Notes Kupiainen
Notes Kupiainen
Notes Kupiainen
INTRODUCTION TO THE
RENORMALIZATION GROUP
Antti Kupiainen
1 Ising Model
We discuss first a concrete example of a spin system, the Ising model. This is a simple
model for ferromagnetism, i.e. the phenomenon that certain materials (e.g. iron) are can
stay permanently magnetized even in the absence of an external magnetic field. We can
think of such materials consisting of elementary magnetic moments, residing in atoms
located at a crystal lattice. A magnetic moment m ∈ R3 interacts with a magnetic field
B ∈ R3 with energy
E = −m · B
(· is the scalar product). Energy is minimized by having m parallel to B.
In Ising model one simplifies by letting each m take only two values, parallel or an-
B
tiparallel to the field: m = µ |B| σ, σ ∈ {1, −1} and µ > 0 a constant. The variable
σ is called "spin" where the terminology comes from the fact that the atomic magnetic
moments often come from the spin degree of freedom of electrons.
The crystal lattice is modeled by a regular lattice which for definiteness we take to be
Z i.e. x = (x1 , · · · , xd ) ∈ Zd with xi ∈ Z. To each x ∈ Zd , we associate a spin variable
d
σx ∈ {−1, 1}. We call σ = {{σx } | x ∈ Zd } a spin configuration on Zd . The set of all spin
d
configurations is denoted by Ω = {−1, 1}Z .
Physical lattices are finite and so we consider also spin configurations on finite subsets
Λ ⊂ Zd , |Λ| < ∞ where |Λ| be the number of elements in Λ. . We denote by ΩΛ = {−1, 1}Λ
the spin configurations in Λ. In practice |Λ| is very large (> 1023 ) so we need to study
the |Λ| → ∞ limit, so-called thermodynamic limit. For simplicity we mostly take Λ a
cube centered at origin: for L ∈ N let ΛL := {x ∈ Zd | |xi | ≤ L, i = 1, · · · , d} “cube of
side 2L + 1”.
1
Definition 1.1 Let |Λ| < ∞. The Ising Hamiltonian in Λ is HΛ : ΩΛ → R given by
X X
HΛ (σ) = −J σx σy − h σx (1.1)
{x,y}∈BΛ x∈Λ
σ = +1
i.e. we view ZL as the cyclic group of order L. We can also consider ZL1 × ZL2 × · · · × ZLd .
2
where
Λσ σ̄ Λc
Note that the ± boundary conditions are special cases where σ̄x = ±1, ∀x. Also note
that Hσ̄ depends on σ̄x only for x ∈ ∂Λ = {y ∈ Λc |dist(y, Λ) = 1}. Indeed,
X X X
σx σy = σx σy + σx σ̄y .
{x,y}∈B̄Λ {x,y}∈BΛ x∈Λ,y∈Λc ,|x−y|=1
Denote by HΛb.c. (h, σ) the Ising Hamiltonian in Λ with given b.c. as above. We also set
J = 1 since it will play no role below.
In statistical mechanics we view the the spin configurations σ ∈ ΩΛ random variables
whose probability distribution is determined by the Hamiltonian.
Remark. Recall some definitions from probability theory. A measure µ be of total mass
one on a σ-algebra Σ of subsets of some set M is a probability measure. A ∈ Σ is called
an event and µ(A) is the probability of A. Let f : M → R be measurable, where R is
equipped with the Borel σ-algebra, the smallest σ-algebra containing open sets in R. We
say that f is a random variable. The distribution of f is the probability measure ν on R
with ν(B) = µ(f −1 (B)). For the mean of f we use the following notations
Z
f dµ ≡ hf i ≡ E f
R R
and the variance of f is f 2 dµ − ( f dµ)2 .
Definition 1.3 The Ising measure on ΩΛ is the measure (note : ΩΛ is a finite set,
|ΩΛ | = 2|Λ| )
1 b.c. (h,σ)
µb.c.
β,h,Λ (σ) = b.c.
e−βHΛ
Z(β,h,Λ)
b.c.
where Z(β,h,Λ) is the partition function:
b.c. (h,σ)
X
b.c.
Z(β,h,Λ) = e−βHΛ
σ∈ΩΛ
X
so µ is a probability measure, µ(ΩΛ ) = µ(σ) = 1.
σ∈ΩΛ
3
1
Remark. In physics, β = kT where k = Boltzman’s constant and T = temperature : β
is the “inverse temperature”. Note that for β → ∞ (T → 0) the minima of H dominate :
at low temperatures we expect magnetism.
In the real world |Λ| ∼ 1023 , so we inquire
1◦ What happens as Λ → Zd ? Is there a limit measure µ = lim µΛ on Ω ? This is called
Λ→Zd
the thermodynamic limit.
2◦ Does µ depend on b.c. ? How does it depend on β, h ?
We will study µΛ via its correlation functions :
Definition 1.4 Let A ⊂ Zd , |A| < ∞. Denote
Y
σA = σx .
x∈A
4
Every f ∈ C0 (Ω) is a linear combination of the functions σA with |A| < ∞ with the
convention σ∅ = 1. Indeed, if g : {1, −1} → R we may write g(σ) = 21 (g(1) + g(−1)) +
1
2
σ(g(1) − g(−1)). Letting Px± f := 12 (f σx =1 ± f σx =−1 ) the desired representation follows
by expanding the product in
Y
f= (Px+ + σx Px− )f
x∈A
where A is the set of x s.t. f depends on σx . From this and (2.1) we conclude
exists for all f ∈ C0 (Ω). C0 (Ω) is a vector space and ` defines a linear map C0 (Ω) → R.
Clearly we have
`(1) = 1 (2.3)
and
Moreover, since |hf iΛ | ≤ supσ∈Ω |f (σ)| (where we take Λ to be large enough to include
the support A of f ; the sup is actually a max) we conclude
kf k = sup |f (x)|.
x∈M
A linear map ` : C(M ) → R is called a state on C(M ) if (2.3)-(2.5) hold. To get into this
setup we need to discuss the topology of Ω.
d
Ω is a compact metric space: {−1, 1} is compact, so Ω = {−1, 1}Z is also compact
in the product topology by the Tychonov theorem (see [35] chap. 4 or [25] chap. 5). A
metric compatible with this topology is e.g.
X
d(σ, σ 0 ) = 2−|x| |σx − σx0 |
x∈Zd
i.e. two spin configurations are close if they agree in a big box around origin. Using the
Stone-Weierstrass theorem (see [35] chap. 4) one then shows that C0 (Ω) is dense in C(Ω).
Thus our linear functional ` extends to C(Ω) and defines a state there.
5
Homework : Check all this !
We may now use the basic real analysis to get our infinite volume measure. Recall
that on a compact metric space we may consider the σ-algebra of Borel sets of M i.e.
the smallest σ-algebra containing the open sets and Borel measures which are measures
defined on this algebra. Given a Borel probability measure µ on M the linear map
Z
`µ (f ) = f dµ
on C(M ) satisfies clearly the conditions (2.3)-(2.5) and defines a a state on C(M ). Con-
versely:
Theorem 2.1 (Riesz Representation Theorem) Given a state ` on C(M ) there is
a Borel probability measure µ on M such that
Z
`(f ) = f dµ
Proof See Rudin, Real and Complex Analysis, chap 2 [37] or Reed-Simon, vol. 1, chap. 4
[35]
Exercise Prove this directly for M = ΩΛ !
We have thus obtained the following
Corollary 2.2 Suppose that for all finite A ⊂ Zd the limit limL→∞ hσA iσ̄ΛL exists. Then
there is a probability measure µσ̄ such that this limit equals
Z
hσA i := σA dµσ̄ .
σ̄
We call h−iσ̄ (or µσ̄ ) infinite volume state and the question of thermodynamic limit
is : find all infinite-volume states by taking different b.c. We’ll see that for all b.c. the
Λ → Zd limit exists (at least, through subsequences), but there can be many different µ0 s.
Consider the Ising model above for d = 1. Let first h = 0 and consider Hσ̄
L+1
X
βHLσ̄ := −β σi−1 σi
i=−L
where σ−L−1 = σ̄−L−1 ≡ σ − and σL+1 = σ̄L+1 ≡ σ + (note that these may vary with L).
Consider first the partition function:
X L+1
Y
ZLσ̄ := eβσi−1 σi .
σ−L ,··· ,σL =±1 i=−L
6
Let T = (Tσσ0 )σ,σ0 =±1 be the 2 × 2 matrix:
!
0 eβ e−β
(Tσσ0 ) = (eβσσ ) = .
e−β eβ
Then,
X
ZLσ̄ = eβσ−L−1 σ−L (T 2L )σ−L σL eβσL σL+1 = (f − , T 2L f + )
σ−L ,σL =±1
± ±
where we use the scalar product (f, g) ≡ σ fσ gσ and f ± are the vectors (eβσ , e−βσ ).
P
T is a symmetric matrix with has eigenvalues the roots of (eβ − λ)2 − e−2β i.e.
2 cosh β := λ1 and 2 sinh β := λ2 . The corresponding orthonormal eigenvectors are
e1 = √12 (1, 1)T , e2 = √12 (1, −1)T and introducing the corresponding orthogonal projec-
tions ! !
1 1 1 1 1 −1
P1 = , P2 =
2 1 1 2 −1 1
we may write
T = λ1 P1 + λ2 P2 .
Since P1 P2 = 0 and Pi2 = Pi
and
7
It has the L → ∞ limit
1
F σ̄ = − log(2 cosh β)
β
which is independent on σ̄. Note that log ZL is extensive i.e., once divided by |Λ|, it has
the thermodynamic limit.
Let us next consider the magnetization:
1 X −βHLσ̄ (σ)
hσx iσ̄L = e σx .
ZLσ̄ σ
where in the last formula σ denotes the diagonal matrix σδσσ0 . But P1 σP1 = 0 so the first
term in the numerator vanishes and combining with (3.1) we obtain
i.e. 2-point correlation function decays exponentially (here we had hσx i = 0). ξ = α1 is
called the correlation length. Note that ξ → ∞ as β → ∞. At low temperatures, the 1d
8
model gets more correlated.
It is easy to show now (exercise) : let x1 < x2 < · · · < x2n . Then
2n n
!
Y X
h σxi iσ̄L L→∞
−→ exp −α (x2i − x2i−1 )
i=1 i=1
i.e. all the correlation functions have a limit which is independent on σ̄ and equals
2n
Y n
Y
h σ xi i = hσx2i−1 σx2i i. (3.3)
i=1 i=1
We get
Theorem 3.2 For the 1d Ising model the thermodynamic limit exists and is independent
on b.c. (for all β ∈ [0, ∞)). The correlation functions of the ∞-volume state are given by
(3.3) with
hσx σy i = (tanh β)|x−y|
1
Exercise. Calculate ZLb.c. (β, h) and F (β, h) = − limL→∞ 2βL
log ZLb.c. (β, h). Show the
limit is independent of b.c. and analytic in β, h.
Remark. T is called the transfer matrix. What was important above was that
(f σ̄ , e1 ) > 0. This is a general fact, following from the Perron-Frobenius theorem (see e.g.
[39] and [21], Theorem 3.3.2, for an extension to compact operators).
Exercise. Consider the d-dimensional Ising model in the cylinder [−L, L] × Zd−1 ` i.e.
we have periodic boundary conditions in d − 1 dimensions with period ` and σ̄ boundary
conditions in the two ends of the cylinder. Write the partition function and the correlations
in terms of a transfer matrix and using the Perron-Frobenius theorem show the limits exist
as L → ∞ (but ` fixed) and are independent on σ̄. We’ll see later the last statement is
not true in low temperatures if we took ` = L.
Remark : What is shown above means that the 1d Ising model has no phase transition
i.e. there is a unique infinite volume state. We’ll discuss this later.
Ising model in two dimensions. Onsager succeeded calculating the free energy F (β)
in closed form in 1944 [32]. The result is not an analytic function of β. At the positive
√
real axis F (β) has a point of non analyticity at β = βc given by tanh(2βc ) = 1/ 2
where the second derivative of F (the specific heat) diverges logarithmically ∂β2 F (β) ∼
const. log |β − βc |. Subsequently he showed the existence of phase transition (see below)
9
by computing the magnetization m(β) = hσ0 i+ in closed form. It vanishes if β ≤ β0 and
is nonzero for β > β0 . Furthermore as β ↓ βc m(β) ∼ (β −βc )1/8 . He also calculated the 2-
point correlation function which has exponential decay for β 6= βc and at βc hσ0 σx i ∼ |x|1/4
as x → ∞. These results were revolutionary. It was the first proof of phase transitions
from first principles. The critical exponents (1/8 and 1/4 above and others) were different
form naive expectations (mean field theory) and called for explanation. This was finally
achieved with Wilson’s renormalization group theory 30 years later [44].
Other exactly solvable models exist in d = 2 (see Baxter’s book [4]). However, this
is very exceptional and, in d > 2, there are practically none so other methods are called
for. We will first develop methods to study hσA i for β small (high temperature) and β
large (low temperature). These methods work for all d and for much more general models
than the Ising model. The high and low temperature expansions are still the best ways
to numerically to study the critical point.
d
Recall our setup: Ω = {−1, 1}Z and the boundary condition on the boundary of
Λ ⊂ Zd , |Λ| < ∞, is given by a configuration σ̄ ∈ Ω.
Note that to say σ̄ = +1 means that we have on the boundary of Λ a positive magnetic
field that tends to force σx to be +1. The question is : can this result in a positive hσx i
as Λ ↑ Zd ?
Theorem 4.1 (a) There exists β0 > 0 such that if β < β0 then hσA iσ̄β,h,Λ has a limit as
Λ → Zd (via cubes say), independent on σ̄.
(b) Let h = 0, d ≥ 2. There exists β1 < ∞ such that if β > β1 then there are at least two
different infinite-volume (pure) states h−i+ 6= h−i− .
Remarks. 1. (a) means uniqueness of the Gibbs state, see below.
2. (b) means phase transition. See next section.
3. In d = 2 there are only 2 (pure) Gibbs states [1], in d ≥ 3 there are many more [12].
4. Finally, it can be proved, see [3] that β0 = β1 .
We prove (a) in this section and (b) in section 4.3. The result is due to Peierls,
Dobrushin and Griffiths [33, 9, 22].
10
Thus !|A|
X
hσA ib.c.
Λ = σνh (σ) = (tanh h)|A|
σ=±1
factorizes and is Λ-independent. We wish to show that for small β our measure approxi-
mately factorizes. Let h = 0 for simplicity. Write (since σx σy ∈ {−1, 1})
where we recall BΛ denotes the set of all bonds b ⊂ Λ. Thus (with |BΛ | the cardinality of
BΛ )
f ree Y
e−βHΛ (σ) = (cosh β)|BΛ | (1 + σb tanh β).
b∈BΛ
Insert this in hσx σy iΛ , cancel (cosh β)|BΛ | in numerator and denominator and expand the
product over b: X X Y
σx σy σb (tanh β)|B|
σ∈ΩΛ B⊂BΛ b∈B
hσx σy ifΛree = X X Y .
σb (tanh β)|B|
σ∈ΩΛ B⊂BΛ b∈B
Definition. Given B ⊂ BΛ , a family of bonds, let ∂B denote the set of sites x ∈ Λ that
occur an odd number of times in the bonds in B.
