2η) 0
lim X(z)/ sinh(z + 2η) =
. (5.118)
z→∞ 0 exp(−2η) 0 0
0 0 0 1
The representation of the braid group (5.118) leads to a link polynomial equivalent
to the linking number.
Let us now apply the gauge transformation (5.93) with = 1 and κ = 1/2
to the weights given by equations (5.96) [82]. Then, from the transformed
This matrix representation of the braid group leads to the Jones polynomial with
q = exp(2η) [82].
We note that based on the representations of the braid group which are
derived from the Boltzmann weights of exactly solvable models, we can construct
various invariants of knots and links (see, for reviews, [83, 112, 113]).
K − K −1
KX± K −1 = q ±2 X± [X+ , X− ] = . (5.120)
q − q −1
(K) = K ⊗ K
(X+ ) = X+ ⊗ I + K ⊗ X+ (5.121)
(X− ) = X− ⊗ K −1 + I ⊗ X− .
The operation
(·) is called the comultiplication. In the quantum group, the
comultiplication does not commute with the exchange operator σ defined by
σ a ⊗ b = b ⊗ a. However, there is an operator R which satisfies
∞
xn
expq x = q −n(n−1)/2 . (5.124)
n=0
[n]!
, b
, c
). Let us denote by τ
and τ
, b
, c
), respectively. Then, we
can show that if the three sets of Boltzmann weights satisfy the Yang–Baxter
equations given in section 5.2.3, then the transfer matrices τ
and τ
commute.
Let us now explicitly discuss the commutation relation. We first introduce the
monodromy matrix. It is an N ranked tensor, whose (α, β) elements are defined
as follows:
(Tα,β )ab11 ,...,a
,...,bN =
N
w(β, b1 |a1 , c2 )w(c2 , b2 |a2 , c3 ) · · · w(cN , bN |aN , α).
c2 ,...,cN
(5A.1)
The transfer matrix is given by the trace of the monodromy matrix
τ = tr(T ) = Tα,α . (5A.2)
α=1,2
b1 b2 bN
a ,...,a
Figure 5A.1. The matrix element (α, β) of the monodromy matrix (Tα,β )b1 ,...,bN .
1 N
Let us denote by T
and T
, b
, c
. The
γ ,δ
(Tγ
1 ,γN+1 Tδ
(δ1 , b1 |e1 , d2 )w
(d2 , b2 |e2 , d3 ) · · · w
(δ1 , b1 |e1 , d2 ))
c2 ,...,cN d2 ,...,dN e1 ,...,eN
· (w
(c2 , e2 |a2 , c3 )w
(d2 , b2 |e2 , d3 )) · · ·
· (w
(cN , eN |aN , γN+1 )w
cj ,c
Here S(aj , bj )dj ,dj+1
j+1
has been defined by
cj ,c
S(aj , bj )dj ,dj+1
j+1
= w
(cj , ej |aj , cj +1 )w
c ,c
We define the matrix element Md00,d11 as follows:
c ,c
Md00,d11 = w(d0 , d1 |c0 , c1 ). (5A.6)
c ,c
Here we assume that Md00,d11 denotes the matrix element for column (c0 , d0 ) and
row (c1 , d1 ) of the matrix M. Multiplying the matrix M to the product Tα,β
T
γ ,δ
and applying the Yang–Baxter relation N times, we can derive the following:
γ ,c
Mδ00,d11 Tc
1 ,γN+1 Td
1 ,δN+1 = Tc
1 ,cN+1 Td
1 ,dN+1 MdNN ,δN+1
c ,γN+1
. (5A.7)
c1 ,d1 cN ,dN
= T
T
M. Open and
closed circles denote the Boltzmann weights w
and w
(cj , ej |aj , cj +1 )w
(dj , bj |ej , dj +1 ). (5A.9)
ej
M −1 = T
T
. (5A.12)
Noting tr(MT
T
M −1 ) = tr(T
T
= τ
τ
. (5A.13)
QUANTUM SYSTEMS
6.1 Introduction
By quantum integrable systems, we will mean systems with a sufficient number of
higher conserved quantities including the Hamiltonian of the model. Such a notion
of integrability in the Liouville sense allows the system to be described through
action-angle variables with the conserved quantities, which are now operators,
playing the role of action variables. For integrable systems, the conserved
quantities, being functionally independent, should form a commuting set of
operators [cn , cm ] = 0, n, m = 1, 2, . . . , N, such that their total number matches
with the degrees of freedom of the system. For example, a one-dimensional lattice
model of l-sites describing d-mode pseudoparticles should have the number of
conserved quantities N = dl. Note that, for spin- 12 chains, we have d = 1, while
spin-1 and electron models account for d = 2. In this review, we will stick to
single mode d = 1 systems for simplicity and consider mainly periodic lattice
models with N < ∞, where the algebraic structures can be seen in their exact
form. At the lattice constant
→ 0, the field models will be generated from their
exact lattice versions, whenever possible. Integrable field models with N → ∞
consequently need to have an infinite number of conservation laws.
Integrable systems, therefore, are restrictive systems with a very rich
symmetry. The beauty of such models is that they allow exact solutions for
the eigenvalue problem simultaneously for all conserved operators including
the Hamiltonian. Moreover, such one-dimensional quantum systems are also
related to the corresponding two-dimensional classical statistical models with a
fluctuating variable. Therefore, parallel to a quantum mechanical model, one can,
are known as ultralocal models, while the integrable models for which
this ultralocality condition does not hold are classified as non-ultralocal
models. Note that in expressions like (6.2), different auxiliary spaces
mean different tensor products like L1j (λ) = Lj (λ) ⊗ I and L2j (µ) = I ⊗
Lj (µ). The ultralocal property (6.2) generally reflects the involvement of
canonical operators with commutation relations like [u(x), p(y)] = iδ(x − y) or
N
A(λ), B(λ)
Ta (λ) = Laj (λ) T (λ) ≡ . (6.5)
C(λ), D(λ)
j =1
Note that the local and global QYBEs have exactly the same structural form.
Invariance of the algebraic form for the tensor product of the algebras, as revealed
here, indicates the occurrence of the coproduct related to a deep Hopf algebra
one gets the twisted trigonometric and rational R-matrix solution of (6.2), which
12 = b(λ) eiθ , R 21 = b(λ) e−iθ . Apart
may be given by (6.7) with the difference R12 21
from these R-matrices, there can be an elliptic R-matrix solution, for example
that related to the XYZ spin chain and the eight-vertex model [1]. All the
models we consider here, however, are associated with trigonometric or rational
m2
ut t − uxx =
sin(αu)
α (6.11)
ip m sin(λ − αu)
L(λ) = p = u̇.
m sin(λ + αu) −ip
ut t − uxx = eiαu
p ξ eiαu
L(ξ ) = i 1 iαu (6.12)
e −p
ξ
[u(x), p(y)] = iδ(x − y).
where S' and ca± , a = 1, 2 are some operators, the algebraic properties of
which are specified later. The structure of (6.29) becomes clearer if we notice
the decomposition Ltrig (ξ ) = ξ L+ + ξ −1 L− , where L± are spectral-parameter
(anc)
[S 3 , S ± ] = ±S ±
1
[S + , S − ] = (M + sin(2αS 3 ) + M − cos(2αS 3 )) (6.30)
sin α
[M ± , ·] = 0
√
with M ± = ± 12 ±1(c1+ c2− ± c1− c2+ ) behaving as central elements with arbitrary
values of cs. As we have previously mentioned, the integrable systems are
[S 3 , S ± ] = ±S ± [S + , S − ] = [2S 3 ]q (6.35)
with the known form of its coproduct recovered easily from (6.31). The simplest
representation S' = 12 σ' for this case derives from (6.29) the integrable XXZ spin
chain (6.15). On the other hand, representation (6.33) with the corresponding
reduction of (6.34) as
1 1
g(u) = [1 + cos α(2u + 1)] 2
2 sin α
with a suitable choice of parameters s, κ recovers the Lax operator of the lattice
sine-Gordon model (6.16) directly from (6.29) and at its field limit the field Lax
operator (6.11). Note that the spectral dependence in ± appearing in (6.29) can
S + = −κA S − = κA†
1
S 3 = −N κ = −i(cot α) 2 (6.38)
with (6.31) reducing directly to the relation (6.37), which thus gives a new
integrable q-boson model. It is important to note now that either using (6.33)
which simplifies (6.34 ) to g 2 (u) = [−2u]q or directly taking the mapping of the
q-bosons to standard bosons:
1
[2N]q 2
A=ψ N = ψ † ψ,
2N cos α
[S + , S − ] = 0 [S 3 , S ± ] = ±S ± (6.39)
which evidently generates from the same general Lax operator (6.29) the discrete-
time or relativistic quantum Toda chain (6.19). Note that, (iii) and (iv) give the
two different Lax operators found in [14] and [26] for the relativistic Toda chain.
Cases (i) and (ii), however, could be used for constructing non-ultralocal quantum
models, namely the light-cone SG and the mKdV models [23].
ca± → ca± e−iθSk Sk± → S̃k± = e−i 2 θSk Sk± e−i 2 θSk
3 1 3 1 3
(6.41)
[s + , s − ] = 2m+ s 3 + m− [s 3 , s ± ] = ±s ± (6.42)
where m+ = c10 c20 , m− = c11 c20 + c10 c21 and cai , i = 0, 1 are central to (6.42). Note
that (6.42) is a generalization of spin as well as the bosonic algebra and its
coproduct can be obtained as a limit of (6.31). Consequently, the general Lax
operator (6.29) is converted into
0
(anc) c1 (λ + s 3 ) + c11 s−
Lrat (λ) = (6.43)
s+ c20 (λ − s 3 ) − c21
and the quantum R-matrix (6.7) is reduced to its rational form (6.9).
We would see that the ultralocal integrable systems belonging to the rational
class can be generated in a similar way now from the Lax operator (6.43) with
algebra (6.42), all sharing the same rational R-matrix (6.9). It is not difficult
to check by a variable change (u, p) → (ψ, ψ † ) that at the limit α → 0 (6.33)
with only NN interactions, where all non-local factors are cancelled out due to
the relevant properties of the permutation operator. Similarly taking the higher
derivatives of ln τ (λ), the higher conserved quantities cj , j = 2, 3, . . . , N, can
be constructed for these regular models. Note that the conserved operator cj
involves interactions with j + 1 neighbours.
For the simplest case of n = 2, we may take the Lax operator as the R-matrix
given by (6.7), which satisfies clearly the regularity condition L(0) ≡ R(0) = P
for both trigonometric and rational cases. Moreover, for (6.8), the part L
jj +1 (0)
in (6.47) introduces anisotropy reproducing the Hamiltonian of the XXZ spin
chain (6.15). However, since for the rational case (6.9) L
jj +1 (0) = 1, using the
(2) j j +1
expression Pjj +1 = 2βα=1 Eαβ Eβα ≡ 12 (Hjj σ
+1 + 1), where Hjj +1 = σ
σ 'j σ'j +1 ,
(6.47) is clearly reduced to the isotropic spin- 2 Hamiltonian H = j Hjj
1 σ σ
+1
(6.23).
It is intriguing to note that, for the rational class, the same form of the
(n)
Hamiltonian H = N j =1 Pjj +1 with several higher values of n describes most of
the important integrable models, though their physical forms are given mainly
through various representations of the permutation operator. For example, for
n = 3 corresponding to the SU (3) group, we can express the permutation operator
3 j +1
(3) j '
Pjj +1 = βα=1 Eαβ Eβα through spin-1 operators S giving a variant of the
integrable spin-1 model:
H= S'j S'j +1 + (S'j S'j +1 )2 with = +1. (6.48)
j
†
H tJ
= tJ
Hjj +1 = −tP cσj cσj +1 + H.C. P
j j σ =↑↓
1
+ J Sj Sj +1 − nj nj +1 + nj + nj +1 (6.49)
4
where H (a)t J , a = 1, 2 are t–J Hamiltonians (6.49) along two legs. Adding a
suitable Hrung to (6.51) with [H, Hrung] = 0, defining the interaction along the
rung, we finally obtain the integrable t–J ladder model.
Apart from these applications of integrable systems with a similar structure
from the algebraic point of view, we should mention some other important models
like the Hubbard model and the Kondo problem, which also fall into the class of
exactly solvable problems in one dimension [36]. Employing further twisting and
gauge transformations on multi-fermion or multi-spin integrable models, one can
generate another type of integrable model of current interest [37]. The importance
of solvable models in physical systems, their relevance to experiments and related
issues are discussed in [38]. For detailed and involved applications of the Bethe
ansatz technique including that for the theory of correlation functions to various
We have constructed spin and boson (including q-spin and q-boson) models
through the realization of particular Lax operators with inequivalent auxiliary
and quantum spaces. However, in the case of finite-dimensional higher-rank spin
representations, there exists an intriguing method, known as the fusion method,
for obtaining higher-spin models by fusing elementary R-matrices such as (6.7).
Thus, by fusion of only the quantum spaces, one can construct spin-s Lax
operators with a spin- 12 auxiliary space, which is also directly obtainable from
(6.29) as a particular realization. Fusing the auxiliary spaces further, the higher-
spin Lax operator with a spin-s auxiliary space may be constructed as
s
s
Lab = (Pa+ ⊗ Pb+ ) Raj bk (λ + iα(2s − k − j ))(Pa+ ⊗ Pb+ )g2s (λ)
j =1 k=1
(6.52)
+
with Pa(b) as the symmetrizer in the fused spin-s space a(b) and g2s some
normalizing factor [39]. For the rational R-matrix corresponding to (6.9), one
obtains, from (6.52), the integrable spin-s Babujian–Takhtajan model [39] which,
for s = 1, may be given in the same form as Hamiltonian (6.48) but with = −1.
Similarly, for the trigonometric case (6.8), the fused model would correspond to
an integrable anisotropic higher-spin chain.
It should, however, be stressed that such a fusion technique, as far as we
know, has not yet been formulated for bosonic and q-bosonic models. Such an
extension, at least for the restricted values of q, therefore, needs more attention.
The same unified scheme as that described in section 6.3 for constructing
integrable models may also be used to indicate various directions for generating
new integrable classes of quantum and statistical models.