11
Example.
y
∂B = {x, y, z, w}
x w
Z
B
X
Since σ n = 0 if n is odd, we get
σ=±1
X
(tanh β)|B|
B:∂B={x,y}
hσx σy ifΛree = X .
(tanh β)|B|
B:∂B=∅
This is the starting point of the high temperature expansion. We use it to prove the
existence of a finite correlation length.
set as above and ∂B 0 = ∅. Given B, P is not unique, but let us choose it arbitarily, and
call the choice P (B).
We may then rewrite the numerator as
0
X X
(tanh β)|P | (tanh β)|B |
P B 0 ∈B(P )
where the sum over P runs through the paths joining x and y and B(P ) consists of sets
B 0 ⊂ BΛ such that ∂B 0 = ∅ and P (P B 0 ) = P . Since B(P ) ⊂ {B ∈ BΛ |∂B = ∅} each
S
12
term in this sum occurs in the denominator and we have
|B 0 |
P
B 0 ∈B(P ) (tanh β)
P |B|
≤1
B:∂B=∅ (tanh β)
and so
X
0 ≤ hσx σy iΛ ≤ (tanh β)|P | . (4.1)
P
To control the sum over P write as above P = {(xi , xi+1 )}ni=1 , x1 = x, xn+1 = y. Given
xi there are at most 2d nearest neighbours xi+1 and n is at least |x − y|. Thus
X
(4.1) ≤ (2d tanh β)n ≤ Ce−|x−y|/ξ
n≥|x−y|
if 2d tanh β < 1 .
Remarks.
1. We could have chosen P self-avoiding and so could have had only 2d − 1 choices :
xi+1 6= xi−1 . Thus (2d − 1) tanh β < 1 suffices.
√
2. In d = 2 , tanh βc = 2 − 1 = 0, 4142.... We got tanh β < 31 which is not that bad. As
d → ∞, our estimate tanh βc ∼ 1/(2d − 1) becomes exact by combining [16] and [18], see
[6].
We now slightly modify the previous argument to prove part (a) of Theorem 4.1.
Proof of Theorem 4.1 (a) We prove uniqueness (independence on b.c) of the limit.
Existence is analogous. Consider σ̄, σ̄ 0 ∈ Ω and
σ̄
X
e−βHΛ (σ) σA
0 σ
hσA iσ̄Λ − hσA iσ̄Λ = X σ̄
− (same with primes)
e−βHΛ (σ)
σ
σ̄ σ̄ 0 0
X
e−βHΛ (σ)−βHΛ (σ ) (σA − σA0 )
σ,σ 0
= X σ̄ σ̄ 0 0
. (4.2)
e−βHΛ (σ)−βHΛ (σ )
σ,σ 0
We expand slightly differently (to understand why, see Remark after the Proof).
0 0
eβ(σx σy +σx σy ) = e−2β (1 + fxy )
0 0
fxy = eβ(σx σy +σx σy +2) − 1.
Note that
13
using 0 ≤ ez − 1 ≤ zez for z ≥ 0, and 0 ≤ σx σy + σx0 σy0 + 2 ≤ 4. Expand in powers of
f as before, but only in the numerator : let B̄Λ be the set of bonds intersecting Λ (i.e.
b = {x, y}, either x or y is in Λ), then
X X Y
(σA − σA0 ) fb
0 B⊂B̄Λ σσ 0 b∈B
hσA iσ̄Λ − hσA iσ̄Λ = XY (4.4)
(1 + fb )
σσ 0 b∈B̄Λ
b1 b2 b3 bn
is antisymmetric.
Hence, the nonvanishing terms in (4.4) have B’s such that B includes a connected path P
of bonds joining A to Λc .
Bound the numerator as
Y Y Y Y Y
|(σA − σA0 ) fb | ≤ 2 fb fb ≤ 2(4β)|P | (1 + fb ) fb
b∈B b∈P b∈B\P b∈P b∈B\P
using (4.3) in the last inequality (Recall that fb ≥ 0 !!); and so, picking, as in the proof
of Theorem 4.2, for each B, a choice P (B) of P ,
XX Y X X X Y Y
| (σA − σA0 ) fb | ≤ 2 (4β)|P | (1 + fb ) fb (4.5)
B σ,σ 0 b⊂B P σσ 0 B 0 ∈B(P ) b∈P b∈B 0
so
X XY
(4.5) ≤ 2 (4β)|P | (1 + fb ). (4.6)
P σσ 0 b∈B̄Λ
14
P Q
The denominator in (4.2) equals σσ 0 b∈B̄Λ (1 + fb ), so combining (4.2) and (4.6) we
obtain
0
X XX X
| hσA iσ̄Λ − hσA iσ̄Λ | ≤ 2 (4β)|P | ≤ 2 (4β)|P | . (4.7)
P x∈A y∈∂Λ P :x→y
Taking 8dβ < 1 the sum over paths is bounded by C(8dβ)|x−y| and then
0
| hσA iσ̄Λ − hσA iσ̄Λ | ≤ C|A| |∂Λ|(8dβ)dist(A,∂Λ) . (4.8)
where ∂Λ is the boundary of Λ. For Λ = ΛL a cube of side L |∂Λ| ∝ Ld−1 , dist(A, ∂Λ) ∝ L
and thus (4.8) is bounded by ≤ C|A|Ld−1 δ L with δ < 1. This tends to zero as L → ∞. .
Remark Note that the positivity of fxy was crucial in (4.5). This is why we did not use
the previous expansion. This method generalizes to a very general class of Hamiltonians,
see [7].
Exercises.
1. Estimate |hσA iσ̄Λ − hσA iσ̄Λ0 | in for Λ ⊂ Λ0 the same way:
0
X X
e−βHΛ (σ) e−βHΛ0 (σ ) (σA − σA0 )
σ∈ΩΛ σ 0 ∈ΩΛ0
hσA iσ̄Λ − hσA iσ̄Λ0 = X X 0
.
e−βHΛ (σ)−βHΛ0 (σ )
σ∈ΩΛ σ 0 ∈ΩΛ0
0
Expand e−βHΛ (σ)−βHΛ0 (σ ) as above and note that only B 0 s connecting A to Λc contribute.
Use this to show that the limit as Λ % Zd exists.
2. Similarily, write
1 X −βH(σ)−βH(σ0 ) 0
X 0
|hσX σY iσ̄Λ − hσX iσ̄Λ hσY iσ̄Λ | = e [σX − σX ][σY − σY0 ]/ e−βH(σ)−βH(σ ) ,
2 σσ0 σσ 0
0
expand e−βH(σ)−βH(σ ) as above and prove:
since here only B 0 s connecting X and Y will contribute i.e. all correlations functions
decay exponentially.
15
say in d = 2, the configuration σx = 1 ∀x 6= x0 , σx0 = −1:
+ + +
+ − +
+ + +
This has energy βH = βHminimum +8β (4dβ in general dimension). Consider a connected
region R of −:
+ + + + +
+ − − − +
+ − − − +
+ − − − +
+ + + + +
This has H = Hmin + 2|∂R| where |∂R| = # bonds {x, y} with x ∈ R, y ∈ Rc . We say
that there is a contour around R. Let us formalize this. Given σ ∈ ΩΛ , let
plaquettes.
Now a bond b in Zd defines a unique (d − 1)-cell b∗ "orthogonal" to it in the dual
lattice:
b*
b*
b b
d=3
d=2
i.e. C ∗ (σ) is a set of bonds (d=2) or plaquettes (d=3) in the dual lattice.
16
We say two bonds b 6= b0 are connected if |b ∩ b0 | = 1 (i.e. they share one site) and two
plaquettes p 6= p0 are connected if |p ∩ p0 | = 2 (i.e. they share a bond). Given a set B of
bonds (plaquettes) consider the graph G(B) with vertex set B and edges {b, b0 } if b and b0
are connected. Let Gα be the connected components of G(B) and Bα the vertices of Gα .
We call Bα the connected components of B. We say B is connected if G(B) is connected.
By some abuse we say B and B 0 are disjoint if no b ∈ B, b0 ∈ B 0 are connected.
We will now characterize the connected components of C ∗ (σ).
Definition 4.3 A contour γ is a connected set γ ⊂ B̄Λ∗ , such that (d=2) each site x of
the dual lattice belongs to an even number of bonds b∗ , with b∗ ∈ γ or (d=3) each bond
of the dual lattice belongs to an even number of plaquettes b∗ ∈ γ. Hence contours are
closed paths (d=2) or closed surfaces (d=3). We say a family Γ of contours is compatible
if all γ ∈ Γ are connected and all γ, γ 0 ∈ Γ are disjoint.
Example.
Allowed :
Not allowed :
Lemma 4.4. The family Γ(σ) of connected components of C ∗ (σ) is a compatible family
of contours. Conversely, given a compatible family Γ of contours there is a unique σ such
that Γ = Γ(σ).
Proof. C ∗ (σ) has the property that x belongs to an even number of bonds (see figure).
Hence its connected components have this property.
Conversely, given Γ, let x0 ∈ Λc (so σx0 = +1) and let P be any path of bonds from x0
to x. Let N (P ) be the number of b ∈ P with b ∈ C(σ), i.e. the number of times P crosses
contours (see figure). Put σx = (−1)N (P ) . This is well defined since if P 0 is another path
then N (P ) − N (P 0 ) is even (show!).
b*1
b1
u v
b2 We have since σ 2 = 1, σu σv σv σw σw σy
b*4
σy σu = 1 = σb1 σb2 σb3 σb4 ⇒ even num-
x b*2
ber of σbi = −1
b4
y w
b3
b*3
17
Λ
- +
+ +
+
Let us denote by GΛ the set of compatible families Γ of contours γ ∈ B̄Λ∗ . The Lemma
implies:
X
−βHΛ+ (σ) = βNΛ − 2β |γ|
γ∈Γ(σ)
and
X Y
ZΛ+ = eβNΛ e−2β|γ| (4.9)
Γ∈GΛ γ∈Γ
Nx (Γ) = # of γ ∈ Γ surrounding x.
4.3 Magnetization
We prove Theorem 4.1. (b) by proving
Theorem 4.3. (Peierls argument). There exists β1 < ∞, δ > 0 such that for β > β1
−
hσ0 i+
Λ = −hσ0 iΛ ≥ δ
for all Λ.
Proof. We have
hσ0 i+
Λ = P(σ0 = 1) − P(σ0 = −1) = 1 − 2 P(σ0 = −1).
Now P(σ0 = −1) equals the probability that there are an odd number of contours sur-
rounding origin which in turn is bounded from above by the probability that there exists
18
a contour surrounding origin. Denote by G0Λ ⊂ GΛ those compatible families of contours
for which there exists γ0 ∈ G0Λ surrounding origin. Hence
X Y X Y
P(σ0 = −1) ≤ e−2β|γ| / e−2β|γ| .
Γ∈G0Λ γ∈Γ Γ∈GΛ γ∈Γ
Since each term in the sum in the numerator occurs in the denominator, we get
X ∞
X
−2β|γ|
P(σ0 = −1) ≤ e = e−2βn #{γ : γ surrounds 0, |γ| = n}. (4.11)
γ surrounds 0 n=2d
and so
#{γ : γ surrounds 0, |γ| = n} ≤ (2n)d c2n
d .
Hence, for β > log cd the series in (4.11) converges, uniformly in Λ and tends to zero as
β → ∞. The claim follows. Obviously hσ0 i− +
Λ = −hσ0 iΛ .
Lemma 4.4. . The number of connected subsets of B̄Λ∗ of cardinality n containing b∗ is
bounded by c2n
d with c2 = 4 and c3 = 12.
Proof. A set B ∗ ⊂ B̄Λ∗ is connected if and only if the graph G(B ∗ ) is. By Lemma 4.5.
any such graph is covered by doing a walk on B̄Λ∗ with starting point b∗ , length 2n and
jumps between connected vertices. In 2d each b∗ has 4 such neighbors and in 3d 12. The
number of such walks is c2n
d . The claim follows.
Lemma 4.5 (Köningsberg bridge Lemma ) Let G be a finite connected graph and let α0
be a vertex of G. Then there is a path starting and ending at α0 which includes each line
of G only twice.
Proof. Induction in the cardinality of the vertex set v(G). Let v(G) = n and {α1 , α2 } be
an edge of G. Consider the graph G0 = G\{α1 , α2 }. If G0 is connected, by induction there
is walk w : α0 → α0 visiting all its edges twice. The walk w0 : α0 → α1 → α2 → α1 → α0
where the first α1 is the first visit of w to α1 is the desired walk.
If G0 is disconnected and α0 and α1 are in the same connected component induction
provides two walks α0 → α0 and α2 → α2 . Then the desired walk is α0 → α1 → α2 →
α2 → α1 → α0 .
19
Remark 1 hσ0 i 6= 0 is called spontaneous symmetry breaking : HΛ (σ) is invariant (if
h = 0) under σ → −σ except for the b.c. (HΛf ree and HΛper are). As Λ % Zd the b.c. have a
finite effect ! Another way to state this : Let h 6= 0. Construct limΛ%Zd h−iσ̄Λ,h ≡ h−ih . It
will be independent on σ̄ (see below). Then limh↓0 h−ih = h−i+ , limh↑0 h−i− . In particular,
let m(h) = hσ0 ih . Then m(h) is discontinuous at h = 0. This is called a first order phase
transition.
Remark 2 The existence of the limit Λ % Zd can be proved along the same lines as for
β small above.
Numerator : x
Denominator :
Try to cancel the so-called “vacuum graphs”, i.e. the ones not involving x or y. Decompose
S
B into connected components: B = α Bα , Bα with Bα connected sets of bonds (as defined
in Section 4.3). Denote by B a family of connected, mutually disjoint (i.e. disconnected)
sets of bonds and by B the set of such families. In the numerator, we have one component,
say B1 , such that ∂B1 = {x, y}. So
X X Y XY
hσx σy ifΛree = ρ(B1 ) ρ(B)/ ρ(B).
B1 connected ∂B1 ={x,y} B:{B1 }∪B∈B B∈B B∈B B∈B
We have defined
ρ(B) = (tanh β)|B| .
The algebra we will perform does not depend on the explicit expression of ρ. The cancel-
20
lation is performed by the following trick. Consider the partition function
∞ n
XY X 1 X Y
Z = ρ(B) = ρ(Bα ) (4.12)
B∈B B∈B n=0
n! α=1
(B1 , B2 , . . . , Bn )
Bi ∩ Bj = ∅
∞ n
X 1 X Y Y
= ρ(Bα ) χ(Bα , Bβ ) (4.13)
n=0
n! α=1 α<β
(B1 , . . . , Bn )
( no constraint)
where the n! comes from writing the sum over the unordered sets B = {B1 , . . . , Bn } as
one over ordered ones (B1 , B2 , . . . , Bn ) and on the second line we introduced
(
0 Bα ∩ Bβ 6= ∅
χ(Bα , Bβ ) =
1 Bα ∩ Bβ = ∅
so that the last product imposes the constraint on the first line.
Note that if we didn’t have the constraint in (4.13) we could do the sum
∞ n
X 1 X Y X
ρ(Bα ) = exp( ρ(B)). (4.14)
n=0
n! α=1 B⊂BΛ
(B1 , . . . , Bn )
no constraint)
(
We will now expand the constraint to derive a generalization of (4.14). To achieve this
we write χ(Bα , Bβ ) = 1 − η(Bα , Bβ ) with
(
0 Bα ∩ Bβ = ∅
η(Bα , Bβ ) = .
1 Bα ∩ Bβ 6= ∅
Then
Y X Y
(1 − η(Bα , Bβ )) = (−η(Bα , Bβ )) (4.15)
α<β G (α,β)∈G
Note that this forces the corresponding Bα ’s to form a conneted network since η 6= 0 only
[
if Bα ∩ Bβ 6= ∅ . Let Ci = Bα .
α∈v(Gα )
21
Reorganize the sum (4.13) with (4.15),(4.18) inserted :
sum over k and a sum over k connected sets (C1 , . . . , Ck ).
sum over n ≥ k
sum over all families of k subsets of {1, . . . , n} (i.e. v(Gi ) above)
sum over Bα ’s
sum over connected graphs on those subsets.
So, one gets:
∞ X 1
X X X n! 1
Z = (∗)
k=0 (C1 ,··· ,Ck ) n≥k
n! n1 ! · · · nk ! k!
(n1 , · · · , nk )
X
ni = n
i
Yk X X ni
Y Y
·
ρ(Bα ) (−η(Bα , Bβ ))
i=1 (B1 , · · · , Bni ) G α=1 (α,β)∈G
ni
[
Bj = Ci
connected graph on
{1, · · · , ni }
j=1
The combinatorial factor (*) is the number of ways to choose a family of k subsets of
{1, . . . , n} of sizes {n1 , . . . , nk }. Indeed, this number equals
n−1
! ! ! X
n n − n1 n − n1 − n2 n− ni 1
··· i=1 k!
n1 n2 n3
nk
n! (n − n1 )! 1 n! 1
= · · · = Qk
n1 !(n − n1 )! n2 !(n − n1 − n2 )! k! i=1 ni !
k!
!
n
where comes from choosing the n1 elements of set 1 etc. and we divide by k!
n1
because order of the sets does not matter.
Thus
∞ k ni
X 1 X X Y 1 X X Y Y
Z = ρ(Bα ) (−η(Bα , Bβ ))
k=0
k! n!
i=1 i α=1
(C1 ,··· ,Ck ) (n1 ,··· ,nk ) (B1 , · · · , Bni ) G (α,β)∈G
ni
[
Bj = Ci
connected on
{1, · · · , n}
j=1
∞
!k
X 1 X X
= f (C) = exp f (C)
k=0
k! C
with
n
X X 1 X Y Y
f (C) = ρ(Bα ) (−η(Bα , Bβ )) (4.17)
n
n! α=1
(B1 , · · · , Bn ) G (α,β)∈G
connected on
[
Bj = C
{1, · · · , n}
22
We have achieved our goal. With our assumptions, the exponent is O(Λ). The main
estimate that guarantees this is
Lemma 4.6 Let |ρ(B)| ≤ |B| . There exists 0 s.t < 0 ⇒ |f (C)| ≤ (2)|C| .
(2 is arbitrary here). (Recall, B is always connected here).
Remark 3 Note that this is non-trivial. We can not estimate in (4.17) | − η| ≤ 1. Indeed,
2
the number of graphs on {1, · · · , n} is the subsets of {1, · · · , n} × {1, · · · , n} i.e. 2n and
2
the number of connected graphs is also ≥ ean a > 0 which is much bigger than n! ≤ nn so
the sum does not converge absolutely. One needs to account for cancellations. Note that
P Q Q
all G (−η) = χ(Bα , Bβ ) which is ≤ 1. Connectedness of G makes estimates harder.
As an example consider the trivial case of Ising model in a box of side 2. There is only
one nontrivial term in the high temperature expansion for the partition function:
(−1)n
ρ(B)n
P
Z = 1 + ρ(B) = elog(1+ρ(B) = e− n n
with ρ(B) = (tanh β)4 . On the other hand, in the expression for f (B) we have Bi = B
for all i and so
X ρ(B)n X Y
f (B) = (−1)G
n
n! G {α,β}∈G
where G runs through connected graphs on {1, . . . , n}. Comparing these two expressions
n
we see that this sum equals (−1)
n
.
Given the Lemma, write
X X X 1 X
f (C) = f (C) ≡ fb,Λ
C⊂BΛ b∈BΛ C⊂BΛ :b∈C
|C| b∈B Λ
where we noted that in the second sum each C ⊂ BΛ is counted |C| times. Then
∞
X 1
|fb,Λ | ≤ (2)n Sn
n=1
n
23
!
Also, we get : X X
ρ(B) exp f (C)
∂B={x,y} C∩B=∅
hσx σy ifree
Λ = X
exp( f (C))
C
X X
= ρ(B) exp f (C) ≡ ρ̃(B) (4.18)
∂B={x,y} C:C∩B6=∅
and |ρ̃(B)| ≤ |B| eC|B| ≤ (2)|B| . (4.18) is a version of the high temperature expansion.
It can be further processed but we won’t do it here.
• At h = 0, β = βc one expects
c
hσx σy i ∼|x−y|→∞
|x − y|a
At d = 2, a = 1/4 (exact)
d=3 a = 1.0364 (numerical, see wikipedia Ising critical exponents)
d≥4 a=d−2 (proven for d > 4)
Because of the behaviour for d ≥ 4, one usually writes a = d − 2 + η, and one calls η the
“anomalous exponent”. We will explain the exponent d − 2 below.
Rigorous results include : βc is unique, ξ < ∞ up to βc , Λ % Zd limit exists for all β and
h, and much else. The deepest is η = 0 for d > 4.
There are two different kinds of phase transitions in the Ising model:
1. First order transition For h 6= 0 let m(h, β) = hσx i where the expectation is in the
unique state (this state is translation invariant so hσx i is independent of x). If β > βc
(low temperatures) then limh↓0 m(h, β) = − lim m(h, β) > 0 i.e. m is discontinuous at
h↑0
h = 0 (if β ≤ βc , m is continuous).
24
h
unique state
-1 -1
βc β
unique state
2 states
Then m(β) > 0 for β > βc , m(β) = 0 for β < βc . What about m(βc )? m is continuous at
βc : limT % Tc m(T ) = 0 = m(Tc ) (we use T = β −1 as a parameter).
However, | dm
dT
|→ ∞ as T % Tc . One has m(T ) ∼ |Tc − T |β where β (not equal to T −1
here!) is a critical exponent for the magnetization. In d = 2, β = 18 (exact), in d = 3,
β = 0.3265 (numerical), in d ≥ 4 one expects β = 12 (this is proven for d ≥ 5). The
continuity or discontinuity of the order parameter (m here) is refered to as second order
and first order transition. Note also that ξ remains finite at a first order transition but
becomes infinite at second order transition.
25
Remark 1. The states above are translation invariant i.e.
hσA i = hσA+x i ∀A ⊂ Zd , ∀x ∈ Zd
where A + x = {a + x|a ∈ A}. Thus Ising model has two translation invariant states in
low temperature and one in high temperature or nonzero field.
Ising model has also non-translation invariant ones. One considers so called Dobrushin
boundary conditions, see the figure.
These force a contour (γ in the figure) joining the points where the boundary condition
changes. It turns out that in d = 2 one does not get a new state this way [1]. In a box
of side the contour fluctuates strongly: the expected deviation from the horizontal line is
1
L 2 and an observer sitting at origin will see either a plus state or a minus state: in the
L → ∞ limit the state is a convex combination:
with some p ∈ (0, 1). However, in d = 3 at low temperatures the fluctuations are O(1) in
L and one does get a new state [12] with an interface : hσx i > 0 for x far above, hσx i < 0
for x far below. However it is believed that there is a βr > βc , the roughening transition
point so that the interface state disappears for β < βr due to large fluctuations of the
interface.
Remark 2 Why is h 6= 0 unique ? Let e.g. h > 0 Let us try to construct – a state:
26
P P
The weight of σ = −1 i.e. the configuration with no contours is e−β(− σx σy −h σx ) =
eβ|B̄Λ | e−βh|Λ| . The weight of the configuration where we have a contour at the boundary
i.e. γ = ∂Λ :
2
equals eβ|B̄Λ | e−βh( h |∂Λ|−|Λ|) . For Λ a L−box |∂Λ| ∼ Ld−1 , |Λ| ∼ Ld so the contour is more
probable! Hence, we expect that the Λ → Zd limit is the same as that of the + boundary
condition state. One can make a proof along these lines, see [20].
Remark 3 What about h which is not constant :
X X
βH = β(− σx σy + hx σx ).
x∈Λ
Let us consider hx ’s random : let each hx be a random variable with zero mean : hx = 0
and different hx , hy are independent, identically distributed. Thus we pick a configuration
of h = {hx } randomly from such ensemble and consider the Gibbs state with this h. What
do we expect ? Consider low temperature and say + boundary conditions. A contour γ
costs now an energy and it weight is
X
exp(−β[2|γ| + 2 hx ])
x∈Intγ
+
–
Int γ
X
Let γ be an L-cube, so |γ| ∝ Ld−1 , how about hx ? This is a sum of Ld independent
x∈Int γ
random variables, therefore of size ∝ Ld/2 since its variance is
X X X X
E( hx )2 = Ehx hy = Eh2x + Ehx Ehy = Ld Eh2
x,y x x6=y
27
Thus, if d − 1 > d2 i.e. if d > 2, the magnetic field is unlikely to suppress contours and we
expect 2 states. For d ≤ 2 it seems to have a chance.
Theorem 4.8 For β large, Eh2 small, d ≥ 3 our model has 2 states: one with hσx i > 0
and one with hσx i < 0. For d ≤ 2 there is only one Gibbs state (this holds with Probability
one in h).
(δ is Kronecker delta). Thus to minimize energy nearest neighbours want to be the same.
For q = 2 this is Ising model: δσσ0 = 21 (1 + σσ 0 ). The model has Sq symmetry (= group
of permutations on q objects): Let π ∈ Sn (so π is a bijection from {1, · · · , q} into itself),
and Tπ : ΩΛ → ΩΛ : (Tπ σ)x = πσx . Then H(Tπ σ) = H(σ). At low temperatures we have
q phases:
Exercise. Devise a Peierls argument to construct a state h−ip with boundary condition
σ̄x = p so that as β → ∞, P(σ0 = p) → 1.
At high temperatures there is a unique infinite volume limit which can be constructed
with a high temperature expansion. We can copy our arguments for the Ising model by
writing
e−βδσx ,σy = e−β (1 + fxy )
where fxy = eβ(1−δσx ,σy ) − 1 = δσx ,σy (eβ − 1). Then
Y X Y
eβ|BΛ | e−βHΛ (σ) = (1 + δσx ,σy (eβ − 1)) = δσx ,σy (eβ − 1) (4.19)
{x,y}∈BΛ B⊂BΛ {x,y}∈B
We have 0 ≤ fxy ≤ β(1 + fxy ) and the thermodynamic limit and decay of correlations can
be done as in the Ising case
Exercise. Check this!
28
The high temperature expansion has actually an interesting structure. Let us consider
the partition function i.e. (up to a trivial multiplicative factor) the sum of (4.19) over σ.
The set of bonds B defines a graph G with vertex set v(G) = Λ and edge set e(G) = B.
Let Gα be the connected components of G. The Kronecker deltas force all the spins to
be equal in each v(Gα ) and the spin sum factorizes over the v(Gα ). We obtain then
X X
eβ|BΛ | e−βHΛ (σ) = (eβ − 1)|e(G)| q n(G) (4.20)
σ G
where n(G) is the number of connected components of G (note that each x ∈ Λ not
belonging to any of the bonds in B is a connected component Gα with v(Gα ) = {x} and
e(Gα ) = ∅). Eq. (4.20) is called the Fortyuin-Kasteleyn representation for the partition
function of Potts model. It has an interpretation of a percolation model, see [23].
As the Ising model Potts model as a critical (inverse) temperature βc separating the
high temperature phase from the low temperature phase. In d = 2 and q ≤ 4 the
correlation length ξ is infinite at βc and finite elsewhere and the transition at βc is second
order. For q > 4 the transition is 1st order and ξ < ∞. This should be the case in d > 2
and q > 2 as well. This is proven for q large enough in all dimensions, see [28].
b) O(N )-models. Here σx ∈ S N −1 = sphere in RN i.e. σx is a vector ~σx ∈ RN , k~σx k2 = 1.
We take X
H=− ~σx .~σy .
|x−y|=1
This has O(N ) symmetry: Let R ∈ O(N ) i.e. R = N × N orthogonal matrix, RT R = 1|.
Then H(R~σ ) = H(~σ ) where (R~σ )x = R~σx . The N = 2 case is called the XY -model,
N = 3 the classical Heisenberg model. These are very interesting. First, there is no
symmetry breaking when d = 2 : there are no states with h~σx i =6 0. This is called the
Mermin-Wagner theorem [31, 14]. We’ll prove it in section 5 below.
There are infinitely many low-temperature states, parametrized by unit vectors n̂ [18].
For d = 2, N = 2 and N ≥ 3 behave qualitatively differently. For N = 2 (XY -model
or plane rotator), there is a critical temperature βc where ξ becomes ∞: For β < βc
ξ(β) < ∞ and for β ≥ βc , ξ(β) = ∞. The correlation function has a peculiar decay if
β > βc : hσx σy i ∼ |x − y|−a(β) where a(β) depends on β [27, 19].
For d = 2, N ≥ 3, one expects ξ(β) diverge as ecβ when β → ∞ (which is proven as a
lower bound).
29
c) Gauge theories Consider bonds b with orientation b = (x, y) 6= (y, x) and denote
(y, x) = b−1 .
b
x y
The spins are indexed by oriented bonds b and take values σb ∈ SU (N ) (N × N unitary
matrices, det σb = 1). Moreover we let σb−1 = (σb )−1 . Hence the state space is ΩΛ =
(SU (N ))BΛ .
b1
b4 b2
Let p be a plaquette and choose arbitarily an orientation :
b3
Put :
1
Sp = Re Tr σb1 σb2 σb3 σb4 = (Tr σb1 σb2 σb3 σb4 + Tr σb1 · · · σb4 )
2
1
= (Tr σb1 · · · σb4 + Tr (σb1 · · · σb4 )−1 )
2
(since Tr U = Tr U + = Tr U −1 for U ∈ SU (N )) and
X
HΛ = Sp .
p⊂Λ
This Hamiltonian has a huge symmetry, the gauge symmetry : Let GΛ = ×x∈Λ SU (N )
i.e. g ∈ GΛ is a map Λ 3 x → gx ∈ SU (N ) i.e. at each point x ∈ Λ choose a matrix
gx ∈ SU (N ). GΛ is the gauge group in Λ. It acts on spins as follows. For σ ∈ ΩΛ , g ∈ GΛ ,
let
(U (g)σ)b = gx σb gy−1
where b = (x, y). Then we have
so HΛ (U (g)σ) = HΛ (σ).
x1 x2
x4 x3
This is a big symmetry and in all dimensions it remains unbroken. The reason being
that a local symmetry such as U (g), g ∈ GΛ does not change the Hamiltonian, even with
arbitrary boundary conditions, provided those b.c. are on the boundary of Λ0 , with Λ
30
strictly included in Λ0 . Then, the fact that the gauge symmetry remains unbroken is an
immediate consequence of the DLR equation (5.4) below.