In all previous constructions, the central elements in the ancestor models (6.29)
or (6.43) are chosen as constant parameters. However if they are chosen to be
site-dependent (or even time-dependent) functions, we can get an inhomogeneous
class of models. In these cases, the cs would be attached with site indices cj in the
Lax operators and, similarly, in (6.34) the Mj± would appear as functions, leading
to the corresponding inhomogeneous extensions of known integrable models, e.g.
the inhomogeneous lattice sine-Gordon, Liouville, Toda chain or the NLS model.
However, since the local algebra remains the same as in the original model,
they have the same quantum R-matrices. Although similar, inhomogeneous Toda
chain, NLS models etc were originally proposed as classical systems, they seem
to be new and have not been studied so far as quantum models. Recall that the
impurity models proposed earlier [45] fall into this class and are obtained by a
particular choice of inhomogeneous cj s which amounts to shifting the spectral
parameter. Implementing the same idea on the XXX spin chain, we notice that,
if in its construction along with ca0 = 1 we also choose c21 = −c11 = j resulting
again in m+ = 1, m− = 0, we get the same form of Lax operator but with a
shift Lj (λ − j ), resulting in the Gaudin model (6.24). Similarly higher-spin
representations as well as the su(1, 1) variant would yield other generalizations
of the same model. The commuting set of Hamiltonians for the Gaudin model
Remarkably, the Gaudin model may be mapped into the integrable BCS model,
which is of immense contemporary interest [28].
Physically, such inhomogeneities may be interpreted as impurities, varying
external fields, incommensuration etc.
"s, m̄|S 3 |m, s# = mδm,m̄ "s, m̄|S ± |m, s# = fs± (m)δm±1,m̄ . (6.53)
c1 |m# =
m (0)−1
m (0)|m# c2 |m# = (
m (0)−1
etc, we obtain their respective values, where one may take H = c1 or any other
combination of cs as the Hamiltonian, depending on the concrete model. This
powerful method which is applicable to both integrable quantum and statistical
systems, requires, however, explicit knowledge of the associated Lax operator
and the R-matrix.
It may be noted that the off-diagonal element B(λ) (C(λ)) of the
monodromy matrix (6.5) acts generally as a creation (annihilation) operator
Therefore, the ABA finally solves the eigenvalue problem for τ (λ), yielding
m m
m (λ) = f (λa − λ) α(λ) + f (λ − λa ) β(λ) (6.57)
a=1 a=1
where the Bethe equation (6.56), which is also equivalent to the singularity-free
condition of the eigenvalue (6.57) serves, in turn, as the set of equations for
determining the parameters λa .
Note that in both these equations, α(λ) = ("0|L̂11j (λ)|0#)
N and β(λ) =
22 N
("0|L̂j (λ)|0#) are the only model-dependent parts given by the action of
the upper and lower diagonal operator elements L̂ii j (λ), i = 1, 2, of the Lax
operator of the model on the pseudovacuum. For vertex models, for which the
ABA formulation runs in parallel, the Lax operator elements in these equations
It is remarkable that the rest of the terms in (6.57) and (6.56) are given solely
through the R-matrix elements f (λ) and, therefore, depend only on the related
class (6.8) or (6.9). Recall that in integrable models, as described in section 6.3,
the R-matrix remains the same for all models belonging to a particular class, while
the L-operators differ and may be obtained through various reductions from the
same ancestor Lax operator.
Therefore, taking the Lax operator elements in (6.57) and (6.56) as
those from the general Lax operator (6.29), one may consider the previous
eigenvalue and the Bethe equation to be the unifying equations for exact solution
of all integrable ultralocal quantum and statistical models constructed here.
Consequently, models like the DNLS, SG, Liouville and the XXZ chain together
with the six-vertex model, belonging to the trigonometric class (6.8) should share
similar eigenvalue relations with individual differences appearing only in the form
of the α(λ) and β(λ) coefficients. Thus, this deep-rooted universality in integrable
systems helps us to solve the eigenvalue problem for the whole class of models
and for the full hierarchy of their conserved currents in a systematic way. Let us
present the explicit example of the XXZ chain with Lax operator (6.15), defining
|0# as all spin-up state which gives α(λ) = sinN (λ + α), β(λ) = sinN λ in Bethe
equation (6.56) (with a shift λ → λ + α/2) resulting in
sin(λa + α/2) N m
sin(λa − λb + α)
= (6.58)
sin(λa − α/2) b=a
sin(λa − λb − α)
a=1
sin(λa − λ − α/2)
m
sin(λ − λa + 3(α/2))
+ sinN λ (6.59)
a=1
sin(λ − λa + α/2)
2 1
1 L2j 1
Z 21
1
2
a)
L 1j L 2j
1
1
1 1
2 1 2 1
R12 Z 21 Z 12 R12
2
2
L 2j b) L 1j
[1,k]
[1,k]
T1 T2
1
1
1 1
1 2 k 1 2 k
1
1
2 Z 12
2
R12 Z 21 R12 2
2
1 2 k
1 2 k
[1,k] C) [1,k]
T2 T1
Figure 6.1. Pictorial description of (a) braiding relation (6.61), (b) local braided
QYBE (6.62) for the Lax operators Laj (λa ) and (c) global braided QYBE for
(
Ta[1,k] (λa ) = kj =1 Laj (λa ), k < N. Note that putting Z = 1, i.e. removing the braiding
by undoing the crossing of the broken lines, one can recover the corresponding pictures
for the ultralocal models [1], namely ultralocality condition (6.2) and the local (6.3) and
global QYBE (6.6), respectively.
The next step is the global extension of the BQYBE for the monodromy
matrix (6.5) and it is not difficult to check that due to the braiding relation
[k,j ]
(6.61),
(k the form of the BQYBE is preserved for global matrices like Ta (λ) =
j =1 Laj (λ) (see figure 6.1(c)). However, since for the periodic boundary
condition, one imposes LaN+1 (λ) = La1 (λ), the Lax operators Laj (λ) for j = 1
and j = N again become NN entries and, hence, modify the equation due to the
appearance of an extra Z-matrix from the braiding relation (6.61), leading finally
to the global BQYBE:
−1 −1 −1 −1
R12 (λ − µ)Z21 T1 (λ)Z12 T2 (µ) = Z12 T2 (µ)Z21 T1 (λ)R12 (λ − µ). (6.63)
(1) Current algebra in the WZWN model [53]. The model involves the non-
ultralocal current algebra
γ
{L1 (x), L2 (y)} = [C, L1 (x) − L2 (y)]δ(x − y) + γ Cδ
(x − y) (6.64)
2
+ + + −1
Rq21 L1j L2j = L2j L1j Rq12 L1j L2j +1 = L2j +1 (Rq12 ) L1j . (6.65)
For the details and an interesting quantum group relation of this model, the readers
are referred to the original works [53].
(2) Coulomb gas picture of conformal field theory [54]. The Drinfeld–
Sokolov linear problem—Qx = L(x)Q—describing this system may be given
in the simplest case by the linear operator L(x) = v(x)σ 3 − σ + with a non-
ultralocal property due to the current-like relation {v(x), v(y)} = δ
(x − y).
Discretized and quantized forms of the current-like operator defined through the
braiding matrix. Generalization of this model for SU (N) has also been similarly
constructed in [54].
ultralocal integrable system [57]. The Bethe ansatz solution of the quantum
It may be shown also that (6.69) obeys exactly the previous BQYBE and the
braiding relation with the trigonometric R-matrix (6.8) and the braiding matrix
= eiασ ⊗σ and, consequently, represents a genuine quantum integrable non-
(−) 3 3
Z12
ultralocal model.
Some other non-ultralocal models known in the literature need an
introduction to braiding beyond NN, the basic formulation of which can be found
in [50, 51]. Examples of such models with the same braiding between any two
different sites are: (i) integrable model on moduli space [59], (ii) supersymmetric
models [51, 60], (iii) braided algebra [49], (iv) non-ultralocal extension of the
YBE [61] etc. A unified description of them can be found in [51, 62].
[1] Baxter R 1981 Exactly Solved Models in Statistical Mechanics (New York: Academic
Press)
Deguchi T 2003 Chapter 5 in this book
[2] Faddeev L D 1980 Sov. Sci. Rev. C 1 107
de Vega H J 1989 Int. J. Mod. Phys. A 4 2371
Jimbo M (ed) 1989 Yang–Baxter Equation in Integrable System (Singapore: World
Scientific)
Foerster A, Links J and Zhou H Q 2003 Chapter 8 in this book
[3] Korepin V E, Bogoliubov N M and Izergin A G 1993 QISM and Correlation
Functions (Cambridge: Cambridge University Press)
[4] Reshetikhin N Yu, Takhtajan L A and Faddeev L D 1989 Algebra Analysis 1 178
Faddeev L D 1995 Int J. Mod. Phys. A 10 1845
[5] Sklyanin E 1988 J. Phys. A: Math. Gen. 21 2375
[6] Kulish P and Sklyanin E 1982 Lecture Notes in Physics vol 151 (Berlin: Springer)
[7] Sklyanin E, Takhtajan L and Faddeev L 1979 Teor. Mat. Fiz. 40 194
[8] Faddeev L D and Takhtajan L A, 1986 Lecture Notes Physics vol 246, ed H de Vega
et al (Berlin: Springer)
[9] Kundu A and Basu-Mallick B 1993 J. Math. Phys. 34 1252
[10] Takhtajan L A and Faddeev L D 1979 Russ. Math. Surveys 34 11
[11] Izergin A G and Korepin V E 1982 Nucl. Phys. B 205 [FS5] 401
[12] Kundu A and Basu-Mallick B 1992 Mod. Phys. Lett. A 7 61
[13] Basu-Mallick B and Kundu A 1992 Phys. Lett. B 287 149
[14] Kundu A 1994 Phys. Lett. A 190 73
[15] Suris Yu B 1990 Phys. Lett. A 145 113
[16] Ablowitz M J and Ladik J F 1976 J. Math. Phys. 17 1011
[17] Kundu A 1999 Phys. Rev. Lett. 82 3936
[18] Sklyanin E K 1989 J. Sov. Math. 47 2473
[19] Kundu A and Ragnisco O 1994 J. Phys. A: Math. Gen. 27 6335
[20] Enol’skii V, Salerno M, Kostov N and Scott A 1991 Phys. Scr. 43 229
Enol’skii V, Kuznetsov V and Salerno M 1993 Physica D 68 138
[21] Drinfeld V G 1986 Proc. Int. Cong. Mathematicians (Berkeley) vol 1 p 798
Chari V and Pressley A 1994 A Guide to Quantum Groups (Cambridge: Cambridge
University Press)
[22] Faddeev L D and Tirkkonen O 1995 Nucl. Phys. B 453 647
[23] Kundu A 2002 Phys. Lett. B 550 128
[24] Macfarlane A J 1989 J. Phys. A: Math. Gen. 22 4581
Biederharn L C 1989 J. Phys. A: Math. Gen. 22 L873
[25] Shnirman A G, Malomed B A and Ben-Jacob E 1994 Phys. Rev. A 50 3453
[26] Inoue R and Hikami K 1998 J. Phys. Soc. Japan 67 87
[27] Sogo K, Uchinomi M, Akutsu Y and Wadati M 1982 Prog. Theor. Phys. 68 508
Wadati M, Deguchi T and Akutsu Y 1989 Phys. Rep. 180 247
[28] Dukelsky J and Schuck P 2001 Phys. Rev. Lett. 86 4207
Delft J V and Poghossian R 2002 Phys. Rev. B 66 134502
Foerster A, Links J and Zhou H Q 2003 Chapter 8 in this book
[29] de Vega H and Woynarovich F 1992 J. Phys. A: Math. Gen. 25 4499
7.1 Introduction
The study of integrable models constitutes an important area of theoretical
physics. Integrable models in condensed matter physics describe interacting
many-particle systems. The most prominent examples are interacting spin and
electron systems which include several real materials of interest. Integrable
models, because of their exact solvability, provide a complete and unambiguous
understanding of the variety of phenomena exhibited by real systems. Integrability
in the quantum case implies the existence of N conserved quantities where N is
the number of degrees of freedom of the system. The corresponding operators
including the Hamiltonian commute with each other. More specifically, integrable
models are also described as exactly solvable since the ground-state energy and
the excitation spectrum of the models can be determined exactly. Historically,
the first example of the exact solvability of a many-body problem was that of
a spin- 12 quantum spin chain [1]. The technique used to solve the eigenvalue
problem is now known as the Bethe ansatz (BA) named after Hans Bethe who
formulated it. The demonstration of integrability, namely the existence of N
commuting operators can be made in the more general mathematical framework
of the quantum inverse scattering method (QISM) [2]. The BA has been used
extensively to obtain exact results for several quantum models in one dimension.
Examples include the Fermi and Bose gas models in which particles on a
line interact through delta-function potentials [3], the Hubbard model [4], one-
dimensional (1D) plasma which crystallizes as a Wigner solid [5], the Lai–
Sutherland model which includes the Hubbard model and a dilute magnetic model
as special cases [6], the Kondo model [7], the single impurity Anderson model [8],
the supersymmetric t–J model (J = 2t) etc [9]. The BA method has further
been applied to derive exact results for classical lattice statistical models in two
dimensions.
where x1 , x2 denote the locations of the two particles and k1 , k2 are the momentum
variables. The wavefunction can alternatively be written as
where θ12 is the scattering phase shift. The BA generalizes the wavefunction
(equation (7.1)) to the case of N particles and is given by
N
ψ= A(P ) exp i kPj xj x1 < x2 < · · · < xN . (7.3)
P j =1
N
y y
HXY Z = [Jx Six Si+1
x
+ Jy Si Si+1 + Jz Siz Si+1
z
] (7.4)
i=1
where Siα (α = x, y, z) is the spin operator at the lattice site i, N is the total
number of sites and Jα denotes the strength of the exchange interaction. Consider
the spins to be of magnitude 12 . The eigenvalue problems corresponding to
the isotropic chain (Jx = Jy = Jz = J ) and the longitudinally anisotropic chain
(Jx = Jy = Jz ) were originally solved using the CBA. Later, the same solutions
were obtained using the formalism of QISM [13, 15]. Baxter [18] calculated the
ground-state energy of the fully anisotropic model (equation (7.4)) and Johnson
et al [19] found the excitation spectrum. The results were derived on the basis of a
special relationship between the transfer matrix of the exactly-solved 2D classical
lattice statistical eight-vertex model and the fully anisotropic quantum spin
Hamiltonian HXY Z . Later, the same results were obtained by the ABA approach
of the QISM. The Ising (Jx = Jy = 0) and the XY (Jz = 0) Hamiltonians are
special cases of HXY Z .
Consider the isotropic Heisenberg exchange interaction Hamiltonian in one
dimension:
N
H =J S'i · S'i+1 (7.5)
i=1
with periodic boundary conditions. The sign of the exchange interaction
determines the favourable alignment of the nn spins. J > 0 corresponds to the
antiferromagnetic (AFM) exchange interaction due to which nn spins tend to
be antiparallel. If J < 0 (equivalently, replace J by −J in equation (7.5) with
The Bethe quantum numbers Ii s are integers (half-integers) for odd (even) r. For
a state specified by {I1 , . . . , Ir }, the solution (z1 , . . . , zr ) can be obtained from
equation (7.7). The energy and momentum wavenumber of the state are given by
E − EF r
2
=− (7.8)
i=1 1 + zi
J 2
2π
r
k = πr − Ii (7.9)
N i=1
with EF = JN/4. For the AFM ground state, the Bethe quantum numbers are given
by
N 1 N 3 N 1
{Ii } = − + , − + , . . . , − . (7.10)
4 2 4 2 4 2
In the thermodynamic limit N → ∞, the exact ground-state energy has been
computed as
Eg = NJ (− ln 2 + 14 ). (7.11)
The AFM ground state serves as the physical vacuum for the creation of
elementary excitations. These excitations are not the spin-1 magnons but
spin- 12 spinons [23]. The spinons can be generated systematically by suitable
modifications of the vacuum array of the BA quantum numbers (equation (7.10))
(for details see [22, 23]). For even N, spinons are always created in pairs, each
such pair originating from the removal of one magnon from the ground state.
Since the spinons are spin- 12 objects, the lowest excitations consisting of a pair
of spinons are four-fold degenerate, three triplet (S = 1) and one singlet (S = 0)
excitations. The energy can be written as E(k1 , k2 ) = (k1 ) + (k2 ) where the
spinon spectrum (ki ) = (π/2) sin ki and the total momentum k = k1 + k2 . At a
fixed total momentum k, one gets a continuum of scattering states. The lower
boundary of the continuum is given by (π/2)| sin k| with one of the ki
s equal
to zero. The upper boundary is obtained for k1 = k2 = k/2 and is given by
π| sin k/2|. Figure 7.1 gives an example of a two-spinon configuration.
The BA results are obtained in the thermodynamic limit. In this limit, the
energies and the momenta of the spinons just add up, showing that they do not
interact. Since the spinons are excited in pairs, the total spin of the excited state is
an integer. Inelastic neutron scattering study of the linear chain S = 12 Heisenberg
AFM (HAFM) compound KCuF3 has confirmed the existence of unbound spinon
pair excitations [24]. It is to be noted that in the case of a ferromagnet, the
low-lying excitation spectrum consists of a single magnon branch whereas the
AFM spectrum is a two-spinon continuum with well-defined lower and upper
boundaries.
The dynamical properties of a magnetic system are governed by the time-
dependent pair correlation functions or their spacetime double Fourier transforms
known as the dynamical correlation functions. An important time-dependent
correlation function is
G(R, t) = "S'R (t) · S'0 (0)#. (7.12)
where q' and ω are the momentum wavevector and energy of the spin excitation
and µ = x, y, z. For a particular q' , the peak in S µµ ('
q , ω) occurs at a value of ω
which gives the excitation energy. At T = 0,
µ
q , ω) =
S µµ (' Mλ δ(ω + Eg − Eλ ) (7.14)
λ
is the transition rate between the singlet (Stot = 0) ground state |G# and the triplet
(Stot = 1) states |λ# [25]. Exact calculation of the dynamical correlation functions
in the BA formalism is not possible. Bougourzi et al [26] have used an alternative
approach, based on the algebraic analysis of the completely integrable spin
chain and have calculated the exact two-spinon part of the dynamical correlation
function S xx (q, ω) for the 1D S = 12 AFM XXZ model. In this model, the
Ising part of the XXZ Hamiltonian provides the dominant interaction. Karbach
et al [27] have calculated the exact two-spinon part of S zz (q, ω) for the isotropic
Heisenberg Hamiltonian. In both cases, the size of the chain is infinite. The
exact form of the two-spinon contribution to the dynamical correlation function
S xx (q, ω) of the S = 12 XXZ HAFM chain is complicated and is given by
)
ω ω 2 − χ 2 ω2 ϑ 2 (β c ) |tan(q/2)|−c
0 0 A −
xx
S(2) (q, ω) = 1+ (7.16)
8I ω ω − ω0
2 2 2 c
c=± ϑd (β− ) Wc
where
JK πK
1 − k
I= sinh χ≡ and k, k
≡ 1 − k2
π K 1 + k
γ = πK
/K, t ≡ 2β/K
and ϑd (x) is a Neville theta-function. The derivation of
xx (q, ω) involves generating the two-spinon states from the spinon vacuum,
S(2)
namely the AFM ground state, with the help of spinon creation operators
and expressing the spin fluctuation operator S µ (q) in terms of the spinon
creation operators. The two-spinon part is expected to provide the dominant
contribution to the dynamical correlation function (7.14). For example, in the
case of the isotropic Heisenberg Hamiltonian, the two-spinon excitations account
for approximately 73% of the total intensity in S zz (q, ω). The two-spinon
triplet excitations play a significant role in the low-temperature spin dynamics
of quasi-1D AFM compounds like KCuF3 ,Cu(C6 D5 COO)2 .3D2 O, Cs2 CuCl4
and Cu(C4 H4 N2 (NO3 )2 ) [24, 28]. These excitations can be probed via inelastic
neutron scattering and, hence, a knowledge of the exact dynamical correlation
function is useful. The two-spinon singlet excitations cannot be excited in neutron
scattering because of selection rules (the spinon vacuum |G# is a singlet and
the excited state |λ# in equation (7.15) is a triplet). Linear chain compounds
like CuGeO3 exhibit the spin-Peierls transition [29]. The transition gives rise to
lattice distortion and consequently to a dimerization of the exchange interaction.
Exchange interactions between successive pairs of spins alternate in strength.
There is a tendency for the formation of dimers (singlets) across the strong bonds.
One can construct an appropriate dynamical correlation function in which the
dimer fluctuation operator (DFO) replaces the spin fluctuation operator S µ (q).
The DFO connects the AFM ground state to the two-spinon singlet and not to the
two-spinon triplet.
Two well-known physical realizations of the 1D S = 12 Ising–Heisenberg
compounds are CsCoCl3 and CsCoBr3 . Several inelastic neutron scattering
measurements have been carried out on these compounds to probe the low-
temperature spin dynamics [30]. In these compounds, the Ising part of the
XXZ Hamiltonian is significantly dominant so that perturbation calculations
around the Ising limit are feasible. Near the Ising limit, the exact two-spinon
dynamical correlation function S xx (q, ω) is identical in the lowest order to the
first-order perturbation result of Ishimura and Shiba (IS) [31]. The IS calculation
provides physical insight on the nature of spinons. The Ising part of the XXZ
Hamiltonian is the unperturbed Hamiltonian and the XY part constitutes the
perturbation. The two-fold degenerate Néel states are the ground states of the
where d(l) = (N/π)| sin(πl/N)| is the chord distance between the pair of spins
separated by l sites on a ring with N equally spaced spins. Pij is the spin
exchange operator, Pij = (2S'i · S'j + 12 ). The model is exactly solvable and the
key results are: the ground state has a form similar to the fractional quantum
Hall ground state, the ground state is a quantum spin liquid (QSL) and the
elementary excitations are the spin- 12 spinons obeying fractional statistics, the
thermodynamics as well as the various dynamical correlation functions can be
calculated exactly. The latter calculations are possible because of the simple
structure of the eigenspectrum.
A correct analysis of the BA equations for the S = 12 HAFM in one
dimension gave rise to the concept of spinons which has subsequently been
verified in experiments. Approximate methods like spin wave theory fail to predict
the spinon continuum, thus pointing to the importance of integrable models in
providing the correct physical picture. The existence of spinons in dimensions
greater than one is a highly debatable issue. No precise statement can be made due
to the lack of exact results in d > 1. The issue is of considerable significance in
where the full lines represent dimers (VBs) and the sum runs over all the
plaquettes of the lattice. The first term of the Hamiltonian is the kinetic part
representing the flipping of a pair of parallel dimers on the two bonds of a
plaquette to the other possible orientation, i.e. from horizontal to vertical and
vice versa. The second term counts the number of flippable pairs of dimers in any
dimer configuration and is analogous to the potential term of the Hamiltonian.
The ground state of the QDM on the square lattice is not, however, a QSL except
at the special point t = V . Moessner and Sondhi [41] have studied the QDM on
the triangular lattice and shown that, in contrast to the square lattice case, the
ground state is an RVB state with deconfined, gapped spinons in a finite range
of parameters. Recently, some microscopic models of 2D magnets have been
proposed [42], the low-lying excitations of which are of three types: spinons,
holons and ‘vortex-like’ excitations with no spin and charge, dubbed as visons.
Some of these models are related to the QDM. Two integrable models [42, 43]
which share common topological features with the microscopic models in two
dimensions have been constructed and have applications in fault-tolerant quantum
computation. The models, however, cannot resolve the issue of spinons in two
dimensions as quantum numbers like the total S z are not conserved in these
models. The search for microscopic models in two dimensions, with spinons as
the elementary excitations, acquires particular significance in the light of recent
experimental evidence of the spinon continuum in the 2D frustrated quantum
antiferromagnet Cs2 CuCl4 [44]. The ground state of this compound is expected
to be a QSL with spinons and not magnons as the elementary excitations. Exactly
solvable models in two dimensions are needed for a clear understanding of the
origin of the experimentally observed spinon continuum.
Real materials are often anisotropic in character. The anisotropy may be
present in the exchange interaction Hamiltonian itself or there may be additional
terms in the Hamiltonian corresponding to different types of anisotropy. A
well-known anisotropic interaction, present in many AFM materials, is the
Dzyaloshinskii–Moriya (DM) interaction with the general form
N
y y x
HDM (
) = − (σ x σ − σi σi+1 ) (7.25)
2 i=1 i i+1
where the vector√D ' in equation (7.24) is in the z-direction. The new anisotropy
parameter is δ/ 1 +
2 where δ is the anisotropy parameter of the original
XXZ Hamiltonian. With changed boundary conditions, the model is still BA
solvable. In fact, in the thermodynamic limit (N → ∞), the boundary conditions
do not affect the critical behaviour. Thus, the Hamiltonian, which includes both
the XXZ Hamiltonian and the DM interaction, has the same critical properties
and
√the phase diagram as the XXZ Hamiltonian with the anisotropy parameter
δ/ 1 +
2 .
We next turn our attention to spin-S (S > 12 ) quantum spin chains. The spin-S
Heisenberg exchange interaction Hamiltonian in one dimension is not integrable.
A family of Heisenberg-like models has been constructed for S = 1, 32 , 2, 52 , . . .
etc for which the spin-S quantum Hamiltonian is given by
Hs = Q(S'i · S'i+1 ) (7.26)
i
with θ varying between 0 and 2π. The biquadratic term has been found to be
relevant in some real integer-spin materials. There are two gapped phases: the
Haldane phase for − 14 π < θ < 14 π and a dimerized phase for − 34 π < θ < − 14 π
[53]. At θ = − 14 π, the model is integrable and the gap vanishes to zero. This point
separates the two gapped phases, Haldane and dimerized, which have different
symmetry properties. Thus, a quantum phase transition occurs at ϑ = − 14 π from
the Haldane to the dimerized phase. The integrable model provides the exact
location of the transition point. The point θ = 14 π corresponds to the Hamiltonian
which is a sum over the permutation operators and is again exactly solvable.
The Haldane phase includes the isotropic Heisenberg chain (θ = 0) and the
Affleck–Kennedy–Lieb–Tasaki (AKLT) Hamiltonian (tan θVBS = 13 ) [54]. The
latter model is not integrable but the ground state is known exactly. The ground
state is described as a valence bond solid (VBS) state in which a VB (singlet)
covers every link of the chain. Since the gap does not become zero for 0 ≤ θ ≤
θVBS , there is no phase transition in going from one limiting Hamiltonian to the
other. Thus, the isotropic Heisenberg and AKLT chains are in the same phase.
The doped cuprate systems exhibit a variety of novel phenomena in their
insulating, metallic and superconducting phases. A full understanding of these
phenomena is, as yet, lacking. There is currently a strong research interest in
doped spin systems. The idea is to look for simpler spin systems in which the
consequences of doping can be studied in a less ambiguous manner. The spin-1
HG nickelate compound Y2 BaNiO5 can be doped with holes on replacing the off-
chain Y3+ ions by Ca2+ ions. Inelastic neutron scattering (INS) measurements on
the doped compound provide evidence for the appearance of new states in the HG
[55]. The structure factor S(q), obtained by integrating the dynamical correlation
function S(q, ω) over ω, acquires an incommensurate, double-peaked form in
the doped state [56]. Frahm et al [57] have constructed an integrable model
describing a doped spin-1 chain. In the undoped limit, the spectrum is gapless and
so the HG of the integer spin system is not reproduced. It is, however, possible to
reintroduce a gap in the continuum limit where a field-theoretical description of
the model is possible. The model has limited relevance in explaining the physical
features of the doped nickelate compound. Another interesting study relates to the
appearance of magnetization plateaus in the doped S = 1 integrable model [58].
The location of the plateaus depends on the concentration of holes. Experimental
evidence of this novel phenomenon has not been obtained so far.