2. The last example had a 4-spin interaction. Let us now generalize to give the general
formalism of bounded spin models. Thus, let M be a compact metric space, define as
d
before the infinite configuration space Ω = MZ , and, in a finite volume ΩΛ = MΛ . Ω is
still a compact metric space (by Tychonov), with metric
X
D(σ, σ 0 ) = 2−|x| d(σx , σx0 )
x∈Zd
Actually this is the free boundary condition Hamiltonian. (Note that e.g. in gauge the-
ories we rather consider spins on bonds, not sites so these definitions should be changed
appropriately).
A priori measure : Let ν be a Borel probability measure on M. Thus, in Ising model
we had ν(−1) = ν(1) = 21 , in O(N )-model ν is the uniform measure on S N −1 and in
SU (N )-Gauge theories : ν is the Haar measure on SU (N ). The a priori measure in MΛ
is the product measure
Y
dνΛ (σ) = dν(σx ) (4.21)
x∈Λ
31
where ZΛ = ΩΛ e−βHΛ (σ) dνΛ (σ). For more general boundary conditions we pick as before
R
by compactness of MX .
Exercise. Let d = 1 and Φ finite range R. Show that for all β the correlation functions
hF (σ)iσ̄Λ have thermodynamic limit which is independent on σ̄. Here F depends on finitely
many spins. Moreover show the correlations decay exponentially. Hint: proceed as in 1d
Ising model by introducing a transfer matrix Tσ,σ0 where σ, σ 0 ∈ MR . For simplicity you
may take M a finite set.
Sometimes finite range is not general enough. There are various classes of potentials,
that allow for the Λ % Zd limit of the free energy, the convergence of high temperature
expansions etc.
using the definition (4.25) i.e. the interaction energy of σ0 with all other spins is finite.
X
Note that translation invariance implies kΦk0 = kΦX k for any y ∈ Zd .
X⊂Zd
y∈X
32
Lemma 4.12 Let Φ ∈ B0 . Then
Proof We have
X X X 1 X X
ΦX (σ) = ΦX (σ) ≤ kΦX k = |Λ| kΦk0
|X ∩ Λ|
X:X∩Λ6=∅ y∈Λ X ⊂ Zd y∈Λ d
X ⊂Z
y∈X y∈X
We have then
Theorem 4.13 Let Φ ∈ B0 . Then the free energy
1
F = − lim log ZΛσ̄L (ΛL = L − box)
L→∞ Ld β
Λ
1◦ dist (y, Λc ) ≤ L0
2◦ dist (y, Λc ) ≥ L0
L0
L
Then the contribution of 1◦ to (4.26) is bounded by
1
CL0 Ld−1 kΦk0 ≤ 2 Ld (4.27)
if we take CL0 /L ≤ 12 .
33
The second sum with the constraint 2◦ is bounded by
X
2Ld kΦX k ≡ Ld δ(L0 ).
X:0∈X, d(X)≥L0
P
Since the sum X:0∈X kΦX k converges, δ(L0 ) → 0 as L0 → ∞. Hence pick L0 so that
δ(L0 ) ≤ 21 and then L so that (4.27) holds. Then we have
(4.26) ≤ Ld
and thus
d 0 0 d
e−βL ZΛσ̄ ≤ ZΛσ̄ ≤ ZΛσ̄ eβL .
So, for L > L()
0
L−d | log ZΛσ̄L − log ZΛσ̄L | ≤ β.
Since this holds for all > 0
0
lim L−d | log ZΛσ̄L − log ZΛσ̄L | = 0.
L→∞
0
Obviously we may replace ZΛσ̄L by the free boundary condition version above.
b) Existence. It suffices to consider the free boundary condition theory. Let > 0. Then
∃R such that kΦ − Φ(R) k ≤ where
(
(R) ΦX diam (X) < R
ΦX = .
0 diam (X) ≥ R
So
|HΛ,Φ − HΛ,Φ(R) | ≤ |Λ|
and thus 1 1
log ZΛ,Φ − log ZΛ,Φ(R) ≤ β.
|Λ| |Λ|
Thus, it suffices to consider Φ of finite range, say R. Let L2 > L1 > R. We will compare
L−d −d
2 log ZΛL2 to L1 log ZΛL1 . Write L2 = nL1 + L with L < L1 and n ∈ N. Then
ΛL2 is a union of nd disjoint L1 -boxes Λ(i) with i = 1, . . . , nd and a region Λ of volume
|Λ| ≤ CL1 Ld−1 (i)
2 . Let σ|Λ(i) ≡ σ . We get
nd
X
0 d −1 −1
HΛL (σ) − HΛ(i) (σ ) ≤ CRLd−1
(i) d d−1
1 n kΦk0 + CL1 L2 kΦk0 ≤ C L2 (L1 + L1 L2 )
i=1
where HΛ(i) is the free boundary condition Hamiltonian. The first term in the middle
expression bounds the contribution of ΦX with X ∩ Λ(i) 6= ∅, X ∩ (Λ(i) )c 6= 0 for some i
and the second one the ΦX with X ∩ Λ 6= ∅. Thus,
R −β P H (σ(i) )
0 d −1 −1 e Λ(i) dνΛL2 0 d −1 −1
e−C βL2 (L1 +L1 L2 ) ≤ R −βH (σ) ≤ eC βL2 (L1 +L1 L2 ) . (4.28)
e ΛL
dνΛL2
34
Qnd (i)
Now using dνΛL2 (σ) = dνΛ (σΛ ) i=1 dνΛL(i) (σ ) we get
Z
HΛ(i) (σ (i) ) d
P
e−β dνΛL2 = (ZΛL1 )n
we get
1 nd
d log ZΛL2 − d log ZΛL1 ≤ C 0 β(L−1 −1
1 + L1 L2 )
L2 L2
Since | Ln2 − 1
L1
| ≤ 1
L2
this implies
1 1
d log ZΛL2 − d log ZΛL1 ≤ C 00 β(L−1 −1
1 + L1 L2 )
L2 L1
and then L1 → ∞
C
provided Jxy has enough decay as |x − y| → ∞ : e.g. |Jxy | ≤ [|x−y|+1] d+ will do.
There is a very important system where Φ 6∈ B0 : the Coulomb gas. It is like (4.29), σx
= charge at x i.e. + or − charges and
X 1
HΛ (σ) = σx σy .
|x − y|
x, y ∈ Λ
x 6= y
(
1
|x−y|
if X = (x, y) X X
1
Hence ΦX = and ΦX = |y|
= ∞. Nevertheless, the Λ % Zd
0 otherwise 0∈X y6=0
−βH
limit exists. The secret is in signs : note that e gets very small contribution from σx =
+1 ∀x or σx = −1 ∀x. The nonzero contribution comes from alternating configurations
(neutral ones : σ is charge). See Simon [42], p. 121.
35
Let us next consider High-temperature uniqueness. Let us proceed as in the Ising
model case. We first write
Y Y Y
e−βHΛ = e−βΦX = e−βkΦX k eβ Φ̃X
X⊂Λ X⊂Λ X⊂Λ
(4.30)
where we note
Φ̃X = kΦX k − ΦX ≥ 0.
Disregarding the σ-independent constant we then expand
Y Y XY
eβ Φ̃X = (1 + fX ) = fXα
X⊂Λ X⊂Λ {Xα } α
where the sum is over all families {Xα } of subsets of Λ and fX = eβ Φ̃X − 1 satisfies
0 ≤ fX ≤ β Φ̃X (1 + fX ).
We can now proceed as in the proof of Theorem 4.1.(a) to study the σ̄ or Λ dependence
of hFA (σ)iσ̄Λ where FA is a continuous function on MA . For instance eq. (4.7) will be
replaced essentially by
XY n
βkΦXi k
P i=1
where P = {X1 , . . . , Xn } is a connected path of sets from A to Λc i.e. A ∩ X1 =
6 ∅,
c
Xi ∩ Xi+1 6= ∅ and Xn ∩ Λ 6= ∅. To control this sum we need to assume a bit more from
the potential, namely that
X
kΦk1 := |X|kΦX k < ∞.
0∈X
Exercise. Show that this condition allows to bound the above sum if β is small enough
and show it tends to zero as d(A, Λc ) → ∞.
Indeed one can prove
Proposition 4.14. Let kΦk1 < ∞. Then, for β small enough the correlation functions
hFA (σ)iΛ converge as Λ ↑ Zd to a σ̄-independent limit and
|hFA GB i − hFA ihGB i| → 0
as dist (A, B) → ∞.
See [5] for more details. Assuming more on the decay of ΦX allows one to get infor-
mation about the decay of correlations. Eg. if we assume that for some α > 0
X
kΦk2 := eα|X| kΦX k < ∞
0∈X
0
|hFA GB i − hFA ihGB i ≤ Ce−α d(A,B)
with 0 < α < α.
36
5 Gibbs States and DLR Equations
Finally, let us define what one means by Gibbs states in the general framework. Recall
the discussion in Section 2 of limits of finite volume correlations. That can be repeated
in the general framework. Thus if we knew that for all finite X ⊂ Zd and all f ∈ C(MX )
Λn
Λ
Write
X X
HΛσ̄ n (σ) = ΦX (σ ∨ σ̄) = HΛσ∨σ̄ (σ) + ΦX (σ ∨ σ̄) ≡ HΛσ∨σ̄ + H̃
X∩Λn 6=∅ X ⊂ Λc
X ∩ Λn 6= ∅
37
But Z Z
σ∨σ̄ −β H̃ σ̄
ZΛσ∨σ̄ e−β H̃ = dνΛ (σ)e −βHΛ
e = dνΛ (σ)e−βHΛn (σ)
so (5.3) becomes Z Z
0 0
Eσ̄Λn (f ) = dµσ̄Λn (σ) dµσ∨σ̄
Λ (σ )f (σΛ )
as claimed.
Another way to say this is the following. The measure µσ̄Λn projects to a measure
(µσ̄Λn )Λ on MΛ by the formula
Z Z
f dµΛn = f d(µσ̄Λn )Λ
σ̄
38
c
for all continuous f on MΛ (here σ̄Λc means just the restriction of the configuration σ̄
to Λc , where f is defined). Thus, if g ∈ C(MΛ ), then
Z Z
gdµ = gdµσ̄Λ ≡ hgiσ̄Λ
σ̄,Λ
where g(σΛc ) = f dµσΛΛc is easily seen to be a continuous function (check!). Hence the
R
Remarks. In the general theory of Gibbs states one can now show the following.
Let Φ ∈ B0 and GΦ be the set of Gibbs states of Φ. Then :
a) GΦ is a convex set : µ1 , µ2 ∈ GΦ ⇒
Example. Ising model : the translation invariant pure phases are µ+ and µ− ,
general translation invariant Gibbs state µ = sµ+ + (1 − s)µ− . For example µf ree =
µperiodic = 21 (µ+ + µ− ) (Why ?).
39
d) Pure phases have correlations that tend to 0 at infinity:
Theorem 5.2 Let µ be a Gibbs measure for the XY -model. Then µ is invariant under
a global rotation of the spins. I.e. for any f ∈ C(Ω) and R ∈ SO(2),
Z Z
f (σ)dµ(σ) = f (Rσ)dµ(σ)
Proof 1. The idea is as follows (we follow here [26], see [42], p. 296). Consider say f (σ)!=
cos ϕ sin ϕ
F (σ0 ), depends on σ0 only. Let R(ϕ) be rotation by angle ϕ, R(ϕ) = .
− sin ϕ cos ϕ
!
cos θx X
Go to angular variables σx = , so H = − cos(θx − θy ), and the a priori
sin θx hxyi
dθx
measure is on [0, 2π].
2π
We want to prove :
hF (θ0 + ϕ)i = hF (θ0 )i ∀φ,
D d E
it suffices to prove F (θ0 + ϕ) = 0 , ∀F smooth. Now h−i is approximatively
dϕ ϕ=0
θ̄
h−iΛ some large Λ, some b.c. θ̄ because Gibbs measures are limits of such measures. Thus
we could change variables
D Eθ̄ D Eθ̄+ϕ
F (θ0 + ϕ) = F (θ0 )
Λ Λ
So the ϕ dependence is now in the b.c. This is tricky to control, so let us try to change
variables by slowly rotating the spins : rotate θ0 + ϕ → θ0 and the rest a bit less such
40
that on ∂Λ there is no rotation : Let g : Λ → R g(0) = 1, g(x) = 0 for x near ∂Λ. Then,
let (τϕ θ)x = θx + g(x)ϕ
D Eσ̄ D Eσ̄
F (θ0 + ϕ) = F ((τϕ θ)0 )
Λ Λ
Z
1 h i Y dθ
x
= σ̄ F (θ0 ) exp −βHΛσ̄ (τ−ϕ θ)
ZΛ 2π
↑
θx → θx − g(x)φ = (τ−ϕ θ)x
σ̄
d σ̄ d σ̄
⇒ h F (θ0 + φ)iΛ = −β H (τ−ϕ θ) F (θ0 )
dϕ ϕ=0 dϕ ϕ=0 Λ Λ
writing
d σ̄ σ̄ d σ̄
β HΛ (τ−ϕ θ)e−βHΛ (τ−ϕ θ) = − e−βHΛ (τ−ϕ θ)
dϕ dϕ
and integrating by parts, we get :
D d 2 E D d2 E
σ̄ σ̄
β H = H (τ−ϕ θ) .
dϕ ϕ=0 Λ dϕ2 ϕ=0 Λ
But
d2 σ̄ d2 X
H (τ −ϕ θ) = − cos θx − θy − ϕ g(x) − g(y)
dϕ2 Λ dϕ2
hxyix∈Λ,y∈Λ
Thus,
D E
d X 2
dϕ ϕ=0 F ≤ C g(x) − g(y) (5.5)
hxyi
x, y ∈ Λ
where C is g-independent. (5.5) holds for all g : Λ → R, g(0) = 1, g = 0 near ∂Λ. Also
(5.5) holds for any Λ (C is Λ-independent).
To make the above argument rigorous, we use the DLR-equation to write
Z Z Z
dF dF σ̄
dµ = dµ (σ) dµ̃Λc (σ̄)
dϕ dϕ Λ
41
i.e. an average over σ̄ of what we did. Then (5.5) follows as above.
and this equals zero. Thus, let h ∈ C0∞ , h(0) = 1 (smooth function h : R2 → R with
compact support). Let the support of h be contained in ΛR (= cube, center 0, side R).
Put gL : Z2 → R , gL (x) = h( R
L
x) (x ∈ Z2 ). Then gL (0) = 1, gL = 0 near ∂ΛL and
2 2
X
2
XX R R
(gL (x) − gL (y)) = h x −h (x − ei )
2 i=1
L L
x∈Z
~x
∇h (x) = − 2+ for |x| ≥ 1 and = 0, |x| < 1, ⇒
|x|
Z Z ∞
2 2
(∇h(u)) d u = 2π2
r−2−2 rdr = π → 0,
1
as → 0
42
We will denote the spin by φ(x) = φx i.e. the spin configuration is φ : Zd → R. Let
us consider a simple Hamiltonian
X X X
HΛ (φ) = − φ(x)φ(y) + a φ2 (x) + λ φ4 (x) (6.1)
{x,y}∈BΛ x∈Λ x∈Λ
which again has the φ → −φ symmetry, and consider the probability measure
1 −βHΛ (φ) Y
e dφ(x) (6.2)
ZΛ x∈Λ
r
Call a − d = 2
and so
1 X rX 2 X 1 X
HΛ (φ) = (φx − φy )2 + φx + λ φ4x + nx φ2x . (6.3)
2 2 2 x∈∂Λ
{x,y}∈BΛ
It will be much more convenient to work with the periodic boundary conditions instead
of the free ones above. In that case the last term is missing from (6.3) and BΛ contains
also the bonds joining opposite faces of the cube ΛL .