An electron in a solid, localized around an atomic site, has three degrees
of freedom: charge, spin and orbital. The orbital degree of freedom is relevant
to several transition metal oxides which include the cuprate and manganite
systems. The latter compounds on doping exhibit the phenomenon of colossal
The nn intra-chain and the rung exchange interactions are of strength J and
JR respectively. When JR = 0, one obtains two decoupled AFM spin chains for
which the excitation spectrum is known to be gapless. For all JR /J > 0, a gap
JR JR
Figure 7.2. A two-chain ladder. The rung and intra-chain nn exchange interactions are of
strength JR and J respectively.
where k is the momentum wavevector. The SG, defined as the minimum excitation
energy, is given by
SG = ω(π) (JR − J ). (7.31)
The two-spin correlations decay exponentially along the chains showing that the
ground state is a QSL. The magnons can further form bound states. Experimental
evidence of two-magnon bound states has been obtained in the S = 12 two-
chain ladder compound Ca14−x Lax Cu24 O41 (x = 5 and 4) [67]. The family of
compounds Srn−1 Cun+1 O2n consists of planes of weakly-coupled ladders of
(n + 1)/2 chains [68]. For n = 3 and 5, respectively, one gets the two-chain
and three-chain ladder compounds SrCu2 O3 and Sr2 Cu3 O5 respectively. For
the first compound, experimental evidence of the SG has been obtained. The
latter compound has properties similar to those of the 1D Heisenberg AFM
chain [69]. A recent example of a spin ladder belonging to the organic family
of materials is the compound (C5 H12 N)2 CuBr4 , a ladder system with strong
rung coupling (JR /J 3.5) [70]. The phase diagram of the AFM spin ladder
in the presence of an external magnetic field is particularly interesting. In the
absence of the magnetic field and at T = 0, the ground state is a QSL with a
gap in the excitation spectrum. At a field Hc1 , there is a transition to a gapless
LL phase (gµB Hc1 =
SG , the spin gap, µB is the Bohr magneton and g the
Landé splitting factor). There is another transition at an upper critical field Hc2
to a fully polarized FM state. Both Hc1 and Hc2 are quantum critical points.
The quantum phase transition from one ground state to another is brought
about by changing the magnetic field. At small temperatures, the behaviour
of the system is determined by the crossover between two types of critical
behaviour: quantum critical behaviour at T = 0 and classical critical behaviour
at T = 0. Quantum effects are persistent in the crossover region at small finite
temperatures and such effects can be probed experimentally. In the case of the
ladder system (C5 H12 N)2 CuBr4 , the magnetization data obtained experimentally
exhibit universal scaling behaviour in the vicinity of the critical fields Hc1 and
Hc2 . In the gapless regime Hc1 < H < Hc2 , the ladder model can be mapped onto
an XXZ chain, the thermodynamic properties of which can be calculated exactly
Bose and Gayen [71] have studied a frustrated two-chain spin model with
diagonal couplings. The intra-chain and diagonal spin–spin interactions are of
equal strength J . It is easy to show that for JR ≥ 2J the exact ground state
consists of singlets (dimers) along the rungs with the energy Eg = −3JR N /4
where N is the number of rungs. Xian [72] later pointed out that as long as
JR /J > (JR /J )c 1.401, the rung dimer state is the exact ground state. At
JR /J = (JR /J )c , there is a first-order transition from the rung dimer state to the
Haldane phase of the S = 1 chain. Kolezhuk and Mikeska [73] have constructed
a class of generalized S = 12 two-chain ladder models for which the ground
state can be determined exactly. The Hamiltonian H is a sum over plaquette
Hamiltonians and each such Hamiltonian contains various two-spin as well as
four-spin interaction terms. They have further introduced a toy model which has
a rich phase diagram in which the phase boundaries can be determined exactly.
The standard spin ladder models with bilinear exchange are not integrable.
For integrability, multispin interaction terms have to be included in the
Hamiltonian. Some integrable ladder models have already been constructed [74].
We discuss one particular model proposed by Wang [75]. The Hamiltonian is
given by
J1
N
J2 N
H= ['
σj · σ'j +1 + τ'j · τ'j +1 ] + σ'j · τ'j
4 i=1 2 j =1
U1 N
+ σj · σ'j +1 )('
(' τj · τ'j +1 )
4 j =1
U2 N
+ σj · τ'j )('
(' σj +1 · τ'j +1 ) (7.32)
4 j =1
where σ'j and τ'j are the Pauli matrices associated with the site j of the upper
and lower chains respectively. N is the total number of rungs in the system. The
ordinary spin ladder Hamiltonian is obtained from equation (7.32) when the four
spin terms are absent, i.e. U1 = U2 = 0. For general parameters J1 , J2 , U1 and U2 ,
the model is non-integrable. The integrable cases correspond to U1 = J1 , U2 = 0
or U1 = J1 , U2 = −J1 /2. Without loss of generality, one can put J1 = U1 = 1,
1 N
H= (1 + σ'j · σ'j +1 )(1 + τ'j · τ'j +1 )
4 j =1
J N
1 1
+ σj · τ'j − 1) +
(' J− N. (7.33)
2 j =1 2 2
Three quantum phases are possible. For J > J+c = 2, the system exists in the
rung dimerized phase. The ground state is a product of singlet rungs. The SG
is given by
SG = 2(J − 2). For J+c > J > J−c , a gapless phase is obtained with
three branches of gapless excitations. J+c is the quantum critical point at which
a QPT from the dimerized phase to the gapless phase occurs. In the vicinity
of the quantum critical point, the susceptibility and the specific heat can be
calculated using the thermodynamic BA. From the low-temperature expansion
of the thermodynamic BA equation, one obtains
1 1
C ∼T 2 χ ∼ T −2 (7.34)
This is a gapless phase with two branches of gapless excitations. For U = −12 , a
similar phase diagram is obtained. Note that the ladder model may equivalently
be considered as a spin-orbital model with σ' and τ' representing the spin and the
pseudospin.
Doped ladder models are toy models of strongly correlated systems [65].
In these systems, the double occupancy of a site by two electrons, one with
spin up and the other with spin down, is prohibited due to strong Coulomb
correlations. In a doped spin system, there is a competition between two
processes: hole delocalization and exchange energy minimization. A hole moving
in an antiferromagnetically ordered spin background, say the Néel state, gives
rise to parallel spin pairs which raise the exchange interaction energy of the
system. The questions of interest are: whether a coherent motion of the holes
is possible; whether two holes can form a bound state; the development of
superconducting (SC) correlations; the possibility of phase separation of holes etc.
Some of these issues are of significant relevance in the context of doped cuprate
systems in which charge transport occurs through the motion of holes [76]. In the
SC phase, the holes form bound pairs with possibly d-wave symmetry. Several
proposals have been made so far on the origin of hole binding but there is, as
The C++ and C+iσ are the electron creation and annihilation operators which act in
iσ
the reduced Hilbert space (no double occupancy of sites),
++ = C + (1 − ni−σ )
Ciσ iσ
(7.36)
+iσ = Ciσ (1 − ni−σ )
C
where σ is the spin index and ni , nj are the occupation numbers of the ith and j th
sites respectively. The first term in equation (7.35) describes the motion of holes
with hopping integrals tR and t for motion along the rung and chain respectively.
In the standard t–J ladder model, i and j are nn sites. The second term contains
the usual AFM Heisenberg exchange interaction Hamiltonian. The t–J model,
thus, describes the motion of holes in a background of antiferromagnetically
interacting spins. A large number of studies have been carried out on t–J
ladder models. These are reviewed in [65, 66]. We describe briefly some of the
major results. The SG of the undoped ladder changes discontinuously on doping.
n. (7.37)
a
The two chains of the ladder are labelled by a = 1, 2 and µ is the chemical
(a)
potential coupling to the number of electrons in the system. Ht −J is the t–J
(a) (a) (a)
Hamiltonian (7.35) for a chain plus the terms nj + nj +1 where nj is the total
number of electrons on site j :
(1) (2)
Hint = − [Ht –J ]jj +1 [Ht –J ]jj +1 . (7.38)
j
Hrung includes the t–J Hamiltonian (7.35) corresponding to a rung and a Coulomb
interaction term V j n(1) (2)
j nj . The possible basis states of a rung are the
following. When no hole is present, a rung can be in a singlet or a triplet spin
configuration. When a single hole is present, the rung is in a bonding (|σ+ #)
or antibonding (|σ− #) state with |σ± # ≡ √1 (|σ 0# ± |0σ #) and σ =↑ or ↓. The
2
rung can further be occupied by two holes. Frahm and Kundu have studied the
phase diagram of the ladder model at low temperatures and in the strong coupling
regime JR * 1, V * µ + |tR | near half-filling. In this regime, the triplet states
are unfavourable. By excluding the triplet states and choosing J = 2t = 2, the
Hamiltonian H (7.37) can be rewritten as
5
H =− jj +1 − Al Nl + constant (7.39)
j l=1
Integrable models have a dual utility. They serve as testing grounds for
approximate methods and techniques. Also, they are often models of real systems
and provide rigorous information about the physical properties of such systems.
Integrable models are sometimes more general than what is necessary to describe
real systems. In such cases, an integrable model corresponds to an exactly solvable
point in the general phase diagram. The point may be a quantum critical point
at which transition from one quantum phase to another occurs or the integrable
model may be in the same phase as a more realistic model. In the latter case,
the physical properties of the two models are similar. In this review, we have
discussed the physical basis of some integrable spin models with special focus on
the relevance of the models to real systems. The Heisenberg spin chain is probably
the best example of the essential role played by exact solvability in correctly
interpreting the experimental data. The concept of spinons owes its origin to
the exact analysis of the BA equations. The theoretical prediction motivated
the search for real spin systems in which experimental confirmation could be
made. In this review, examples are also given of systems for which the links
between integrable models and experimental results are not well established. A
major portion of the review is devoted to physical systems which exhibit rich
phenomena, like the systems with both spin and orbital degrees of freedom and
undoped and doped spin ladder systems, where the need for integrable systems is
particularly strong. These systems exhibit a variety of novel phenomena, a proper
understanding of which should be based on rigorous theory. Two-dimensional
spin systems with QSL ground states have been specially mentioned to explain the
recent interest in constructing integrable models of such systems. The review is
meant to be an elementary introduction to the genesis and usefulness of integrable
models vis-à-vis physical spin systems. Future challenges are also highlighted to
motivate further research on integrable models.
There are some AFM spin models which are not integrable but for which the
ground states and, in some cases, the low-lying excited states are known exactly.
The most prominent amongst these are the Majumdar–Ghosh (MG) chain [87]
and the AKLT [54] model respectively. The MG Hamiltonian is defined in one
dimension for spins of magnitude 12 . The Hamiltonian includes both nn as well
References
[1] Bethe H 1931 Z. Phys. 71 205
See also Mattis D C (ed) 1993 The Many Body Problem: An Encyclopedia of Exactly
Solved Models in One Dimension (Singapore: World Scientific) for an English
translation of Bethe’s paper
[2] Takhtajan L A and Faddeev L D 1979 Russian Math. Surveys 34 11
Faddeev L D 1980 Sov. Sci. Rev. C 1 107
[3] Lieb E and Liniger W 1963 Phys. Rev. 130 1605
Lieb E 1963 Phys. Rev. 130 1616
Gaudin M 1967 Phys. Lett. A 24 55
Yang C N and Yang C P 1969 J. Math. Phys. 10 1115
[4] Lieb E and Wu F Y 1968 Phys. Rev. Lett. 20 1445
[5] Sutherland B 1975 Phys. Rev. Lett. 34 1083
Sutherland B 1975 Phys. Rev. Lett. 35 185
[6] Sutherland B 1975 Phys. Rev. B 12 3795
[7] Andrei N 1980 Phys. Rev. Lett. 45 379
[8] Wiegmann P B 1981 J. Phys. C: Solid State Phys. 14 1463
Filyov V M, Tsvelick A M and Wiegmann P G 1981 Phys. Lett. A 81 175
[9] Bares P A and Blatter G 1990 Phys. Rev. Lett. 64 2567
[10] González T, Martin-Delgado M A, Sierra G and Vozmediano A H 1995 Quantum
Electron Liquids and High-Tc Superconductivity (Berlin: Springer) ch 10
[11] Sutherland B 1985 Exactly Solvable Problems in Condensed Matter and Relativistic
Field Theory ed B S Shastry, S S Jha and V Singh (Berlin: Springer) p 1
[12] Gaudin M 1971 Phys. Rev. Lett. 26 1301
[13] Takhtajan L A 1985 Exactly Solvable Problems in Condensed Matter and Relativistic
Field Theory ed B S Shastry, S S Jha and V Singh (Berlin: Springer) p 175
[14] Bogoliubov N M, Izergin A G and Korepin V E 1985 Exactly Solvable Problems in
Condensed Matter and Relativistic Field Theory ed B S Shastry, S S Jha and V
Singh (Berlin: Springer) p 220
See also Korepin V E, Bogoliubov N M and Izergin A G 1993 QISM and Correlation
Functions (Cambridge: Cambridge University Press)
Australia
8.1 Introduction
The current realization of nanotechnology as a viable industry is presenting a
wealth of challenging problems in theoretical physics. Phenomena such as Bose–
Einstein condensation, entanglement and decoherence in the context of quantum
information, superconducting correlations in metallic nanograins, soft condensed
matter, the quantum Hall effect, nano-optics, the Kondo effect and Josephson
tunnelling phenomenon are all emerging to paint a vast canvas of interwoven
physical theories which provide hope and expectation that the emergence of new
nanotechnologies will be rapid in the short-term future. A significant tool in the
evolution of the theoretical aspects of these studies has been the development
and application of potent mathematical techniques, which are becoming ever
increasingly important as our understanding of the complexities of these physical
systems matures.
One approach that has recently been raised to prominence in this regard is
that of the exact solution of a physical model. The necessity of studying the exact
solution has been demonstrated through the experimental research on aluminium
grains with dimensions at the nanoscale level. The work of Ralph, Black and
Tinkham (RBT) [1] in 1996 detected the presence of superconducting pairing
correlations in metallic nanograins which manifest as a parity effect in the energy
spectrum dependent on whether the number of valence electrons on each grain
is even or odd. A naı̈ve approach to describe these systems theoretically is to
apply the theory of superconductivity due to Bardeen, Cooper and Schrieffer
(BCS) [2]. Indeed, the BCS model is the appropriate model for these systems
but the associated mean field treatment fails. This is because a mean field
First we will review the basic features of the quantum inverse scattering method
[14, 15]. The theory of exactly solvable quantum systems in this setting relies on
the existence of a solution R(u) ∈ End(V ⊗ V ), where V denotes a vector space,
which satisfies the Yang–Baxter equation acting on the three-fold tensor product
space V ⊗ V ⊗ V :
with b(u) = u/(u + η) and c(u) = η/(u + η). Here, P is the permutation
operator which satisfies
P (x ⊗ y) = y ⊗ x ∀ x, y ∈ V .