Remark. The Ising model is a limit of this model : take
r r 2
= −2λ so φ + λφ4x = λ(φ2x − 1)2 − λ
2 2 x
q
λ −λ(φ2 −1)2
and π
e −→
λ→∞ δ(φ2 − 1) (in the sense of distributions).
7 Gaussian Integrals
7.1 Definitions and elementary properties
Consider first the case λ = 0. Then our measure is Gaussian.
43
Definition 7.1 Let A be a real symmetric n×n matrix which is (strictly) positive definite
(i.e. all eigenvalues > 0 i.e. (φ, Aφ) > 0 ∀φ ∈ Rn , φ 6= 0). The Gaussian measure on Rn
with covariance A−1 and mean 0 is the probability measure
1 − 1 (φ,Aφ)
e 2 dµ(φ) =
Dφ
Z
Q X
where φ = (φ1 , · · · , φn ), Dφ = i dφi , (φ, Aφ) = φi Aij φj .
ij
n
Thus, each φi : R → R given by φi (φ) = φi is a random variable, with mean 0 and
variance hφ2i i = (A−1 )ii . Also, (φ, f ) is a random variable, with variance (f, A−1 f ).
Some calculations
1. Consider the partition function
Z
1
Z= e− 2 (φ,Aφ) Dφ.
Rn
2. The correlations functions h kα=1 φiα i where iα ∈ {i, . . . , n} (note: several iα may be
Q
the same) can be computed using the generating function : Let f ∈ Rn . Define
Z
1 1
S(f ) = he(φ,f )
i= e− 2 (φ,Aφ)+(φ,f ) Dφ,
Z
k k
Y Y ∂
h φiα i = S(f ).
α=1 α=1
∂fiα f =0
But
Z
1 1 −1 −1 1 −1
S(f ) = e− 2 (φ−A f,A(φ−A f ))+ 2 (f,A f ) Dφ
Z
Z
1
(f,A−1 f ) 1 1 1 −1
= e 2 e− 2 (Ψ,AΨ) DΨ = e 2 (f,A f )
Z
using the change of variables Ψ = φ − A−1 f . Hence, say the 2-point function,
∂ 2 1 (f,A−1 f )
hφi φj i = e2 = (A−1 )ij .
∂fi ∂fj 0
44
In general,
k
Y 0 k odd
m
h φiα i = Y ∂ 1 1 X Y
−1
(f, A f ) = (A−1 )iα iβ k = 2m
∂fiα 0 2 m!
α=1
P {α,β}∈P
where P is a pairing of the set {1, 2, · · · , 2m} i.e. a partition into m sets of size 2. The
sum runs through all such pairings.
Example For n = 4 there are three pairings {{1, 2}, {3, 4}}, {{1, 3}, {2, 4}} and {{1, 4},
{2, 3}}. In general, the number of pairings is 2(2m)!
m m! (show !).
Thus
2m
Y X Y
h φ iα i = hφiα φiβ i
α=1 P {α,β}∈P
Concretely
X
−(∆per φ)x = (φx − φx+u ) (7.2)
|u|=1
where the addition is modulo L. For example in d = 1 we have (−∆φ)x = 2φx − φx−1 −
φx+1 .
d
(7.2) defines as well an operator ∆ in infinite volume i.e. for φ ∈ RZ . Then ∆ is self
d
adjoint on `2 (Zd ) ⊂ RZ
X
`2 (Zd ) = {φ : Zd → C | |φx |2 < ∞}.
x∈Zd
P
Indeed, ∆ is bounded: k∆k ≤ 2d (show!) and in the scalar product (φ, ψ) = x∈Zd φ̄x ψx
we have (φ, ∆ψ) = (∆φ, ψ).
45
d
The periodic Laplacean can be defined in RZ as well. For this consider the set
d
FLper ⊂ RZ of periodic functions : [we use φx or φ(x) below !]
Clearly FLper can be identified with RΛL , namely φ is determined by φ|ΛL , ΛL = {x|xi ∈
{0, . . . , L − 1}}. Clearly ∆ : FLper → FLper so the periodic Laplacean ∆per is the corre-
sponding matrix in RΛL .
d
We will now diagonalize ∆per and ∆ by Fourier series. Consider p ∈ [−π, π]d and φp ∈ RZ
φp (x) = eipx (where px := α pα xα ) . Then,
P
X X
∆φp (x) = (eip(x+u) − eipx ) = (eipu − 1)eipx = −µ(p)φp (x)
|u|=1 |u|=1
where
d
X
µ(p) = 2 (1 − cos pµ ).
µ=1
and
X
ei(p−q)x = Ld δpq for p, q ∈ BL (7.5)
x∈ΛL
Thus, the correlations of the λ = 0 GL-model are given in terms of the (periodic b.c.)
Green’s function GL .
How about L → ∞ ? In (7.6) we have a Riemann sum over cells of size ( 2π
L
)d . So, in
the limit,
dd p eip(x−y)
Z
lim GL (x − y) = d
≡ G(x − y). (7.7)
L→∞ [−π,π]d (2π) µ(p) + r
46
This equals of course the kernel of the operator (−∆ + r)−1 defined on `2 (Zd ). Indeed, on
`2 (Zd ) the operator −∆ has spectrum {µ(p)|p ∈ [−π, π]d } and we can diagonalize it by
Fourier series. If f ∈ `2 (Zd ), let
X
fˆ(p) = e−ipx f (x).
x∈Zd
Proposition 7.5 Let µL be the Gaussian measure on RΛL , with covariance (−∆per +r)−1 .
Then
Z Y XY
−→
φ(xα )dµL (φ) L→∞ G(xα − xβ ). (7.8)
P hαβi
47
which always exists by dominated convergence theorem.
W (f ) has some obvious and less obvious properties:
R
a) |W (f )| ≤ 1dµ = 1.
b) W (0) = 1.
d) Let zα ∈ C, fα ∈ Rn , α = 1, · · · , N . Then,
X Z X Z X
i(φ,fα ) −i(φ,fβ )
zα z̄β W (fα − fβ ) = zα z̄β e e = | zα ei(φ,fα ) |2 dµ ≥ 0.
α,β
For φ, ψ ∈ H put X
(φ, ψ) = φ̄(x)ψ(y)W (x − y)
x,y∈Rn
(this is a finite sum). ( , )is an inner product except that (φ, φ) = 0 does not imply φ = 0.
Let H0 = {φ|(φ, φ) = 0}; H0 is a subspace of H. Put H e = H/H0 = equivalence classes
[φ] = [φ + φ0 ], φ0 ∈ H0 . Then (H, e (·, ·)) is an inner product space which can be completed
into a Hilbert space.
Let for t ∈ Rn , Ut : H → H (Ut ψ)(x) = ψ(x + t). Clearly, Ut : H0 → H0 so one
defines Uet : H e→H e by Uet [φ] = [Ut φ]. (Uet φ, U
et ψ) = (φ, ψ) so U
et is an isometry. We have
U
et+s = U et U e0 = 1|. Also t → U
es , U et is strongly continuous (i.e. ∀ψ ∈ H e t→U et ψ ∈ H
e is
continuous from Rn → H) e because W is continuous:
kU es ψk2 = 2(ψ, ψ) − (U
et ψ − U es ψ) − (U
et ψ, U et ψ) = [2(ψ, ψ) − (ψ, U
es ψ, U et−s ψ) − (ψ, U
es−t ψ)]
This equals
X
ψ̄(x)ψ(y)[2W (x − y) − W (x − y + t − s) − W (x − y + s − t)]
x,y
48
which tends to 0 as t → s. We have now verified the assumptions of
Proof See Reed-Simon [35], vol. 1, Th. VIII.12. [The heuristics behind the proof
is that, if H were finite dimensional, we could easily prove that Ut is differentiable in t,
∂Ut P
|t=0 = iAj , Aj Hermitean, [Ai , Aj ] = 0 and Ut = ei tj Aj . Diagonalize all Aj simulta-
∂tj
(j)
λ1 0 n
.
X
neously ⇒ Aj =
. . ~
⇒ dµφ (λ) =
δ(~λ(i) − ~λ)|φi |2 ].
(j) i=1
0 λn
To finish the proof of the Theorem, using Stone’s Theorem, take φ̃ ∈ H̃
(
1 x=0
φ̃ = [φ], φ(x) =
0 x 6= 0
X Z
W (t) = δ(x)δ(y + t)W (x − y) = (φ̃, Ũt φ̃) = ei(λ,t) dµφ̃ (λ)
x,y∈Rn
other variables :
49
for all Borel sets A ⊂ Rn . A natural σ-algebra on RN is the σ-algebra generated by the
cylinder sets. Denote by B Borel sigma-algebra on R and by B n the one on Rn .
Definition 8.1 A cylinder set on RN is a set of the form SB1 ,··· ,Bn = {x ∈ RN |xi ∈ Bi , i =
1, · · · , n, Bi ∈ B, n < ∞}. The σ-algebra of subsets of RN generated by cylinder sets is
denoded by B N . A measure µ defined on B N is called a cylinder measure. We have then
the
Kolmogorov’s extension theorem. Let {µn }∞ n=1 be a consistent family of measures on
n
B . Then there is a unique cylinder measure µ on B N such that
Proof. Main problem is countable additivity. We can do this with Riesz theorem by the
following trick.
Let Ṙ be the one-point compactification of R (i.e. Ṙ = R ∪ {∞} and open sets are
open sets in R and sets of the form A ∪ {∞} where A ⊂ R is open and Ac ⊂ [−n, n] some
n < ∞, Ṙ is easily seen to be compact). Let
M = ×∞ N
i=1 Ṙ = Ṙ .
As it turns out, we can embed all interesting function (and distribution) spaces into
RN . Let us start with two subsets :
2
Borel sets of ṘN = smallest σ-algebra containing open sets and cylinder sets are generated by open
sets.
50
8.3 Minlos Theorem : Measures on s0
Let ∞
X
sm = {x ∈ R | N
n2m |xn |2 ≡ kxk2m < ∞}
n=1
for m ∈ Z. Let \ [
s= sm s0 = sm
m∈Z m∈Z
Lemma 8.2 The space s with the above topology is a complete metrizable space, i.e. s
has a complete metric d such that d gives the same topology as above.
Then (Exercise) : Show that this distance defines the same topology as above and that
s is complete (Cauchy sequences converge).
Remark. s is an example of a Frechet space i.e. a complete locally convex metric space,
and which is not a Banach space.
Lemma 8.3 s0 is the dual space of s i.e. the space of continuous linear functionals on s.
Proof Let y ∈ s0 . y defines a continuous linear functional `y on s :
∞
X
`y (x) ≡ (y, x) ≡ yn xn .
n=1
Indeed, this converges since |yn | ≤ Cnm some m < ∞ and |xn | ≤ Cm nm for all m. To
prove continuity, let d(x(k) , x) → 0 as k → ∞. Then, by Schwartz’ inequality,
∞
(k)
X X 1
|`y (x ) − `y (x)| ≤ Cnm |xn(k) − xn | = C nm+1 |x(k)
n − xn |
n=1
n
X 1 1/2
≤ C kx(k) − xkm+1
n2
≤ C 0 2m+1 d(x(k) , x) → 0 as k → ∞.
51
Conversely, let ` : s → R be a continuous linear map. Then there exists C < ∞, m < ∞
such that
|`(z)| ≤ Ckzkm (8.12)
for all z. Indeed, by continuity there exists a neighbourhood U of zero in s such that
|`(x)| < 1, for x ∈ U . On the other hand U contains a set {z| kzkm < δ} some m, δ.
Hence, given a z ∈ s we have δ kzkz m ∈ U and so |`(δ kzkz m )| < 1 i.e. |`(z)| ≤ δ −1 kzkm ).
(k)
Let now e(k) ∈ s be given by en = δnk . Put yn = `(e(n) ). By (8.12)
so y ∈ s0 . Given x ∈ s, let
n
X
x(n) = e(k) xk .
k=1
52
Proof. Let W be given. Let u ∈ Rn , put x = (u1 , · · · , un , 0, 0, · · · ) ∈ s. Then W (x) ≡
fn (u) is a function of positive type in Rn ⇒ ∃ probability measure µn on Rn such that
W
Z
Wn (u) =
f ei(y,u) dµn (u)
Rn
{µn }∞
n=1 are consistent since Wn+m ((u1 · · · un , 0, · · · , 0)) = Wn (u) and Wn uniquely deter-
f f f
mines µn (Why ?). Hence ∃µ on RN such that
Z
W (x) = ei(y,x) dµ(y) (8.13)
RN
for all x ∈ s with only finitely many non zero xi . We need to show : µ(s0 ) = 1. In that
case the integral can be restricted to s0 and (8.13) holds for all x ∈ s (because both sides
are continuous in x and the set of x with a finite number of nonzero xi is dense in s).
[∞
µ(s0 ) = 1 : Recall that s0 = sm with
m=−∞
∞
X
sm = {y| n2m yn2 ≡ kyk2m < ∞}.
n=1
µ(sm ) ≥ 1 − . (8.14)
(
−α kyk2m 1 y ∈ sm
since e 2 % χ(y ∈ sm ) ≡
0 y∈6 sm , i.e. kykm = ∞
Z N
!
α X 2m 2
We consider exp − n yn dµ. Choose m:
2 n=1
(recall that W (0) = 1). (Note that m here tends to be < 0 !).
Then, for all x ∈ s,
2
Re W (x) ≥ 1 − − kxk2−m−1 (8.16)
δ
53
since if kxk2−m−1 < δ, (8.16) holds by (8.15) and if kxk2−m−1 > δ, (8.16) holds because
|W (x)| ≤ W (0) = 1 implies Re W (x) ≥ −1.
Now write ! Z
N
α X 2m 2
exp − n yn = ei(y,x) dν(x)
2 n=1 R N
N
Y
2m −1/2
x2n
dν(x) = (2παn ) exp − dxn
n=1
2αn2m
to get
Z N
! Z Z N
! !
α X 2m 2 X
exp − n yn dµ = exp i yi xi dµ(y) dν(x)
2 n=1 RN i=1
Z Z
= W (x)dν(x) = Re W (x)dν(x)
RN
But
∞
Z N Z N
!
X X X 1
kxk2−m−1 dν = n−2m−2 x2n dν(x) = n−2m−2 αn2m ≤ α.
RN n−1 n=1 n=1
n2
Thus !
Z N
α X 2m 2 cα
lim exp − n yn dµ ≥ 1 − −
N →∞ 2 n=1 δ
and our claim follows
Example. Gaussian measures. Let C be a matrix Cij i, j = 1, 2, · · · with |Cij | ≤ aim · j m
X
some a < ∞, m < ∞ and (x, Cx) = xi xj Cij > 0, ∀x 6= 0, for x ∈ s (note (x, Cx)
−n
makes sense since |xi | < bn i all n). Then the function
1
WC (x) = e− 2 (x,Cx)
is of positive type.
Proof W is continuous since |(x, Cx)| ≤ Akxk2n for some n. Hence WC (x) = lim WC (x(k) ),
( k→∞
(k) x i i ≤ k
xi = . But
0 i>k
k
! Z k
!
1 X X
WC (x(k) ) = exp − xi xj Cij = exp i xi yi dµk (y)
2 ij=1 Rk i=1
54
where µk is the Gaussian measure on Rk with covariance matrix Cij , i, j = 1, · · · , k. By
assumption, this is a positive k × k matrix. Hence, let WCk (x) = WC (x(k) ) x ∈ s so WCk is
of positive type and WCk (x) → WC (x) ∀x ∈ s → WC (x) is of positive type.
i.e. f and its derivatives decay faster than any power. Let us define (“creation and
annihilation operators”)
1 d + 1 d
a= √ x+ , a = √ x−
2 dx 2 dx
2
and N = a+ a = − 21 dx
d 1 2 1
2 + 2 x − 2 . For f ∈ S(R), let
Z 1/2
m 2
kf km = |(N + 1) f | dx .
R
N (m, ) = {f ∈ S| kf km < }
55
These satisfy : Z ∞
(φn , φm ) ≡ φn φm dx = δnm
−∞
i.e. the map f → x maps the neighbourhoods to each other. To conclude, we need to
show that this map is a bijection.
N
X
B) Surjection, or the map is onto : Let x ∈ s. Then we need to show that xn φn−1 con-
n=1
verges as N → ∞ to an element f of S with xn = (f, φn−1 ). The sequence N
P
n=1 xn φn−1
N
X N
X
converges in L2 to f ∈ L2 . (N + 1)m xn φn−1 = xn nm φn−1 converges in L2 ⇒ f ∈
n=1 n=1
C ∞ . And kf km = kxkm so f ∈ S.
Definition 8.7 The space of tempered distributions on R is the set S 0 (R) of continuous
linear functionals on S(R).
Let ϕ ∈ S 0 (R). Put yn = ϕ(φn−1 ). Then ϕ continuous means4 : ∃C, m : |ϕ(f )| ≤ Ckf km
4
ϕ continuous : ∀ ∃ N(m, δ) : f ∈ N (m, δ) ⇒ |ϕ(f )| < i.e. kf km < δ ⇒ |ϕ(f )| < . Thus for any
f ∈ S, |ϕ(f )| = |ϕ kffkm δ | kfδkm ≤ δ kf km = Ckf km .
56
∀f ∈ S ⇒ |yn | = |ϕ(φn−1 )| ≤ Ckφn−1 km = Cnm . Hence y ∈ s0 . Conversely, y ∈ s0 gives
∞
X
0
rise to ϕy ∈ S (R) by ϕy (f ) = yn (f, φn−1 ). Indeed, from above, xn = (f, φn ) is in s so
n=1
|ϕy (f )| ≤ Ckxkm for some m and this = Ckf km .
2
Remark 1 We see from above that k · km is a norm on S : kfkm = 0 ⇒ (f, φn ) = 0∀n P
⇒
2 + m + m + 2m
f = 0. Moreover, since kf km = (a a + 1) f, (a a + 1) f = f, (a a + 1) f =
d
of f, (a+ a)k f and (a+ a)k f = (Polynomial in x and dx ) f , we get
X dβ
kf km ≤ Cm sup |xα f (x)| ≡ Cm kf k(m)
α,β≤2m
x dxβ
and conversely (show !). So, ϕ ∈ S 0 ⇐⇒ ∃C, N : |ϕ(f )| ≤ Ckf k(N ) . This is in practice
useful, see Examples 1,2 below.
since
Z Z
ϕ(x)f (x)dx ≤ C (1 + |x|)m |f (x)|dx ≤
Z
1
C (1 + |x|)m+2 |f (x)| dx ≤ C 0 sup(1 + |x|)m+2 |f (x)| ≤ C 00 kf k(m+2)
[1 + |x|]2 x
so ϕ ∈ S 0 .
57
Definition 8.8 1. The cylinder set σ-algebra in S 0 (R) is the smallest σ-algebra containing
all sets
A(f, B) = {ϕ ∈ S 0 (R)|ϕ(f ) ∈ B}
for f ∈ S, B Borel in R. It is just the image of our old cylinder algebra on s0 to S 0 by the
map constructed above. Indeed, the cylinder σ-algebra C in s0 is the smallest σ-algebra
of subsets of s0 that contains the sets σn (B) = {y ∈ s0 |yn ∈ B} where B is Borel in R.
I.e. it is the smallest σ-algebra such that functions
pn : s0 → R , pn (y) = yn
ous bilinear function on S(Rn ) × S(Rn ), with C(f, f ) > 0 for f 6= 0. Important examples
are given by integral kernels
Z Z
C(f, g) = d x dn yC(x, y)f (x)g(y)
n
R
here |C(x, y)|[(1 + |x|)(1 + |y|)]m dn xdn y < ∞ for some m suffices. We are interested in
translation-invariant C’s i.e. C(x, y) = G(x − y). These are most conveniently expressed
in terms of Fourier transform : If f ∈ S(Rn ) we put
Z
fˆ(k) = f (x)e−ikx dn x
and one gets fˆ ∈ S(Rn ). [This is because (xα Dβ f )∧ (k) = iα+β Dα k β fˆ(k), do the details !]
Inverse :
dn k
Z
ikx ˆ
f (x) = e f (k) .
(2π)n
58
Then we have
dn k
Z Z
2 n
|f (x)| d x = |fˆ(k)|2
(2π)n
Z
(f ∗ g)(x) ≡ f (x − y)g(y)dn y
and
dn k
Z Z
f (x)G(x − y)f (y)d xd y = n n
|fˆ(k)|2 G(k)
b
(2π)n
which holds for f, G ∈ S to start with, but extends to much more general G’s.
Example The free Euclidean field with mass m is the Gaussian measure µG where
1
G(k)
b = (k ∈ Rn ).
k2 + m2
dn k
|fˆ(k)|2
R
Clearly S 3 f 7→ is continuous, so
(k 2 + m2 )(2π)n
" Z ˆ 2 n #
1 |f (k)| d k
W (f ) = exp −
2 k 2 + m2 (2π)n
e−m|x|
|G(x)| ≤ C n−1 . (8.18)
|x| 2
To see this, let us consider a ultraviolet cutoff : let Λ > 0 and define
1 Λ2
G
bΛ (k) = .
k 2 + m2 k 2 + Λ2
bΛ (k) −→ G(k)
This is called Pauli-Villars cutoff. Note that G b pointwise, and as elements
Λ→∞
of S 0 .
59
So, by rotational symmetry,
dn k eik1 |x| Λ2
Z
GΛ (x) =
(2π)n k 2 + m2 k 2 + Λ2
So the k1 integral :
q q
k + m = (k1 + i ~k 2 + m2 )(k1 − i ~k 2 + m2 ),
2 2
q q
k + Λ = (k1 + i k + Λ )(k1 − i ~k 2 + Λ2 ),
2 2 ~ 2 2
~k = (k2 , k3 , · · · ).
By Cauchy:
" √ √ #
~2 ~2
dn−1~k e−|x| k +m
2 2
Λ2 e−|x| k +Λ
Z
1
GΛ (x) = p + p
(2π)n−1 2 ~k 2 + m2 m2 + Λ2 2 ~k 2 + Λ2 m2 + Λ2
G(x) = p ;
(2π)n−1 2 ~k 2 + m2
p
this is a L1 function satisfying (8.18). To see this, expand ~k 2 + m2 = m + O(~k 2 ) for
small ~k, and change variables |x|~k = ~k 0 .
p
Note that, a priori, for ϕ ∈ S 0 , ϕ(x) is not defined, only ϕ(f ), f ∈ S. These are mea-
surable functions in S 0 and actually Gaussian random variables on the probability space
(S 0 (Rn ), µ):
1 2
heit(ϕ,f ) i = e− 2 t (f,Gf )
so the variance of (ϕ, f ) is (f, Gf ). Let fi ∈ S, i = 1, · · · , N . Then xi ≡ ϕ(fi ) are jointly
Gaussian :
P P 1 P
hei ti xi i = heiϕ( ti fi ) i = e− 2 ij ti tj (fi ,Gfj )
i.e. x = (x1 , · · · , xN ) ∈ RN is Gaussian with covariance matrix A−1 ij = (fi , Gfj ). Hence
e.g.
Z Z
d d −i[tϕ(f )+sϕ(g)]
ϕ(f )ϕ(g)dµG (ϕ) = − he i = (f, Gg) = dn xdn yf (x)G(x − y)g(y)
dt 0 ds 0
and, more generally,
2N
Z Y X Y
ϕ(fi )dµG (ϕ) = (fi , Gfj )
i=1 P hiji∈P
Z X Y
= dn x1 · · · dn x2N f1 (x1 ) · · · fN (xN ) G(xi − xj )
P hiji∈P
60
Hence we may use the notation
Z X Y
ϕ(x1 ) · · · ϕ(x2N )dµG (ϕ) = G(xi − xj )
P hiji∈P
Remark Indeed, let us try to see what ϕ(x)2 would be. Formally
Z Z
2
ϕ(x) dµG (ϕ) = lim ϕ(x)ϕ(y)dµG = lim G(x − y) = G(0) = ∞.
x→y x→y
This is a first instance of an ultraviolet divergence in quantum field theory. Our G is such
that ϕ(x) is not a nice random variable (it is a “distribution valued random variable”).
We’ll return to this.
eip(x−y) dd p
Z Z
C(x − y) = φ(x)φ(y)dµC (φ) =
[−π,π]d µ(p) + r (2π)d
d
where x, y ∈ Zd , φ ∈ RZ .
We have
0 ≤ C(x − y) ≤ Ce−|x−y|/ξ ξ > 0 , ξ → ∞ as r → 0.
Proof Let R = max |xi − yi |. We may assume R = |x1 − y1 | = x1 − y1 .
i=1,··· ,d
!
π
dd−1 p~
Z Z
1 dp 1
C(x − y) = eip1 R ei~p·(~x−~y) .
−π 2(1 − cos p1 ) + µ(~p) + r 2π (2π)d−1
61
B Im p1
A C
Re p1
−π π
Choose so that : Re2(1 − cos(p1 + i)) + r > 0 i.e. 2(1 − cos p1 cosh ) + r > 0 i.e.
√
2 cosh < 2 + r or < cosh−1 (1 + r/2). [So, for r small, one can choose ∼ r ]. So,
since µ(p) ≥ 0,
C(x − y) ≤ Ce−R
a) We showed C(x) ∼ G(x) as |x| → ∞. Here G(x) = 2-point function of the massless
free field, G(p)
b b for d > 2, defines a continuous function e− 12 (f,Gf )
= p12 . Note that G,
on S : We have
Z
Cd
(f, Gg) = dd xdd y f (x)g(y) ≤ Ckf km kgkm for m large enough.
Rd ×Rd |x − y|d−2
62
b) Consider the function
We call this the scaling limit : take the statistical mechanics at the critical point, look
d−2
at long distances, scale φ (by L 2 above) ⇒ get quantum field theory. In this Gaussian
d−2
model we get : let xi 6= xj i 6= j, and ϕL (x) = L 2 φ(Lx) (for x ∈ (L−1 Z)d ).
Then
Z Y 2N Z Y 2N
−→
ϕL (xi )dµC (φ) L→∞ ϕ(xi )dµG (ϕ).
s0 i=1 S 0 i=1
Here the LHS is with a measure defined on variables φ(x), x ∈ Zd , a lattice model and
the RHS with a measure defined on variables ϕ(x) x ∈ Rd , a continuum model (to make
this precise, observe that (L−1 Z)d ) becomes “dense” in Rd as L → ∞).
Cr (x) ∼ |x|2−d
Stated differently, let us put r = L−2 m2 and consider, after the change of variable pm → p,
eipx dd p
Z
d−2
L Cm2 /L2 (Lx) = 2 2 d
[−Lπ,Lπ]d L µ(p/L) + m (2π)
eipx dd p
Z
−→
L→∞ 2 2 d
≡ Gm2 (x).
Rd p + m (2π)
63
Hence
2N
Z Y Z Y
d−2
L 2 φ(Lxi )dµCm2 /L2 (φ) → ϕ(xi )dµGm2 (ϕ)
s0 i=1 S0
i.e. if we approach the critical point r → 0 in a suitable way and scale distances and φ
we get the continuum theory at nonzero m i.e. correlation length < ∞.
A) The measure !
1 X
exp −λ φ(x)4 dµCΛ (φ)
ZeΛ x∈Λ
on R where µCΛ is Gaussian with covariance CΛ = (−∆Λ +r)−1 and ∆Λ is the Laplacean
|Λ|
in Zd with some boundary conditions on ∂Λ (we have considered periodic and free above;
64
one can consider Dirichlet or Neumann on the lattice too). ZeΛ normalizes the total mea-
sure to 1.
B) The measure !
1 X
exp −λ φ(x)4 dµC (φ)
ZΛ x∈Λ
X
Apply it to the random variable x = λ φ4y to get
y∈Λ
Z !
X X
exp −λ φ4y dµC ≥ exp[−λ hφ4y i]
y∈Λ y∈Λ
where 2
dd p
Z Z
1
hφ4y i ≡ dµC (φ)φ4y = 3C(0) = 3 2
µ(p) + r (2π)d
from our rules for Gaussian integrals. Also, since hφ4y i > 0, we get
exp[−3λ|Λ|C(0)2 ] ≤ ZΛ ≤ 1. (10.1)
Note how the extensivity of ZΛ (= eO(Λ) ) is visible here. Also, note that the free energy :
1
−3λC(0)2 ≤ |Λ| log ZΛ ≤ 0, uniformly in Λ. Similar inequalities hold in the case A) above.
65
perturbatively in λ. Consider e.g. the pair correlation
!
R X
φ(x)φ(y) exp −λ φ(x)4 dµC (φ)
x∈Λ
hφ(x)φ(y)iΛ ≡ ! ≡ G2 (x, y). (10.2)
R X
exp −λ φ(x)4 dµC (φ)
x∈Λ
It is not hard to see that, for |Λ| < ∞, this is C ∞ in λ. Indeed consider e.g. the
X
φ(x)4 and set Fk (λ) := V k e−λV dµC (φ). Since
R
denominator and let V =
x∈Λ
|Fn−1 (λ + ) − Fn−1 (λ)| ≤ V n−1 |e−V − 1|e−λV ≤ ||V n e−(λ−||)V ≤ ||n!δ −n e−(λ−||−δ)V
and e−(λ−||−δ)V dµC (φ) < ∞ for λ − || − δ ≥ 0 we see that Fn−1 (λ) is differentiable with
R
0
Fn−1 (λ) = Fn (λ) and the latter equals the n:th derivative of the denominator. Proceeding
similarly with the numerator one shows that both numerator and denominator are C ∞
and from (10.1) above, ZΛ > 0. Hence G2 is smooth (of course this way we get terrible
bounds for the derivatives as Λ gets large).
Thus, we have Taylor’s expansion for G2 :
N
X
G2 (x, y) = G2,n (x, y)λn + RN
n=0
where lim λ−N RN (λ) = 0. The Taylor coefficients G2,n have a nice graphical representa-
λ→0
tion which we now derive.
Let us start with small n. Denote (10.2) by N
D
.
R
n = 0. N = φ(x)φ(y)dµC + O(λ), D = 1 + O(λ) so G2,0 = C(x − y) our Gaussian
covariance.
n = 1.