In this case, the Yang–Baxter algebra has four elements which we express as
A(u) B(u)
T (u) = . (8.4)
C(u) D(u)
it follows from (8.1) that the transfer matrices commute for different values of the
spectral parameters; viz.
= LU
2 (v)L2 (v)L1 (u)L1 (u)R12 (u − v)
W U W
a(vi ) M
vi − vj − η
= i = 1, . . . , M (8.14)
d(vi ) j =i vi − vj + η
are satisfied. Note that in the derivation of the Bethe ansatz equations, it is required
that vi = vj , ∀ i, j. This is a result of the Pauli principle for Bethe ansatz solvable
models as developed in [20] for the Bose gas. We will not reproduce the proofs
for the present cases, as they follow essentially the same argument as that in [20].
K κη2
µ EJ
= = −κηω = κ.
4 2 2 2
An explicit representation of (8.4) is obtained from (8.16) with the
identification
a(u) = (u + ω)(u − ω)
d(u) = η−2 .
with the eigenstates of the form (8.11) with C(u) given as before. From the Bethe
ansatz equations, we may derive the useful identity
m m N
vi − vj − η
η2 (vi2 − ω2 ) = (8.18)
i=1
v − vj + η
i=1 j =m+1 i
Note that this expression is independent of the spectral parameter u which can be
chosen arbitrarily. The formula simplifies considerably with the choice u = ω, by
employing (8.18), which yields a polynomial form:
N
η2 N 2
E = −κ η−2 η2 (vi − ω + η)(vi + ω) − − ηωN − η−2 .
i=1
4
However, for the purpose of an asymptotic analysis in the Rabi regime, it is more
convenient to choose u = 0, while for the Fock regime we use u = η2 .
Excitations correspond to changing the signs of the leading terms in the Bethe
ansatz roots. To study the asymptotic behaviour for the mth excited state, we set
vi ≈ −η−1 + i + ηδi i = 1, . . . , m
−1
(8.22)
vi ≈ η + i + ηδi i = m + 1, . . . , N
m
1
i = (8.23)
j =i
i − j
N
1
i = − (8.24)
j =m+1
i − j
j =i
which implies
N
N
(N − m)(N − m − 1)
i = 0 i2 = − .
i=m+1 i=m+1
2
It is clear from (8.23) and (8.24) why the Pauli exclusion principle applies in
the present case. In the asymptotic expansion for vi , i is assumed finite. However,
if vi = vj for some i, j , then i = j and (8.23) and (8.24) imply that i , j are
infinite which is a contradiction. Hence, vi must be distinct for different i. Note
also that for this approximation to be valid, we require η−1 * i . However, we
see that |i | is of the order of N 1/2 . Thus, our approximation will be valid for
ηN 1/2 1, which is precisely the criterion for the Rabi region and, consequently,
N cannot be arbitrarily large for fixed η or vice versa.
Now we go to the next order. From (8.18), we find
m
m(m − 1) m(m − N) mω2
δi = − + −
i=1
4 2 2
N
(N − m)(N − m − 1) m(m − N) (N − m)ω2
δi = − + +
i=m+1
4 2 2
Em η2 ω2 (N − 2m) η2 N η2
≈ −N + 2m − + + m(N − m).
κ 2 4 2
The energy level spacings
m = Em − Em−1 are, thus,
η2
m ≈ κ 2 + η ω + (N − 2m + 1) .
2 2
2
One may check that
m /N is of the order of N −1 . This indicates that the Rabi
regime is semiclassical [27]. This value for the gap between the ground and first
excited state agrees, to leading order in η2 N, with the Gross–Pitaevskii mean-field
theory [33] giving a Josephson plasma frequency of ωJ = 2κ(1 + η2 N/2)1/2 .
with
M +1 M(M + 1)
a1+ = − a1− =
2 √2
−(M − 1) ± (M − 1) 13M 2 + 10M + 1
2
a2± =
√12
(M − 1)(M − 2) 2M(M + 1)
a3± = ±
24
M −j +1
aj ± = √ aj −1,± j = 3, . . . , M
(j + 1)j (j − 1)(j − 2)
(M + 1)(2M + 1) M(M + 1)
a1+ = − a1− = −
2 √ 2
−(M − 1)2 ± i(M − 1) 11M 2 + 14M − 1
a2± =
√12
(M − 1)(M − 2) 2M(M + 1)
a3± = ±i
24
M −j +1
aj ± = √ aj −1,± j = 3, . . . , M
(j + 1)j (j − 1)(j − 2)
for the (symmetric) second excited state. The breakdown of the bound state at
−(M − j )η, j = 2, . . . , M results in the higher excited states.
Substituting these results into (8.20) leads us to the asymptotic ground-state
energy
E0 ≈ −2κη−2M(M + 1)
while for the first and second excited states, we have
M2 + M − 2
E1 ≈ κη2 − κη−2
3
5M 2 + 5M + 2
E2 ≈ κη2 + κη−2 .
3
In contrast to the Rabi regime, the Fock regime is not semiclassical, as the ratio
of the gap
and N is of finite order when N is large.
We can perform a similar analysis for odd N. In this case, the gap between
the ground and the first excited states is proportional to κη−2 instead of κη2 .
Furthermore, the ground-state root structure is different in the odd case since not
all the bosons can be bound in pairs. This indicates there is a strong parity effect
in the Fock regime, in contrast to the Rabi regime.
(ii) ω = 0. In this case the root structure is somewhat more complicated
than for ω = 0, so we will not present the details. We remark, however, that our
calculations show that up to order η−2 the ground-state energy eigenvalue takes
the same form as in the case ω = 0. Actually, the leading contribution arising from
the ω term appears only as ω2 η−4 . This means that the results presented here are
applicable for all values of ω (or, equivalently,
µ).
Although it is difficult to define rigorously [27, 34], the relative phase
between Bose–Einstein condensates is useful in understanding interference
experiments [28, 29, 35]. Recall that in Josephson’s original proposal [36]
for Cooper pair tunnelling through an insulating barrier between macroscopic
superconductors, the current is a manifestation of the relative phase between
the wavefunctions of the superconductors. By definition, the relative phase ! is
conjugate to the relative number of atoms in the two condensates n ≡ N1 − N2 .
and
t (0) = δK z + b † K − + bK + − ηK z Nb . (8.26)
Let |0# denote the Fock vacuum state and let |k# denote a lowest weight
state of the su(1, 1) algebra with weight k, i.e. K z |k# = k|k#. On the product state
|"# = |0#|k#, it is clear that B(u)|"# = 0 and
M
vi − η M
vi + η
(0) = k(δ + η−1 ) − kη−1 (8.27)
i=1
vi i=1
vi
(a †)2 a2 2Na + 1
K+ = K− = Kz =
2 2 4
we then find that the Hamiltonian (8.25) is related to (8.26) through
|k = (m + 1)/2# ≡ (a † )m |0#
and J = 2k − 1. A detailed analysis of this model through the exact solution will
be given at a later date.
For the exact solution of the Hamiltonian (8.25) it is necessary to take the
quasi-classical limit η → 0 in the Bethe ansatz equations (8.28). The resulting
Bethe ansatz equations take the form
2k M
1
δ − vi + =2 . (8.30)
vi v − vi
j =i j
M
E = ω(M + k − 1/4) − vi
i=1
M
1
= ω(k − 1/4) − 2k . (8.31)
v
i=1 i
The equivalence of the two energy expressions can be deduced from (8.30). The
eigenstates too are obtained by this procedure. Consider the following class of
where c(v) = (vb † − a † a † /2), |"# = |0# for k = 1/4 and |"# = a †|0# for k =
3/4. In the case when the set of parameters {vi } satisfy the Bethe ansatz equations
(8.30), then (8.32) are precisely the eigenstates of the Hamiltonian.
For i > m, we obtain, from the zero-order terms in the Bethe ansatz equations,
M
1
i = 2
j =m+1
i − j
j =i
which implies
M
i = 0.
i=m+1
and, thus,
M
µi = 2(k + m)(M − m).
i=m+1
Next we look at the Bethe ansatz equations for i ≤ m. The terms in δ give
2k m
1
1+ =2
µi µ
j =i j
− µi
which implies
m
µi = −2km − m(m − 1).
i=1
M
2
M
Em ≈ ω(M + (k − 1/4)) − ω(M − m) − i − µi
i=m+1
ω i=1
2
= ω(m + k − 1/4) + (3m2 − m + 4km − 2kM − 2mM).
ω
The level spacings are
m = Em − Em−1
22
≈ω− (M + 2 − 3m − 2k)
ω
from which we conclude that, in this limit, the model is semi-classical.
Let E denote the ground-state energy (E = E0 for δ * 0, E = EM for
δ 0) and
the gap to the first excited state. Employing the Hellmann–
Feynman theorem, we can determine the asymptotic form of the following zero-
temperature correlations
∂E ∂E
"Na # = 2 θ = −2
∂ω ∂
where θ = −"a † a † b + b†aa# is the coherence correlator. For large N, we
introduce the rescaled variables
δ
"Na # θ
δ∗ =
∗ = "Na #∗ = θ∗ = . (8.33)
N 1/2 N 1/2 N N 3/2
We then have, for δ ∗ * 0,
1
∗ ≈ δ ∗ − "Na #∗ ≈ 0 θ∗ ≈ 0
δ∗
while, for δ ∗ 0,
2 1 1
∗ ≈ −δ ∗ − "Na #∗ ≈ 1 − θ∗ ≈ − .
δ∗ 2(δ ∗ )2 δ∗
This shows that the model has scale invariance in the asymptotic limit. The
scaling properties actually hold for a wide range of values of the scaled detuning
parameter δ ∗ , which is established through numerical analysis [43].
uG
− (u2 − δu − 2k)G
+ (Mu − E/ + δ(k − 1/4))G = 0 (8.34)
F (u) = uG
− (u2 − δu − 2k)G
.
The recurrence relation (8.35) can be solved as follows (cf [19]). Setting
n
gn+1 = g0 xj yj
j =0
with
E − ω(j + k − 1/4)
xj =
(j + 1)(j + 2k)
and substituting into the recurrence relation (8.35), we have
j −M −1
xj xj −1 yj −1 (yj − 1) = .
(j + 1)(j + 2k)
2 (j + 1)(j + 2k)(j − M)
cj =
(E − ω(j + k + 3/4))(E − ω(j + k − 1/4))
which means yj can be expressed as a continued fraction. The requirement that
G is a polynomial function of order M decrees yM = 0, in turn implying
2 M(M + 2k − 1)
yM−1 =
(E − ω(M + k − 1/4))(E − ω(n + k − 5/4))
which is an algebraic equation that determines the allowed energy levels Em . This
procedure can easily be employed to determine the energy spectrum numerically,
without resorting to solving the Bethe ansatz equations. Explicit results can be
found in [43].
L
θj k
τj = 2αSjz + (8.38)
k=j
j − k
with θ = S + ⊗ S − + S − ⊗ S + + 2S z ⊗ S z .
We define a Hamiltonian through
L L L L
1 1 1 1
H =− j τj + 3 τj τk + 2 τj − Cj (8.39)
α j =1 4α j,k=1 2α j =1 2α j =1
L
L
1
=− 2j Sjz − S−S+ (8.40)
j =1
α j,k=1 j k
where
C = S + S − + S − S + + 2(S z )2
is the Casimir invariant for the su(2) algebra. The Hamiltonian is universally
integrable since it is clear that [H, τj ] = 0, ∀j irrespective of the realizations of
the su(2) algebra in the tensor algebra.
In order to reproduce the Hamiltonian (8.36), we realize the su(2) generators
through the hard-core boson (spin- 12 ) representation, i.e.
In this instance, one obtains (8.36) (with the constant term − L j j ) where
g = 1/α as shown by Zhou et al [23] and von Delft and Poghossian [46].
For each index k in the tensor algebra in which the transfer matrix acts, and
accordingly in (8.40), suppose that we represent the su(2) algebra through the
irreducible representation with spin sk . Thus {Sk+ , Sk− , Skz } act on a (2sk + 1)-
dimensional space. In employing the method of the algebraic Bethe ansatz
discussed earlier, we find that
L
u − k − ηsk
a(u) = exp(−αη)
k=1
u − k
L
u − k + ηsk
d(u) = exp(αη)
k=1
u − k
such that the parameters vj now satisfy the Bethe ansatz equations
L
M
2sk 2
2α + = . (8.43)
k=1
vj − k i=j vj − vi
Through (8.42) we can now determine the energy eigenvalues of (8.40). It is useful
to note the following identities:
M M L
vj sk
2α vj + 2 = M(M − 1)
j =1
v
j =1 k=1 j
− k
L
M
sk
αM + =0
j =1 k=1
vj − k
M L
M L L
vj sk sk k
− =M sk .
v − k j =1 k=1 vj − k
j =1 k=1 j k=1
M
E=2 vj . (8.44)
j =1
From this expression, we see that the quasi-particle excitation energies are given
by twice the Bethe ansatz roots {vj } of (8.43). In order to specialize this result to
the case of the BCS Hamiltonian (8.36), it is a matter of setting sk = 1/2, ∀ k.
Finally, let us remark that in the quasi-classical limit the eigenstates assume the
form
M L bj†
|"# = |0#.
v − j
i=1 j =1 i
The construction given here can also be applied on a more general level.
Taking higher spin representations of the su(2) algebra produces models of BCS
systems which are coupled by Josephson tunnelling, as described in [47, 48].
One can also employ higher-rank Lie algebras, such as so(5) [49] and su(4)
[50,51] which produce coupled BCS systems which model pairing interactions in
nuclear systems. For the general case of an arbitrary Lie algebra, we refer to [52].