XZ
N = C(x − y) − λ φ(x)φ(y)φ(z)4 dµC
z∈Λ
and XZ
D =1−λ φ(z)4 dµC .
z∈Λ
66
6!
has 23 ·3!
= 15 pairings and it equals
So, X
N = C(x − y)(1 − 3λ|Λ|C(0)2 ) − 12λ C(x − z)C(y − z)C(0)
z∈Λ
and
D = 1 − 3λ|Λ|C(0)2
Hence altogether X
G2,1 = −12C(0) C(x − z)C(z − y).
z∈Λ
Example. In hφ(z)4 φ(x)φ(y)i we had 2 graphs, G1 has lines {z, z}, {z, z}, {x, y} and G2
lines {x, z}, {y, z}, {z, z}. n(G1 ) = 3, n(G2 ) = 12.
Now, once we sum (10.3) over the points z1 , . . . , zN the order of the points zi will not
anymore matter. We get finally the following graphical rules which we state for arbitrary
correlations :
Definition 10.1 A φ4 - graph with 2m external legs, labelled 1, · · · , 2m, and N unlabelled
vertices is a graph on N points (vertices) such that each vertex has 4 lines attached (one
67
line can start and end at a vertex and is counted as two here), 2m lines have one end with
no other lines attached.
Examples.
2m = 2, N = 1
2m = 2, N = 2
2m = 2, N = 3
2m = 4, N = 1
2m = 4, N = 2
More : etc.
, ,
where G run through (2m, n)-graphs, n(G) is the number of pairings giving that G and
the amplitude A(G) corresponding to the graph G is the following expression:
1. Label the n vertices of G as z1 , · · · , zn
2. To each line ` = {u, v} of G put C` = C(u − v).
X Y
A(G) = C` . (10.4)
z1 ,··· ,zn ∈Λ `
We denote for short A by .
68
Example. We get to O(λ) :
Remark. Note the presence of disconnected graphs like that are cancelled upon
normalization.
Homework : Prove :
G2 = − 12λ + λ2 (α +β +γ ) + O(λ3 )
and find α, β, γ
Theorem 10.2
Z
4
P
n φ(x)φ(y)e−λ φ (x) dµC
d n
X
c
Z = (−) n(G)A(G)
dλn λ=0
P 4
e−λ φ (x) dµC G
X
c
where the sum runs over connected graphs.
In the denominator · ·
69
I.e. connected components have either both x and y or neither.
Note that disconnected graphs are here proportional to a power of the volume :
X
= C(x − y) C(0)2 = |Λ|C(0)2 C(x − y)
z∈Λ
R
so hF i0 = F dµC .
Then,
d X X
hF iλ = − hφ(x)4 F iλ + hφ(x)4 iλ hF iλ
dλ x
d X
i.e. hF iλ = − hφ(x)4 ; F iλ where we denote by
dλ x
Then hF ; Gi0 = C(xγ − xδ ) where each pairing P has at least one pair {γ, δ}
P {γ,δ}∈P
such that γ ∈ A, δ ∈ B.
Proof Obvious, since hF ihGi has those pairings where no such pairs occur and hF Gi
has all pairings.
Now (Prove !)
dn n
X
n
hF iλ = (−1) hF ; φ(x1 )4 ; φ(x2 )4 ; · · · ; φ(xn )4 i0
dλ λ=0
x ···x 1 n
where
X Y Y
hF1 ; F2 ; · · · ; FN i0 = (−1)|π|−1 (|π| − 1)! h Fi i0
π α i∈πα
|π|
where π = {πα }α=1 is a partition of {1, · · · , N } into |π| subsets.
70
Example. hF1 ; F2 ; F3 i = hF1 F2 F3 i−hF1 F2 ihF3 i−hF1 F3 ihF2 i−hF2 F3 ihF1 i+2hF1 ihF2 ihF3 i
−1
where for each partition into pairs P , there is a {{γi , δi }}N
i=1 ⊂ P that forms a connected
path, connecting the Ai ’s, i.e. there exists a permutation ρ of the Ai ’s such that
A4
A2
A3
A1
Example
The Lemma yields our theorem since there cannot be a connected graph with only one
leg.
4
Y
Problems. 1. Let G4 (x1 · · · x4 ) = h φ(xi )i and define the connected four point function
i=1
Gc4 (x1 · · · x4 ) = hφ(x1 ); . . . ; φ(x4 )i. Prove that
by showing that XY
GN (x1 · · · xN ) = Gc|I| (xI )
π I∈π
71
and that the perturbation expansion of GcN consists of connected graphs with N legs.
3. Let
Z(J) := he(φ,J) i.
Hence !
N N
Y Y ∂
GN (x1 , . . . , xN ) := h φ(xi )i = Z(J).
∂J(xi )
J=0
i=1 i=1
dn
4. Prove that dλ n log Z is sum of connected vacuum graphs with n vertices. (a vacuum
0
graph is a graph with no legs).
How to calculate the amplitude A(G) of a given graph G? First, let us observe that
in our model G2m,n has a Λ → Zd limit : Clearly it suffices to consider a connected graph
which is an expression (let say m = 1)
X
A(x1 , x2 , z1 · · · zn )
z1 ···zn ∈Λ
and A is a product of C(yi − yj ) where the y’s are x’s or z’s. Use |C(x)| ≤ const. e−|x|/ξ ,
and note that, since G is connected, there exists a connected tree graph in G containing
all the vertices and all the legs (a tree graph has no loops).
Example.
z3
z2
−→ x z1 y
X
Now do the zi sums using C(zi − zj )≤ const (independent on Λ) starting at ends of
zj ∈Λ
branches (e.g. z3 above).
72
10.3 Momentum space representation
Actual calculations are easier in Fourier transform. In order not to worry about analysis
let us first work in finite volume Λ = ZdL . Recalling from Section 7.2. we have for p ∈ BL
(defined in (7.3)) X
fˆ(p) = e−ipx f (x).
x∈ΛL
R
where dp is a convenient shorthand for the Riemann sum, converging as L → ∞ to
L
dp where dp is the normalized Lebesgue measure i dp
R Q i
[−π,π]d 2π
. We have, see (7.4, 7.5),
Z
dpeip(x−y) = δxy for x, y ∈ ΛL (10.5)
L
and
X
ei(p−q)x = Ld δpq ≡ (2π)d δL (p − q) p, q ∈ BL . (10.6)
x∈ΛL
where we defined the discrete delta function δL (p − q). Products work out as
Z
ˆ
(f g)ˆ(p) = f ∗ ĝ := dq fˆ(p − q)ĝ(q)
L
and
(f ∗ g)ˆ = fˆĝ.
Example.
X
= C(x − z1 )C(z1 − z2 )3 C(z2 − y) = (C ∗ C 3 ∗ C)(x − y)
z1 z2
with Z Z
G(p)
b = dq drC(q)
b C(r)
b C(p b 2
b − q − r) C(p)
L L
or pictorially
73
q
r
p p
p-q-r
A more useful way to derive the same result is to insert for each C its Fourier represen-
tation:
XZ 5
Y
i[p1 (x−z1 )+p2 (z1 −z2 )+p3 (z1 −z2 )+p4 (z1 −z2 )+p5 (z2 −y)]
= e C(p
b i )dpi
z1 z2 L i=1
Z 5
Y
d d ip1 x−ip5 y
= (2π) δL (−p1 + p2 + p3 + p4 )(2π) δL (−p2 − p3 − p4 + p5 )e C(p
b i )dpi
L i=1
The first delta-function is due to the sum over z1 and the second to the sum over z2 . Note
“momentum conservation” at vertices.
p2 p2
p1 4
X 4
X
p3 p3 p5 p1 = pi , p 5 = pi
i=2 i=2
p4 p4
74
for all y so
X P
b 1 , · · · , pn ) =
G(p G(x1 , x2 , · · · , xn )e−i i p i xi
x1 ,...,xn
X P
= G(0, x2 − x1 , · · · , xn − x1 )e−i i pi xi
x1 ,...,xn
X Pn P
= G(0, x2 , · · · , xn )e−i i=2 pi xi −ix1
e i pi
x1 ,...,xn
X
= ĝ(p2 , . . . , pn )(2π)d δL ( pi )
2) Satisfy momentum conservation at each vertex (coming from the delta functions).
p1 -p-p-p
1 2 3
q1
C(p 1 − p2 − p3 )C(p2 )C(p3 )
b 1 )C(−p
b b b
Z Y3
q2 · dqi C(q
b 1 )C(q
b 2 )C(q
b 3)
i=1
p-q-q q +q -p-p-p
1 2
1 1 2 q3 1 2 3
·C(p
b 1 − q1 − q2 )C(q
b 1 + q2 − p 1 − p2 − p3 )
·C(p
b 1 + p2 − q1 − q2 − q3 ) ≡ Γ(p1 , p2 , p3 )
p+p-q-q-q
1 2 1 2 3
p3 and in x space the graph is
p2
Z 3
i(p1 (x1 −x4 )+p2 (x2 −x4 )+p3 (x3 −x4 ))
Y dpi
G(x1 , · · · , x4 ) = e Γ(p1 , p2 , p3 )
i=1
(2π)d
Remark 2 The expansion is not convergent. Consider e.g. a lattice consisting of one
point i.e. the integral Z ∞
4 r 2
F (λ) = e−λϕ − 2 ϕ dϕ
−∞
75
Then
dn F n
Z
4n − r2 ϕ2 −2n n
√ (4n)!
αn = n
= (−) ϕ e dϕ = r (−1) 2π 2n ∼ C n (2n)!
dλ λ=0 2 (2n)!
P 1
as n → ∞. Thus the Taylor series α λn has αn!n ∼ C n n! i.e. is very badly divergent.
n! n
For λ > 0, there exists rc (λ) such that for r > rc (λ)
For r < rc (λ) there are two phases, where hφ(x)i = ± m 6= 0 and
Let us see how these facts are reflected in the behavior of the perturbation theory when
r = 0. Let us consider the momentum space expressions for the graphs at r = 0. Since
all integrals are over a bounded region [−π, π]d the only problem could come from p = 0
Xd
where µ(p) = 2 (1 − cos pi ) = p2 + O(p4 ) vanishes. Consider the following graph
i=1
76
p = p 1 + p2
Now
q
dd q 1
Z
1
= d
≡ I(p)
[−π,π]d (2π) µ(q) µ(p − q)
p-q
This is integrable in d > 2 if p 6= 0 and behaves as
d−4
|p|
d<4
I(p) ∼p→0 log |p| d=4
constant d>4
n n(d−4)
Thus | {z } = I(p) ∼ |p| which is not integrable if
n
n(4 − d) ≥ d (i.e. for d = 3 if n ≥ 3). Hence e.g. the graph
p p–q
has amplitude
I(q)n dd q
Z
µ(p − q) (2π)d
which is ill-defined for d < 4 and n large.
Gm,n = Gm,n (x1 · · · xm , r) have no limit as r → 0 if d < 4 (at least some of them). For
d ≥ 4 it may be shown that this limit exists. Hence we expect to need radically new ideas
for d < 4. The p ∼ 0 divergencies are called IR (infrared) divergencies.
77
11.2 Ultraviolet
Consider now field theory. Let dµG (ϕ) be the Gaussian measure on S 0 (Rd ) with covariance
G = (−∆+m2 )−1 . We would like to consider the continuum limit of the Ginzburg-Landau
model i.e. to start with finite volume Λ ⊂ Zd we would consider the measure
ϕ(x)4 dd x
R
e−λ Λ dµG (ϕ).
We immediately run into trouble: ϕ(x)4 , as we saw, is not a well defined random variable.
R R
Indeed, ϕ(x)4 dµG = ∞. Actually, we only know ϕ(f ) ≡ ϕ(x)f (x)dx for f ∈ S(Rd ) is
a measurable function on S 0 and has finite moments ϕ(f )n dµ < ∞, ∀n.
R
R
One way to proceed is to regularize the theory. Recall that ϕ(x)4 dµG = 3G(0)2 and
dd p 1
R
G(0) = Rd G(p)
b
(2π)d
. For us, G(p)
b = p2 +m 2 so the divergence is due to insufficient decay
of G(p)
b as |p| → ∞ i.e. UV (ultraviolet) divergence. Thus, let us replace G(p)b by
G
b (p) = G(p)χ(p)
b ≡ G(p)χ
b (p)
2
where χ ∈ S(Rd ) cuts of large |p|. We demand χ(0) = 1 and χ ≥ 0. E.g. χ(p) = e−p
is a good choise. As → 0, G
b (p) → G(p)
b pointwise. Now ϕ(x) is a well defined random
variable :
dd p
Z Z
2n (2n)! n
ϕ(x) dµG = n G (0) , G (0) = G(p)χ(p)b <∞
2 n! (2π)d
[To be pedantic : a priori we know only that ϕ(f ) is measurable, for f ∈ S, so consider
(x−y)2
e− 2
ϕ(f,x ), f,x (y) = . This is a measurable function ` : S 0 → R : ϕ 7→ ϕ(f,x ).
(2π)d/2
Thus lim→0 ` also is measurable and this is what we call ϕ(x) (formally : ϕ(f,x ) =
R R
ϕ(y)f,x (y)dy−→ →0 ϕ(y)δ(x − y)dy = ϕ(x))].
Hence, consider the measure
1 − R [aϕ(x)2 +λϕ(x)4 ]dd x
e Λ dµG (ϕ) (11.1)
ZΛ,
where we added for later purpose also a quadratic term to the Hamiltonian. This measure
is well defined provided a ∈ R, λ > 0. Indeed, as before by Jensen’s inequality we get
2 a2
a
and since aϕ2 + λϕ4 ≥ − 4λ , ZΛ, ≤ e|Λ| 4λ .
The correlation functions are again C ∞ in λ and a, and we have the expansion
N
X
G2k = λn ap G2k,n,p + RN RN = O(λN +1 , aN +1 )
n,p=0
78
given in terms of graphs. The coefficients, G2k,n,p for m2 > 0, have a |Λ| → Rd limit but
not an → 0 limit, as we saw above.
The question of continuum limit will not arise in statistical mechanics where the lattice
spacing is a fixed nonzero quantity. However, in quantum field theory this question arises.
In this case there is no fundamental length and one would like to have a quantum theory
of continuum fields. Moreover these fields depend also on time and are defined on the four
dimensional space time R4 . Brushing aside for the moment the fact that this space time
carries the Minkowski metric instead of the Euclidean one we have been considering and
that in quantum mechanics the correlation functions of fields are replaced by expectation
values of field operators in Hilbert space, it turns out that in case of a scalar field theory
one needs to address precisely the question of → 0 limit as above.
The problem of divergences of the coefficients of perturbation theory arose in the
1930’s when physicists were studying Quantum Electrodynamics (QED), the quantum
theory of electromagnetic field interacting with electrically charged matter. In this theory
there are two sorts of fields, the electromagnetic field A(x) which resembles our φ(x) but
instead takes values in A(x) ∈ R4 . A(x) is the four dimensional electromagnetic vector
potential. It is a Gaussian field with zero mass, thus resembling our r = 0 free field.
The second field describes electrons and has non zero mass. The interaction term in the
Hamiltonian makes the theory non Gaussian; the analogue of our parameter λ is played
by the electron charge e. One can then proceed to derive a perturbation series as above
in powers of e and as above the individual terms are ill defined, diverging due to small
scale (large momentum) behavior of the integrands.