Finally, let us mention that if one reproduces this construction with the su(1, 1)
L-operator (8.9) in place of the su(2) L-operator (8.8), the pairing model for
bosonic systems introduced by Dukelsky and Schuck [53] is obtained.
vi ≈ i + gδi + g 2 µi i = 1, . . . , M.
Substituting this into (8.43) and equating the different orders in g yields
L M
g g2 1 1
vi ≈ i − + −
2 4 k=m+1 j − k i=j j − i
L
M
g2
M 1
E0 ≈ 2 j − gM + .
j =1
2 j =1 k=M+1 j − k
Next we look at the first excited state. In the g = 0 case, this corresponds to
breaking the Cooper pair at level M and putting single unpaired electrons in the
M−1 L
g 2 M−1 1
E1 ≈ M + M+1 + 2 j − g(M − 1) + .
j =1
2 j =1 k=M+2 j − k
g2
M
1
"ni # ≈ .
2 j =1 (j − i )2
Acknowledgments
We are deeply indebted to Ross McKenzie, Mark Gould, Xi-Wen Guan and
Katrina Hibberd for their collaborations on these topics. Financial support
from the Conselho Nacional de Desenvolvimento Cientı́fico e Tecnológico and
Australian Research Council is gratefully accepted.
9.1 Introduction
Much effort has been devoted to the study of integrable quantum spin chains
such as the Heisenberg model [1, 2], t–J [3–5] and Hubbard models [6, 7]
and many more. The appealing feature of these systems is the availability of
exact data for the spectrum and other physical properties despite the truly
interacting nature of the spins (resp. particles). The computational basis for the
work on integrable quantum chains is the Bethe ansatz yielding a set of coupled
nonlinear equations for one-particle wavenumbers (Bethe ansatz roots). Many
studies of the Bethe ansatz equations have been directed at the ground state of
the considered system and have revealed interesting non-Fermi liquid properties
such as algebraically decaying correlation functions with non-integer exponents
and low-lying excitations of different types, i.e. spin and charge with different
velocities constituting so-called spin and charge separation, see [8–10].
A very curious situation arises in the context of the calculation of the
partition function from the spectrum of the Hamiltonian. Despite the validity of
the Bethe ansatz equations for all energy eigenvalues of the previously mentioned
models, the direct evaluation of the partition function is rather difficult. In contrast
to ideal quantum gases, the eigenstates are not explicitly known, the Bethe ansatz
equations just provide implicit descriptions that pose problems of their own kind.
Yet, knowing the behaviour of quantum chains at finite temperature is important
for many reasons. As a matter of fact, the strict ground state is inaccessible
due to the very fundamentals of thermodynamics. Therefore, the study of finite
L
HL = S'j S'j +1 (9.1)
j =1
With these settings the partition function ZL of the quantum chain at finite
temperature T reads as
The rhs of this equation can be interpreted as the partition function of a staggered
six-vertex model with alternating rows corresponding to the transfer matrices
T (τ ) and T (τ ), see figure 9.1. We are free to evaluate the partition function of this
classical model by adopting a different choice of transfer direction. A particularly
useful choice is based on the transfer direction along the chain and corresponding
transfer matrix T QTM defined for the columns of the lattice yielding
In the remainder of this chapter, we will refer to T QTM as the ‘quantum transfer
matrix’ of the quantum spin chain because T QTM is the closest analogue to the
transfer matrix of a classical spin chain. Due to this analogy, the free energy f
per lattice site is given just by the largest eigenvalue max of the QTM
Note that the eigenvalue depends on the argument τ = β/N which vanishes in the
limit N → ∞ requiring a sophisticated treatment.
The main difference to classical spin chains is the infinite dimensionality of
the space in which T QTM is living (for N → ∞). In formulating (9.7) we have
implicitly employed the interchangeability of the two limits (L, N → ∞) and the
existence of a gap between the largest and the next-largest eigenvalues of T QTM
for finite temperature [31, 32].
N
−τ
chain length L
Figure 9.1. Illustration of the two-dimensional classical model onto which the quantum
chain at finite temperature is mapped. The square lattice has width L identical to the chain
length and height identical to the Trotter number N. The alternating rows of the lattice
correspond to the transfer matrices T (τ ) and T (τ ), τ = β/N. The column-to-column
transfer matrix T QTM (quantum transfer matrix) is of particular importance to the
treatment of the thermodynamic limit. The arrows placed on the bonds indicate the type of
local Boltzmann weights, i.e. R- and R-matrices alternating from row to row. (The arrows
do not denote local dynamical degrees of freedom.)
Note that we are mostly interested in which is obtained from (v) simply by
setting v = 0. Nevertheless, we are led to the study of the full v-dependence since
the condition fixing the values of vj is the analyticity of λ1 (v) + λ2 (v) in the
complex plane. This yields
a(vj ) = −1 (9.15)
where the function a(v) is defined by
λ1 (v) φ(v − i)q(v + 2i)
a(v) = = eβh . (9.16)
λ2 (v) φ(v + i)q(v − 2i)
Figure 9.3. Distribution of zeros (◦) and poles (×) of the auxiliary function a(v). All zeros
and poles vj ∓ 2i are of first order, the zeros and poles at ±(2i − iτ ), ±iτ are of order N/2.
Figure 9.4. Distribution of zeros and poles of the auxiliary function A(v) = 1 + a(v). Note
that the positions of zeros (◦) and poles (×) are directly related to those occurring in the
function a(v). There are additional zeros () above and below the real axis. The closed
contour L, by definition, surrounds the real axis and the zeros (◦) as well as the pole at
−iτ .
these zeros are depicted in figure 9.4 (open squares) but for a while they are not of
prime interest to our reasoning. Next we are going to formulate a linear integral
expression for the function log a(v) in terms of log A(v). To this end, we consider
the function
1 1
f (v) := log A(w) dw (9.18)
2πi L v − w
defined by an integral with closed contour L surrounding the real axis, the
parameters vj and the point −iτ in anticlockwise manner, see figure 9.4. Note
that the number of zeros of A(v) surrounded by this contour is N/2 and, hence,
identical to the order of the pole at −iτ . Therefore, the integrand log A(w) does
not show any non-zero winding number on the contour and consequentially the
integral is well defined. By use of standard theorems, we see that the function
f (v) is analytic in the complex plane away from the real axis with asymptotic
behaviour equal to zero. Due to Cauchy’s theorem, there is a jump discontinuity at
the real axis identical to the discontinuity of log A(v) itself. This discontinuity, in
turn, is identical to that of the function log[q(v)/(v + iτ )N/2 ] whose asymptotic
behaviour is equal to 0. We, therefore, have the identity
q(v)
f (v) = log (9.19)
(v + iτ )N/2
which is proved by noting the three properties of the difference function of the
left- and right-hand sides: (i) analyticity on the complex plane with a possible
exception at the real axis, (ii) continuity at the real axis (hence we have analyticity
everywhere) and (iii) zero asymptotics (from this follows boundedness—due to
Liouville’s theorem this bounded function is constant and, of course, equal to
zero).
Thanks to (9.18) and (9.19), we have a linear integral representation of
log q(v) in terms of log A(v). Because of (9.16) the function log a(v) is a linear
as the only singularities of the integrand surrounded by the contour L are the
simple zeros vj and the pole −iτ of order N/2 of the function A. Also, we obtain
1
1 1 1 N/2
[log A(w)]
dw = − +
2πi L v − w j
v − vj − 2i l
v − wl v + 2i − iτ
(9.25)
where the evaluation of the integral is done by use of the singularities outside
of the contour L. We deform the contour such that the upper (lower) part of L
is closed into the upper (lower) half-plane where the relevant singularities are
the simple poles vj + 2i, the zeros wl and the pole iτ − 2i of order N/2 of the
function A.
Next, we take the difference of (9.24) and (9.25), perform an integration by
parts with respect to w and finally integrate with respect to v:
1 1 1
− log A(w) dw
2πi L v − w v − w − 2i
[(v − i(2 − τ ))(v + i(2 − τ ))]N/2
= log ( + constant. (9.26)
l (v − wl )
The constant is determined from the asymptotic behaviour of the lhs for v →
∞ with the result constant = − log A(∞) = − log(1 + exp(βh)). Combining
(9.23), (9.26) and (9.11), we find that
1 log A(w)
log (v) = −βh/2 + dw. (9.27)
π L (v − w)(v − w − 2i)
These formulas, (9.27) and (9.22), are the basis of an efficient analytical and
numerical treatment of the thermodynamics of the Heisenberg chain. There are,
however, variants of these integral equations that are somewhat more convenient
for the analysis, especially for magnetic fields close to zero [37].
C
0.6 0.1
0.0
0.0 0.5 1.0 1.5 2.0
C/T
0.4 T
0.2
0.0
0.0 0.5 1.0 1.5 2.0
T
Figure 9.5. Specific heat coefficient c(T )/T versus temperature T for the spin- 12 XXX
chain. In the inset the specific heat c(T ) versus T is shown.
0.106 0.14
0.12
0.104
χ
0.10
0.102
0.0 0.5 1.0 1.5 2.0
0 0.001 0.002 0.003 0.004 0.005
T
Figure 9.6. Magnetic susceptibility χ at low temperature T for the spin- 12 XXX chain. In
the inset χ(T ) is shown on a larger temperature scale.
2 1 2
K(x) := − 0 (x) := −h − . (9.29)
π x2 + 4 x2 +1
In a similar manner, we obtain for the eigenvalue
∞
1
log = −βh/2 + ρ0 (x) log A(x) dx ρ0 (x) = . (9.30)
−∞ π(x 2 + 1)
x0 x0
Re(v)
Figure 9.7. Depiction of the root (•) and hole (◦) distribution at low temperature (and
h < 0). Note that both distributions touch the axis Im(v) = −1 at the points with real parts
±x0 . For purposes of illustration, the picture has been turned upside down (imaginary axis
is directed downwards) in order to have the holes above the roots.
so-called 1-strings, 2-strings etc but at low temperature we simply have particle–
hole-type excitations in very much the same way as in (free) Fermi systems.
Hence, excitations are constructed by changing roots into holes and vice versa.
The points ±x0 play the role of Fermi points and the number of roots is related
to the spin of the eigenstate of the QTM. The integral equations derived earlier
for the largest eigenvalue are still valid in the general case only if the integration
contour is designed such that it separates the roots from the holes.
:= 0 + K ∗ ξ := 1 + K ∗ ξ. (9.32)
From the last equation and the definition of the Fermi points ±x0 , i.e. (x0 ) = 0,
we have log a(x0 ) = [ξ(x0 ) − 1]2πid or
with an odd step function σ taking values −1, 0, +1 for x < −x0 , |x| < x0 ,
x > x0 , where ◦ denotes the convolution of two functions a ◦ b(x) = I a(x −
y)b(y) dy with integration interval I = ]−∞, −x0 ] ∪ [x0 , ∞[. Defining the
root
hole
η=K ◦σ +K ∗η (9.36)
η
(x) = K(x − x0 )(1 − η(x0 )) + K(x + x0 )(1 + η(−x0 )) + K ∗ η
(x)
= f0
+ K ∗ η
(x) (9.38)
with
f0
= [K(x − x0 ) + K(x + x0 )][1 − η(x0 )] (9.39)
where we have used η(−x) = −η(x), i.e. η is odd. Using this property once again,
we find
2η(x0 ) = η
(x) dx = 1 · η
= ξ · f0
(9.40)
|x|<x0 |x|<x0 |x|<x0
where we have employed the ‘dressed function trick’, i.e. ab0 = a0 b.
Substituting (9.39) into the last term of (9.40) and observing the definition
of ξ in (9.32), we obtain
and, hence,
η(x0 ) = [1 − ξ −1 (x0 )]. (9.42)
where we have again ignored the contribution of the σ term in the integral as it is
an odd function. With the definition of the dressed density ρ satisfying
ρ = ρ0 + K ∗ ρ (9.52)
and using the ‘dressed function trick’, ab0 = a0 b, we obtain, with (9.50),
x0
ρ0 (x)[log a(x) + 2πid] dx
−x0
x0 x0
= −β ρ(x)0 (x) dx + 2πid ρ(x) · 1 dx
−x0 −x0
x0
1
+
ρ(x)[+ K(x − x0 ) + − K(x + x0 )] (9.53)
β (x0 ) −x0
& '$ %
= + [ρ(x0 ) − ρ0 (x0 )] + − [ρ(−x0 ) − ρ0 (−x0 )]
where we have again ignored the contribution of the σ term in (9.50) as it is an
odd function. Combining the last equation with (9.51), we obtain
x0
log = −βh/2 − β ρ(x)0 (x) dx
−x
& 0 '$ %
=: E
x0
ρ(x0 )
+ 2πid ρ(x) dx +(+ + − )
. (9.54)
−x0 β (x0 )
& '$ %
=: ρ
v =
/2πρ|x0 . (9.55)
Acknowledgments
The authors acknowledge financial support by the Deutsche Forschungsgemein-
schaft under grants Kl 645/3-3, 645/4-1 and the Schwerpunktprogramm SP1073.
Re a
1 x
Appendix
Here we want to sketch the calculation of integrals of type (9.46)
x
log
log(1 + ez ) dz (9A.1)
−∞
-
with the contour along the real axis and at the end encircling a certain number n
of odd multiples of πi. We substitute z = log a and obtain the integral
x
log(1 + a)
In (x) := da (9A.2)
a
0
now with a contour starting at the origin and (A) surrounding the origin n times
in narrow loops in a clockwise manner, (B) followed by n larger loops in a
counterclockwise manner around the origin as well as −1 finally ending at x.
Obviously the first part (A) of the contour can be dropped. The simplest case with
n = 1 is illustrated in figure 9A.1.
We want to express In in terms of I0 . The explicit relation we are going to
prove is
In (x) = I0 (x) − 2π 2 n2 + 2πin log x. (9A.3)
The essential ingredient of our computation is the analytic continuation of I0 (x)
in x surrounding 0 and −1 exactly one time in a counterclockwise manner, see
figure 9A.1, giving I1 (x):
x −1 x
log(1 + a) log(1 + a) log(1 + a) + 2πi
I1 (x) = da = da + da
a 0 a −1 a
0 & '$ %
& '$ % ‘straight’
one loop
x
2πi
= I0 (x) + da = I0 (x) − 2π 2 + 2πi log x. (9A.4)
−1 a
0
∞ zn
∞ 0
ne ezn
log(1 + e ) dz =
z
(−1) dz = (−1) n
dz
n=1
n n=1
n
−∞ −∞
∞
1 π2
= (−1)n = (9A.6)
n=1
n2 12
References
[1] Takahashi M 1971 Phys. Lett. A 36 325–6
[2] Gaudin M 1971 Phys. Rev. Lett. 26 1301–4
[3] Schlottmann P 1992 J. Phys. Cond. Mat. 4 7565–78
[4] Jüttner G, Klümper A and Suzuki J 1997 Nucl. Phys. B 486 650
[5] Suga S and Okiji A 1997 Physica B 237 81–3
[6] Kawakami N, Usuki T and Okiji A 1989 Phys. Lett. A 137 287–90
[7] Jüttner G, Klümper A and Suzuki J 1998 Nucl. Phys. B 522 471
[8] Schulz H J 1993 Correlated Electron Systems vol 9 (Singapore: World Scientific)
p 199
[9] Eßler F H L and Korepin V E 1994 Exactly Solvable Models of Strongly Correlated
Electrons (Singapore: World Scientific)
[10] Lieb E H 1995 Advances in Dynamical Systems and Quantum Physics (Singapore:
World Scientific) p 173
[11] Belavin A A, Polyakov A M and Zamolodchikov A B 1984 Nucl. Phys. B 241 333
[12] Bethe H 1931 Z. Phys. 71 205–26
[13] Sutherland B 1970 J. Math. Phys. 11 3183–6
[14] Baxter R J 1971 Phys. Rev. Lett. 26 834
The first of the allowed movements in (10.1) is reversible while the second is
not. A typical question is then for the long-time behaviour of such quantities
as the mean particle density n(t). Trivially n(t) decreases with increasing time
t but different long-time asymptotic behaviours such as n(t) ∼ t −y or n(t) ∼
e−t /τ are conceivable. The oldest approach to this problem was introduced by
Smoluchowski [11] and consists of writing down kinetic equations, e.g. for the
spatially averaged density n(t) and one obtains ∂t n(t) = −λn(t)2 for the problem
at hand, where λ = 4α. With the initial condition n(0) = n0 , the solution n(t) =
n0 (1 + n0 λt)−1 (λt)−1 is easily found and apparently answers the physical
question. A slightly more involved version of this argument does allow for spatial
variation of the density n = n(r, t) and considers a reaction–diffusion equation
the long-time behaviour is still described by the model (10.1), with a renormalized
rate. For late times, one expects the mean exciton density to fall off as a power
law n̄(t) ∼ t −y . The excitons are unstable, with lifetimes of the order 10−3 s.
Their decay produces light of a characteristic frequency whose intensity can be
used to measure n̄(t) while light with a different frequency is emitted if excitons
decay through a pair reaction. This allows one to measure ∂t n̄(t) as well, through
time-resolved experiments down to the picosecond scale. Table 10.1 gives some
results for the exponent y in different materials (the branching ratio B 10% in
the first two lines of table 10.1). Clearly y 1/2 as expected from (10.6) and far
from unity. This is strong evidence in favour of strong fluctuation effects in these
systems and against their description through a reaction–diffusion equation (10.2).
Another aspect becomes apparent if we now briefly consider the triple
annihilation process • • • → ◦ ◦ ◦ combined with single-particle diffusion. The
where w(τ → σ ) are the transition rates between the configurations τ and σ and
are assumed to be given from the phenomenology. In order to rewrite this as a
matrix problem, one introduces a state vector
|P # = P (σ ; t)|σ # (10.11)
σ
1 This does not imply, however, that models containing both binary and multi-site reaction terms
could not have a non-trivial behaviour. For example, the phase structure of the pair contact process
(•• → ◦◦, • • ◦ → • • •) with single-particle diffusion (•◦ ↔ ◦•) is presently controversial and under
very active study [19].
∂t |P # = −H |P # (10.12)
Re En ≥ E0 = 0. (10.16)
(iii) Let |P0 # = |P (0)# be the initial state such that the weights of the individual
configurations satisfy 0 ≤ P (σ ; 0) ≤ 1 and "s|P0 # = 1. Then, for all times
t ≥ 0, one has
and one-point and two-point functions of the nj are then expressed as2
C1 (j ; t) = "ñj #(t) = "s|ñj |P # C2 (j, ; t) = "ñj ñ #(t) = "s|ñj ñ |P #.
(10.21)
Two basic situations are readily distinguished from the spectrum of H . If in
the limit of infinite lattice size L → ∞, the lowest excited states have a finite gap
to the ground-state energy E0 = 0, then the averages (10.21) will approach their
steady-state values exponentially fast on the time-scale τ = 1/ . However, if
there is, in the L → ∞ limit, a continuous spectrum down to E0 = 0, one expects
an algebraic decay of the correlators as the system approaches the steady state.
hold. However, the quotient (1, 1)Hn (q) is defined through the condition [23]
(ei ei+2 )ei+1 (q + q −1 − ei )(q + q −1 − ei+2 ) = 0 i = 1, 2, . . . . (10.25)
1 L L
Sz = σiz S± = Si±
2 i=1 i=1
(10.32)
ln q
i−1
ln q L
Si± = exp σz σi± exp − σz
2 =0 2 =i+1
q 2S − q −2S
z z
[S z , S ± ] = ±S ± [S + , S − ] = . (10.33)
q − q −1
However, set
where
L
iπ
i−1
T ±
=q (1−L)/2
q i−1
exp (σ + 1) σi±
z
(10.36)
i=1
2 =1
and
q L − q −L
[S z , T ± ] = ±T ± {T + , T − } = . (10.37)
q − q −1
D L−1 y y
H =− [σ x σ x + σj σj +1 + (σjz σjz+1 − 1)] (10.38)
2 j =1 j j +1
and coincides with the (ferromagnetic) XXX Heisenberg quantum chain [30].
Certainly, one may now use the Bethe ansatz solution of HXXX to rederive
known results on simple diffusion. The recent interest in this set-up comes from
the insight that the integrability of the associated quantum chains allows us to
make contact with the pre-established algebraic techniques for the treatment of
these [26, 31]. Independently, integrability was also observed to occur in the
transfer matrices for discrete-time dynamics [32, 33]. The enormous possibilities
for non-trivial applications then triggered an ongoing wave of activity, see, e.g.,
[2, 4, 10, 26, 34–39] and references therein.
Following [26, 31], we now give more examples of integrable quantum
Hamiltonians of stochastic systems, restricting ourselves for simplicity to a single
species of particles and to binary reactions only (see section 10.1). The reaction
rates are defined in table 10.2, using the convention of various authors but
unfortunately there is no standard notation. While we prefer a light notation
(slightly modified from [40]) and shall use it here3 , other authors often opt for
a systematic though heavier notation with several indices.
For the time being and for purposes of illustration, let us consider besides
diffusion only those reactions which irreversibly reduce the number of particles
(that is βL,R = νL,R = σ = 0). Define
√ √ ) )
γL γR δL δR DL γL δL
D = DL DR , γ = , δ= , q= = =
D D DR γR δR
(10.39)
= 12 (q + q −1 )(1 + δ − γ ) − α/D, h = 12 (2α/D + γ (q + q −1 )). (10.40)
Note that the ratio of the left and right rates is taken to be the same for diffusion,
coagulation and death processes. We first consider an open chain with L sites.
Then the quantum Hamiltonian becomes
where HXXZ (h,
, δ) is the standard XXZ quantum chain, including bulk and
boundary magnetic fields
1 L−1 y y
HXXZ (h,
, δ) = − σ x σ x + σj σj +1 +
(σjz σjz+1 − 1)
2 j =1 j j +1
1
+ h(σjz + σjz+1 − 2) − (1 − δ)(q − q −1 )(σjz − σjz+1 )
2
(10.42)
which contains the diagonal and diffusive matrix elements while the particle
annihilation terms are contained in
L−1
Hα = −2α q −2j −1 σj+ σj++1
j =1
L−1
Hγ = −γ q −j (ñj σj++1 + q −1 σj+ ñj +1 ) (10.43)
j =1
L−1
Hδ = −δ q −j (q −2 (1 − ñj )σj++1 + qσj+ (1 − ñj +1 ))
j =1
To see this, recall that the XXZ Hamiltonian conserves the number of
particles while the reaction terms irreversibly decrease the total particle number.
Thus, H can be written in a block diagonal form
N0 Xδ Xα
N1 Xγ ,δ Xα
H = . .
(10.45)
N2 Xγ ,δ .
.. ..
. .
where Nn refers to the n-particle states and X are the reaction matrix elements.
Because of the identity
A X
det = det A det B (10.46)
0 B
it follows that the elements of (10.43) do not enter into the characteristic
polynomial of H .
Therefore, the phase diagram for the full Hamiltonian H can be read off
from the well-known spectrum of HXXZ (h,
, δ) [44]. For our purposes, we need
the following [31]. From (10.40), only the portion of the phase diagram where
where the Ij are pairwise distinct integers (half-integers) when r is odd (even)
and
sin((k − k
)/2)
(k, k
) = 2 arctan . (10.53)
cos((k + k
)/2) −
cos((k − k
)/2)
We are interested in the leading finite-size corrections when L → ∞ with r fixed.
The ansatz
2π aj
kj = Ij + 2 + · · · (10.54)
L L
and
2π
r
aj = (I − Ij ).
1 −
=1
Then (10.49) follows. Second, for free boundary conditions, the Bethe ansatz [48]
reproduces equation (10.51) for the energies, while the Bethe ansatz equations for
the quasi-momenta now take the form
1
Lkj = πIj − [(kj , k ) − (−kj , k )] (10.55)
2 =j
L−1
H = −D ei (10.56)
i=1
sinh θ sinh(η − θ )
Ři (θ ) = ei + q = eη (10.58)
sinh η sinh η
Ři (θ )Ři±1 (θ + θ
)Ři (θ
) = Ři±1 (θ
)Ři (θ + θ
)Ři±1 (θ )
[Ři (θ ), Řj (θ
)] = 0 |i − j | ≥ 2 (10.61)
k,n
Ři (θ ) = S,m 11 ⊗ · · · ⊗ 1i−1 ⊗ E mk ⊗ E n ⊗ 1i+2 ⊗ · · · . (10.62)
• ◦ 6
6. @@ eη sinh θ
@ 6
@
◦ •
◦ ◦
@ -?
7. @ sinh θ
@@ 6
• •
Figure 10.2. Diffusion and pair-annihilation of particles in the seven-vertex model and
their Boltzmann weights, for the vertices 5 to 7.
This implies that the vertex model associated with equation (10.56) is a seven-
vertex model. In a vertex model, arrows are attached to the bonds of a square
lattice [50]. In the stochastic model, we associate a particle (•) with an arrow
pointing up/right and no particle (◦) with an arrow pointing down/left. In
figure 10.2 we list together the chemical reactions, the vertex configurations and
their Boltzmann weights. The vertices usually labelled 1 to 4 correspond to no
reaction and are not shown (see [31]). Vertices 5 and 6 correspond to diffusion to
the right and to the left and vertex 7 to pair annihilation. In the leftmost column
of figure 10.2, the state of the particles before the reaction is given as the lower
pair of symbols while the state after the reaction is given by the upper pair of
symbols. The middle column gives the corresponding vertex configuration and
the right column the Boltzmann weight. The Hamiltonian equation (10.56) may
be recovered from H = − dθ d
ln T (θ )|θ=0 .
We have studied in some detail the pair-annihilation process and its integrability.
Still, extracting explicitly information about the long-time behaviour (or the
L
H= 11 ⊗ · · · ⊗ 1j −1 ⊗ Hj,j +1 ⊗ 1j +2 · · · ⊗ 1L . (10.63)
j =1
One can always rescale time such that D = 1. Then the parameters of the XXZ
chain H0 = HXXZ become
= 1 + δ − γ − α and h = α + γ .
In equations (10.63) and (10.64), H is only integrable for δ = 0.
However, the special case α = 0, γ = δ simply corresponds to the radioactive
decay of diffusively moving particles. While for δ = 0, one is back to the
critical Pokrovsky–Talapov line h +
= 1, the associated quantum spin chain
ground state for δ = 0.
has a frozen The one- and two-particle correlators
C1 (t) = L j =1 C1 (j ; t) and C2;n (t) =
L
j =1 C(j, j + n; t) with n fixed, see
equation (10.21), only imply the sectors with r = 1 and r = 2 particles of HXXZ,
respectively. Their long-time behaviour is easily worked out from the results of
section 10.4 and is collected in table 10.3 [40]. Of course, we implicitly assume
that the corresponding amplitudes do not accidentally vanish. At first sight, one
might have expected a simple exponential factor e−2kδt for the k-point correlator
Ck and we already observe that eventual algebraic prefactors are not readily
predicted from the spectrum of H alone. The more complicated form of the
relaxation time for δ > α + γ comes from a bound state in the two-particle sector
of HXXZ, with energy 4δ + 4 − 2
− 2/
, see [50, 51]. More general initial
conditions are discussed in [40].
This also explains the experimental results in table 10.1. The known stochastic
similarity transformations of the form (10.65) leave the parameters
and h of the
XXZ chain invariant, but the results of proposition 10.2 suggest that a stochastic
similarity transformation between systems with different values of
might exist.
See [53] for the extensions to δ > 0 and q = 1.
Equation (10.67) allows long-time behaviour C(t|α, γ ) ∼ t −1/2 to be
recovered from a simple heuristic argument. For pure coagulation, one particle
always remains, thus C(∞|0, γ ) = 1/L in the steady state. Therefore, the steady-
state density C(∞|α, γ ) ∼ L−1 . However, from the spectrum of H , the leading
relaxation time τ = −1 ∼ L2 ∼ ξ 2 , where ξ is identified as the characteristic
spatial length scale. Therefore, C(∞|α, γ ) ∼ ξ −1 ∼ τ −1/2 . The asserted time-
dependent behaviour, therefore, might have been anticipated on dimensional
grounds.
For the pure annihilation model with 2α = DR + DL , one has
= 0. In this case,
H may be diagonalized through a Jordan–Wigner transformation followed by a
canonical transformation [55]. In order for this to work, H may only contain pairs
of particle creation and annihilation operators. For space-independent reaction
rates, the complete list of reaction–diffusion process whose quantum Hamiltonian
H+ is similar through (10.65) to a free-fermion Hamiltonian H is as follows
[10, 40, 53] and shown in table 10.4. Since the transformation (10.65) is spatially
local, these correspondences actually hold in any dimension but free-fermion
methods are only available in one dimension.
where C1 (∞) is the steady-state density and τ = 1/(4 − 4h) is the relaxation time
(see [10] and references therein for more information on solved free-fermion
models).
This kind of analysis was generalized to find those reaction–diffusion
systems which are similar, via a transformation of the type (10.65), to the XXZ
chain [40]. While the full result is too complex to be re-stated here, an interesting
special case is given by the conditions
γR + βL + 2α + DL = νL + δL + 2σ + DR
(10.69)
γL + βR + 2α + DR = νR + δR + 2σ + DL .
In this case, the (usually infinite) hierarchy of equations of motion for the k-point
particle-density correlators Ck (t) = "n1 (t) . . . nk (t)# closes naturally, such that
C˙k (t) only depends on the C (t) with ≤ k. In principle, the equations of motions
for the Ck can then be solved iteratively [35].
O=1
C1(t) O=10
O=0
0.10
0.01
1 10 100 1000
t
Figure 10.3. Evolution of the mean particle density C1 (t) in the symmetric coagulation
model with the production reaction • ◦ • → • • • for several production rates λ. For long
√
times, the asymptotic behaviour is C1 (t) ∼ 1/ t for all values of λ (after [59]).
the empty-interval method [7, 12]. Consider a periodic chain with L sites and
lattice spacing a. Let In (t) be the probability that at time t, n consecutive sites are
empty. Then the mean particle density is [12, 52]
C1 (t) = (1 − I1 (t))/a. (10.70)
In order to illustrate the method, we consider the left–right symmetric pure
coagulation model and also take the free-fermion condition γ = D of model D
in table 10.4, but we now add a three-site production reaction • ◦ • → • • • with
rate 2Dλ [59]. The equations of motion for the In (t) read as
I˙1 (t) = 2D(I0 (t) − 2I1 (t) + I2 (t)) − 2Dλ(I1 (t) − 2I2 (t) + I3 (t))
(10.71)
I˙n (t) = 2D(In−1 (t) − 2In (t) + In+1 (t)) 2≤n≤L−1
together with the boundary conditions I0 (t) = 1 and IL (t) = 0 (assuming that
there is at least one particle in the system). The solution of these equations
is straightforward. For example, one obtains for the leading relaxation time
τ −1 = = 2Dπ 2 L−2 + O(L−4 ), in agreement with the results of section 10.4.
The effect of the production term is only transient,√ as illustrated in figure 10.3
for the mean density C1 (t). For λ = 0, C1 (t) ∼ 1/ t is of course expected from
equations (10.6) and (10.67). While the free-fermion condition γ = D is essential
for the method to work, we also see that the presence of the production term poses
no problem at all for the closure of the equations of motion (10.71)4.
4 This term cannot, e.g. by a similarity transformation, be turned into a term treatable by either free-
fermion or full integrability methods.
where h and h
are constants. In this case, the couplings in H are space-
dependent. The extension of the similarity/enantiodromy approach to this more
general setting remains to be done. Extensions of the empty-interval method to
interactions on more than two sites are studied in [60].
While the empty-interval method, as such, does not work for the pair-
annihilation process, the method has been generalized recently [61]. We briefly
explain the idea using the left–right symmetric pair-annihilation process with the
free-fermion condition α = D (model A or B) as example. Let Gn (t) be the
probability that, at time t, one has on n consecutive sites an even number of
particles. The mean particle density is C1 (t) = (1 − G1 (t))/a. Furthermore, let
Fn (t) (Hn (t)) be the probability that a segment of n consecutive sites with an
even (odd) number of particles is followed by the presence of a particle at the
(n + 1)th site. From the relations
Ġn (t) = 2D(Fn−1 − Hn−1 + Hn − Fn ) = 2D(Gn−1 (t) − 2Gn (t) + Gn+1 (t))
(10.74)
follow. They can be solved by standard methods. Reaction terms parametrized
by σ, νL,R , βL,R (see table 10.2) and even the reaction ◦ • ◦ → • • • can be
added [61]. Correlators are studied in [62].
In view of the practical success of these techniques it is perhaps not
completely futile to ask whether there might a be a systematic way to identify
these ‘empty-interval’ or related variables?
were found [26] through the quotients (P , M)HL−1 (q) as realized through the
Perk–Schultz model. They are collected in table 10.5. The following conventions
apply.
(1) For the first reaction in all models and the second reaction in models B, C, D
(with n < m understood), the reaction to the right (left) occurs with rate
R (L ).
(2) For models E, F the sum r + s has to be taken modN. If in this case
r + s = 0 mod N, the rate is L + R . If r + s = 0 mod N as well as for
the third reaction in models C, D the rate is R . In model D, it is also
assumed that, in the third reaction, pairs (r, s = r ± 1) never have an element
in common. If the products on the right are interchanged (e.g. A1 A1 → A1 ◦
in model C), the rate is L .
(3) In the third reaction in model F, the rates are ± respectively such that
+ + − = L + R .
√ √
One defines D = L R = 1 and q = R / L . The Hecke algebra
quotient (P , M)HL−1 (q) according to the realization as a Perk–Schultz quantum
chain equation (10.27) [24, 26] is also indicated.
From table 10.5 and theorem 10.3, we see that the simple diffusion model A
has, up to degeneracies, the same spectrum as the XXZ chain used in section 10.4
to describe biased diffusion of a single species of particles •. In the same way, the
spectrum of model E is, up to degeneracies, the same as the one found for pair
annihilation in section 10.4, with 2α = DL + DR .
For illustration, we briefly consider model E for N = 3. Each site may
contain either a particle of type A or B or be empty (◦). Single particles may
diffuse to the right A◦ → ◦A, B◦ → ◦B with a rate R or similarly to the
left with rate L . On encounter, between like particles the reactions AA → ◦B
and AA → B◦ occur with rates R and L , respectively, and similarly BB →
◦A, A◦. Two unlike particles react AB → ◦◦ with rate L + R . In the left–right
symmetric case, the identity of the spectra of H(E) and the one of pair annihilation,
where σ < ρ and σ, ρ ∈ {1, . . . , N} and gσρ ∈ R\{0}, gρσ ∈ R and the xσ are
complex parameters. If, in addition, the set A of generators Dσ admits a linear
Poincaré–Birkhoff–Witt (PBW) basis of ordered monomials Dσk11 Dσk22 · · · Dσknn
with σ1 > σ2 > · · · > σn and kn ∈ N, A is called a diffusion algebra [39].
These conditions imply certain constraints on the gσρ and the xσ , quite
analogously to the Jacobi identities of a Lie algebra. These constraints can be
fully solved and a classification of all diffusion algebras for N species is obtained
[39, 69]. The representation theory of N-species diffusion algebras is just getting
started, see [70].
steady state of the model is not critical, the relaxation towards it occurs through
domain coarsening and is very slow, the typical length scale varying with time as
L(t) ∼ t 1/z , see figure 10.4. Typically, the observables depend algebraically on
the time t passed since the quench, see [8, 9] for (a collection of) recent reviews.
Here we concentrate on the two-time spatio-temporal response function R(t, s; r)
of the time-dependent spin σr (t) at site r with respect to an external magnetic
field h0 (s) applied at the origin 0 at an earlier time s < t. Generically, two-time
quantities such as R(t, s; r) depend on both times t and s and not merely on the
difference τ = t − s. This breaking of time-translation invariance is called ageing.
For ageing systems, an extension of dynamical scaling is possible and allows
one to fix the form of the two-time response function. Specifically, it can be shown
that for a dynamical exponent z = 2 [74, 75, 77]
1+a−λR /2
δ"σr (t)# t −1−a M r2
R(t, s; r) = = r0 (t − s) exp − .
δh0 (s) h=0 s 2 t −s
(10.79)
Here a and λR are non-equilibrium exponents to be determined which
characterize the ageing universality class [78, 83]. Finally, r0 and M are non-
universal constants. We first present evidence that the response function of the
Glauber–Ising model in 2D and 3D is, indeed, given by (10.79). Then we discuss
where this presumably exact result comes from.
We first consider the autoresponse R(t, s) = R(t, s; 0). While R itself is
too noisy to be measured directly, integrated response functions are accessible
through simulations, see [79] and references therein for the details which we skip
over here. For the example, the integrated autoresponse
s
ρ(t, s) = du R(t, u) ∼ s −a fM (t/s) (10.80)
0
ρ (x,µ)
ln(fM(x))
−0.5
(2)
−1 0.22
−1.5
Figure 10.5. Scaling form of the integrated magnetic response in the Glauber–Ising model
as a function of x = t/s below criticality. The symbols correspond to different waiting
times s. The integrated autoresponse is shown in (a) two dimensions at T = 1.5 and in
(b) three dimensions at T = 3. An example of the integrated spatio-temporal response in
two dimensions at T = 1.5 and with µ = 2 is shown in (c). The full curves are obtained
from (10.79). After [79].
where µ is a control parameter. We stress that the scaling function ρ (2) no longer
contains the free non-universal parameter [79]. As an example, we compare in
figure 10.5(c) data from two dimensions taken with µ = 2 with equation (10.79).
Besides the expected scaling, the functional form of the scaling function neatly
follows the prediction. We stress that the position, the height and the width of
the maximum of ρ (2) in figure 10.5(c) are completely fixed. Similar results have
been obtained for other values of µ and in three dimensions as well. This provides
strong evidence that equation (10.79) is exact, at least in this model [79]. Tests of
(10.79) in different universality classes are described in [75].
e1 e1 e1
Figure 10.6. Root space of the complexified conformal Lie algebra conf3 , indicated by the
full and the open points. The double circle in the centre denotes the Cartan subalgebra. The
generators belonging to the three non-isomorphic parabolic subalgebras [82] are indicated
- 1 , (b) a.
by the full points, namely (a) sch -1 .
ge1 and (c) alt
For fixed M, the Schrödinger group is the maximal invariance group on the space
of solutions of equation (10.81). It is defined by the spacetime transformations
(R is a rotation matrix)
αt + β Rr + vt + a
t −→ t
= r −→ r
= αδ − βγ = 1 (10.82)
γt + δ γt + δ
and acts projectively on the solutions φ(t, r) [81]. Let schd be the Lie algebra of
(10.82). Time translations occur in schd and are parametrized by β. If we treat
the ‘mass’ M not as a constant but as another variable, the embedding schd ⊂
confd+2 for the complexified Lie algebras follows [82], where confd+2 is the Lie
algebra of the conformal group in d + 2 dimensions. From the classification of the
parabolic subalgebras of confd+2 we obtain several new subalgebras, called age -
- [82]. For the 1D case, we illustrate in figure 10.6 their definition through
or alt
the root space of conf3 ∼= B2 . These subalgebras still contain the generator for the
dilatations t → b2 t, r → br (which is in the Cartan subalgebra of conf3 ) but no
longer contain time translations (which is in the lower left corner of figure 10.6).
They are candidates for a dynamic symmetry algebra of ageing systems. If we
assume that the two-time response function transforms covariantly under the
- a set of linear differential equations for R(t, s; r) is
- or alt,
action of either age
obtained. Matching their solution with the expected [78] scaling behaviour of R,
we recover equation (10.79) in the special case z = 2.
The functional form of R depends on the fact that the Galilei transformation
of (10.82) is identical to the well-known one of a free particle. It is not trivial at all
that the response function of an interacting field theory such as the Glauber–Ising
and β and α are arbitrary functions. Whether this has a bearing on the
ageing behaviour of non-equilibrium spin systems is still open. Local scale-
transformations generalizing the Schrödinger group (10.82) to general values
of the dynamical exponent z = 2 exist [75]. It can be shown that R(t, s; r) =
R(t, s; 0)!(r(t − s)−1/z ), such that equation (10.79) holds for the autoresponse
R(t, s; 0) if λR /2 is replaced by λR /z and !(v) is given as the solution of a linear
differential equation of fractional order [75].
Acknowledgments
I thank M Pleimling and J Unterberger for their pleasant collaboration on local
scale-invariance and the members of the Groups de Travail Mathématiques/
Physique for keeping my interest in integrable systems alive. This work
was supported by CINES Montpellier (projet pmn2095) and the Bayerisch-
Französisches Hochschulzentrum (BFHZ).
References
[1] Derrida B, Lebowitz J L and Speer E R 2002 Phys. Rev. Lett. 89 030601
[2] Derrida B, Janowsky S A, Lebowitz J L and Speer E R 1993 J. Stat. Phys. 73 813.
Derrida B 1998 Phys. Rep. 301 65
[3] Schmittmann B and Zia R K P 1995 Phase Transitions and Critical Phenomena vol
17, ed C Domb and J Lebowitz (London: Academic Press)
[4] Privman V (ed) 1996 Nonequilibrium Statistical Mechanics in One Dimension
(Cambridge: Cambridge University Press)
[5] Marro J and Dickman R 1999 Nonequilibrium Phase Transitions in Lattice Models
(Cambridge: Cambridge University Press)
[6] Hinrichsen H 2000 Adv. Phys. 49 815
[7] ben-Avraham D and Havlin S 2000 Diffusion and Reactions in Fractals and
Disordered Systems (Cambridge: Cambridge University Press)
[8] Cates M E and Evans M R (ed) 2000 Soft and Fragile Matter (Bristol: IOP Press)
[9] Cugliandolo L F 2002 Preprint cond-mat/0210312
[10] Schütz G M 2000 Phase Transitions and Critical Phenomena vol 19, ed C Domb
and J Lebowitz (London: Academic Press)
[11] Smoluchowski M v 1917 Z. Phys. Chem. 92 129
[12] ben-Avraham D, Burschka M A and Doering C R 1990 J. Stat. Phys. 60 695
[13] Spouge J L 1988 Phys. Rev. Lett. 60 871
Spouge J L 1988 Phys. Rev. Lett. 60 1885 (erratum)