One would like to think about the electron charge (and its mass) as given physical
constants that enter the Hamiltonian describing the dynamics. The divergence of the
perturbation theory indicates that this point of view is incorrect. Rather one should
think about these parameters depending on scale. Thus physicists introduced in the 50’s
the idea of renormalization. We should think about the parameters entering -cutoff
theory describing the charge and mass of the electron in that scale. These can very well
be different from the ones we measure that are interpreted as the charge and mass of
the electron in the measurement scale. Once we have introduced the renormalization
group we will make this picture more precise. For the time being we just remark that the
parameters a and λ entering the measure (11.1) should be allowed to depend on the scale
. Then the → 0 limit question can be posed as follows.
Question. Find three functions Z(), a(), λ() such that
2m
Y
Z()m h ϕ(xi )i
i=1
79
2m
Y
0 2md m
converge, as → 0, to elements of S (R ) (or Z() h ϕ(fi )i , converge for all fi ∈
i=1
S(Rd )). Here h i is the expectation w.r.t. the limit of the measure (11.1), as Λ → Rd ,
a = a() and λ = λ().
Remarks. 1. We allowed (it turns out this often is necessary) for a third parameter that
1
renormalizes the value of the field: ϕ(xi ) is replaced by Z() 2 ϕ(xi ).
Obviously λ() = 0, a() = 0, Z = 1 is a solution. We ask if there are any limits that are
non-Gaussian. The answer is the following:
For d = 2, Z = 1, a = −6λG (0), λ() = λ works.
For d = 3, Z = 1, a = −6λG (0) + αλ2 log (α explicit), λ() = λ works.
For d ≥ 4 there is no non-Gaussian limit.
Our objective is to explain these claims.
1
Remark. χ is a regulator. It cuts off momentum |p| > ∼ very effectively. We could also
use a lattice-cutoff. Thus consider the following Hamiltonian on fields defined on (Z)d :
Let ϕ : (Z)d → R and set
1
X (ϕ(x) − ϕ(y))2 r X d X
H= 2
d + ϕ(x) 2
+ d (a()ϕ(x)2 + λ()ϕ(x)4 ). (11.2)
2 2 x x
|x−y|=
we have
G (x1 , . . . , x2m ) = Z()m (2−d)m G() (x1 /, . . . , x2m /) (11.4)
where
2m
Y
G() (y1 , . . . , y2m ) = h φ(yi )i()
i=1
()
and h−i is the expectation w.r.t. the measure
" #
1 X
lim exp − 2 a()φ(x)2 + 4−d λ()φ(x)4 dµC (φ)
Λ↑Z d ZΛ,
x∈Λ
80
where dµC is the Gaussian measure on unit lattice fields φ ∈ s0 (Zd ) with covariance
C = (−∆ + 2 r)−1 .
Thus our UV problem → 0 is the same thing as staying on a fixed lattice, taking
distances to ∞ (xi /), and going to critical point (2 r → 0). It should now be obvious
that both IR and UV problem have something to do with Statistical Mechanics at the
critical point.
Remark. The -lattice theory can be interpreted as a particular cutoff of the continuum
theory. Namely, the measure with H as in (11.2) with a = λ = 0 is the Gaussian measure
on s0 ((Z)d ) with covariance
X C = (−∆ + r)−1 where the latter is an operator on `2 (Zd )
i.e. (−∆ ϕ)(x) = [ϕ(x) − ϕ(x + i)]−2 for ϕ : (Z)d → R. Concretely
|i|=1
eip(x−y) dd p
Z
1
(x, y) =
−∆ + r [− π , π ]d −2 µ(p) + r (2π)d
which shows that p is cut-off to |pi | ≤ π/. Similarily, instead of the lattice model we may
consider the Statistical Mechanics where ϕ is on Rd but has the cutoff χ, i.e. the measure
1 −λ R ϕ(x)4 dx
lim e Λ dµG1 (ϕ)
Λ→Zd ZΛ
1
with G1 (p) = p2 +r
χ(p).
(as |x − y| > 0) where > 0. We are after a theory that explains the leading term in
this asymptotics and the fact that this term (or the exponent a) is universal i.e. stays
the same when the Hamiltonian is changed (at least under some changes). Thus suppose
there is a class of Hamiltonians that satisfy (12.5). Then the details of these Hamiltonians
affect the sub leading asymptotics i.e. the RHS of (12.5). We can get a special theory
where the scale invariance satisfied by the leading term is exact by taking the scaling limit.
Recall the relations (11.3) and (11.4) between cutoff quantum field theory and statistical
mechanics. Let, for L > 0
a
ϕL (x) = L 2 φ(Lx). (12.6)
81
Then if (12.5) is satisfied we get
A
lim hϕL (x)ϕL (y)i = .
L→∞ |x − y|a
On the other hand ϕL can be seen as a cutoff 1/L field as it depends on x ∈ (L−1 Z)d .
Hence if the LHS is given by an expectation hϕ(x)ϕ(y)i the field ϕ lives on Rd . Thus we
expect the limits, if they exist,
Y
lim h ϕL (xi )i
L→∞
i
to be the correlation functions of a field theory. This field theory is called the scaling
limit of our model. By definition it is scale invariant, e.g.
La hϕ(Lx)ϕ(Ly)i = hϕ(x)ϕ(y)i.
Universality then would mean that several Hamiltonians give rise to the same scaling
limit.
The drawback of this formulation is that, first, the scaling limit is a different object
than the one we started with. The latter one is a fixed (unit) lattice model whereas the
former one is a continuum object. More importantly, we have not given any constructive
approach to the study of the scaling limit: we still need to understand the statistical
mechanics at the critical point, in particular we need to show (12.5). The Renormalization
group addresses these two problems. First, it supplements scaling by another operation,
coarse graining, that allows one to stay in the category of fixed lattice spacing theories.
Second, it provides actually a tool to study the critical theory.
We replace the scale transformation (12.6) by
a
φL (x) = L 2 φaverage (Lx) (12.7)
d d
(12.7) and (12.8) define a map φ ∈ RZ → φL ∈ RZ , i.e. the scaled field is also defined
on Zd . We take for convenience L > 1 odd integer.
• • •
Here L = 3
• • Lx •
• • • Lx + y
82
We see that φ → φL involves :
a) “Coarse graining” : average over details of φ for scales ≤ L.
b) Scaling : φL (x) depends on φ near Lx and we multiply φ by La/2
φ → φL is called the Block-spin transformation. (12.7) and (12.8) define a linear map
d d
CL : RZ → RZ , concretely,
(
a 1 |(Lx − y)i |< L/2
(CL )xy = L 2 −d
0 otherwise
Note that
CL2 = CL2 .
Thus, the iteration of (12.8) n times is the same as doing it once with Ln :
Now observe :
Proposition 12.1 Suppose that (12.5) holds. Then, for maxi |xi − yj | > 1,
∗
h(CLn φ)(x)(CLn φ)(y)i−→
n→∞AG (x, y)
with Z Z
∗
G (x, y) = du dv|x − y + u − v|−a
and the A-term → A du dv|x − y + u − v|−a . We used max |xi − yi | > 1, |(u − v)i | <
R R
83
1
Example For the Gaussian Ginzburg-Landau model with covariance −∆ we take a = d−2
and get n n
n n n(d−2) −2nd
XZ dd p eip(L x−L y+u−v)
hCL φ(x)CL φ(y)i = L L d
d (2π) µ(p)
u,v [−π,π]
−n
dd p eip(x−y+L (u−v))
X Z
−2nd −2n
=L L
u,v [−Ln π,Ln π]d (2π)
d µ(L−n p)
Z Z
1
−→
n→∞ du dv (x − y + u − v),
−∆Rd
using L2n µ(L−n p) → p2 , as n → ∞, and where −∆Rd is the usual −∆ on L2 (Rd ). Now,
1 const
(x − y) = .
−∆Rd |x − y|d−2
1
Remark. In this example, CLn φ(x) are Gaussian, with covariance CLn −∆ (CLn )T . T denotes
the transpose, this is just what we write above, i.e.
n 1 1
X
n T n n
CL (C ) (x, y) = (CL )xz (CL )yw .
−∆ L z,w
−∆ zw
= W (CLT f )
d d
µL is the probability distribution of the block spins CL φ. Thus CL : RZ → RZ induces
a map
CL∗ : B → B
where B are the cylinder measures on s0 (Zd ).
CL∗ is called the Renormalization Group Transformation on measures.
84
12.3 Transformations on Hamiltonians
Concretely, let us work in finite volume ΛN and take ΛN = LN -box centered at the origin.
Thus φ ∈ RΛN , CL φ ∈ RΛN −1 and CLn φ ∈ RΛN −n . Let µ be a probability measure
1 −H(φ) ΛN
dµ(φ) = e d φ
Z
when H is some function H : RΛN → R such that e−H is integrable. Then,
with Z
1 Y
F (Ψ) = e−H(φ) δ(Ψ(x) − (CL φ)(x))dΛN φ
Z x∈ΛN −1
R
(Z normalizes F dΛN −1 Ψ = 1).
a
Let us write this more explicitely : Ψ(x) equals L 2 times the average of φ in the L-cube
centered at Lx. Thus, if we fix Ψ(x), ∀x ∈ ΛN −1 , we need to integrate out all fluctuations
around this average. In other words, let φ0 be a configuration that is constant in each
a
L-cube, and equals the average of φ there. I.e. φ0 (Lx + y) = L− 2 Ψ(x) for all x ∈ ΛN −1 ,
all y with |yi | < L/2. Using our previous notation this equals
a a
φ0 = L− 2 L− 2 +d CLT Ψ = Ld−a CLT Ψ
a
(recall, (CLT )xy = L 2 −d if x ∈ L-cube at Ly, and equals 0 otherwise). Thus φ = φ0 + Z
with Z having 0 average over L-cubes i.e. CL Z = 0. Thus
Z !
1 d−a T
Y
F (Ψ) = e−H(L CL Ψ+Z) δ (CL Z)(x) dΛN Z
Z x∈Λ N −1
Even more concretely, take z ∈ RΛN \LΛN −1 i.e. z(x) ∈ R, ∀x ∈ ΛN where x 6= Lw, w ∈ Zd .
Put
z(x) x ∈ ΛN \LΛN −1
X
Z(x) = − z(Lw + y) x = Lw
|yi |<L/2
Then CL Z = 0 and so
Z
1 d−a C T Ψ+Z)
F (Ψ) = e−H(L dΛN \LΛN −1 z
Z
a concrete integral over “fluctuation variables” z(x).
0
Let F = e−H . The map RL : H → H0 is called the RG Transformation on Hamiltonians,
i.e.
Z
d−a T
RL H = − log e−H(L CL Ψ+Z) dz (12.9)
85
Here H : RΛN → R and RL H : RΛN −1 → R. Note how the volume contracted from an
LN -box to an LN −1 -box.
Remarks. 1. It was trivial to define the RG on measures and in infinite volume : CL∗ µ
is automatically a measure. On Hamiltonians, we need to
a) work in finite volume,
b) make some assumptions on H so that RL H in (12.9) is well defined.
2. RG is called a “group” because CLn = CL n i.e. (CL∗ )n = CL∗ n and (RL )n = RLn . Actually,
it is a semigroup : R1 = id but RL−1 is not defined.
Our Proposition 12.1 says that
Z
−→ G∗ (x, y)
(CLn φ)(x)(CLn φ)(y)dµ n→∞
Thus, we might hope that the measures µLn converge to some measure µ∗ such that
Z
Ψ(x)Ψ(y)dµ∗ (Ψ) = G∗ (x, y)
−→ H∗ ,
(RL )n H n→∞ RL H∗ = H∗
86
we have a bounded spin model and can ask whether RL actually maps Hamiltonians to
Hamiltonians. I.e. suppose µ is a Φ-Gibbs measure, Φ ∈ B. Is CL∗ µ a Gibbs measure for
some Φ0 ? If so, define RL Φ = Φ0 . This is expected to be true in Ising model, but proven
only for small β or large β (where one needs to enlarge B, surprisingly !).
3. For unbounded spins, RL and its iterations have been rigorously studied for the
Ginzburg-Landau model. It turns out that one can, to a certain extent set up a space
of Hamiltonians where RL acts; however, one has to supplement this with a different
representation of µ for large values of φ’s.
For the rest of the time, we will not discuss these problems. Rather we
a) Pretend that RL is defined in some space of H’s and see what implications this could
have (Section 13).
b) Carry out a perturbative analysis of RLn H for H = Ginzburg-Landau model and un-
derstand the IR and UV problem this way (Sections 14-16).
Λ % Zd and HΛ is some finite volume version of H (or more properly, H = {HΛ }Λ⊂Zd ).
2. The RG is defined in K : RL : K → K i.e.
CL∗ µH = µRL H
We also suppose that there is a metric in K and if Hn → H, then µHn → µH .
Let us see what kind of picture of critical phenomena emerges from such assumptions.
Let H∗ be a fixed point of RL : RL H∗ = H∗ .
Definition 13.1 The stable manifold of H∗ in K is
Ms = {H ∈ K | RLn H → H∗ }
What can we say about the decay of correlations of H ∈ Ms ? Let us define
Definition 13.2 H ∈ Ms is critical if
sup |hφ(x)φ(y)iH |e|x−y|/ξ = ∞
x,y
87
for all ξ < ∞.
Hence for critical H there are no A > 0, ξ < ∞ such that |hφ(x)φ(y)iH | ≤ Ae−|x−y|/ξ for
all x, y. X
a
Recall a in the definition of CL : (CL φ)(x) = L 2 L−d φ(Lx + y). We have
|yi |< L
2
88
Next, consider
Mξ = {H|H has correlation length ξ}
For this we assume that all H ∈ K have a well defined ξ, i.e. the limit
1
lim − loghφ(0)φ(x)i = ξ −1 exists
|x|→∞ |x|
Then a) Ms = M∞
b) RL : Mξ → Mξ/L
R2 H0
RH0
Mξ1 /L2
H0 Mξ1 /L
Mξ 1
H∗ R2 H
RH H
Ms
RL takes :
– critical H to H∗ upon iteration
– noncritical H away from H∗ upon iteration.
89
The spectrum of L plays now a crucial rule in the analysis of R near H∗ .
In the known examples, the spectrum of L has the following structure :
i.e. the Hi perturbation increases (λi > 1) or decreases (λi < 1) exponentially in n or
stays constant (λi = 1). Let us interpret this. Suppose first the case where there is only
one λ > 1, with multiplicity 1, say λ1 . Then for |x − y| >> 1
Take n such that λn1 = O(1). Recall that we had one direction in K where we depart
from H∗ upon iteration, namely the one parametrized by ξ. Thus, it is natural to assume
that H∗ + H1 has ξ < ∞ and H∗ + O(1)H1 has ξ = O(1). From above, then
= Ln O(1).
ξ ∼ (T − Tc )−1/α
and −1/α is the so called correlation length critical exponent describing how ξ diverges
as T → Tc 7 .
|x−y|
7 1 −
Note that we got hφ(x)φ(y)iT ∼ |x−y|a e
ξ(T ) for all x, y.
90
Universality. Consider now H = H∗ + Hi , λi < 1. As before,
so, if λi = L−αi , αi > 0, then, since |u−v| ∼ Ln ⇒ L−αi n ∼ |u−v|αi . Thus these directions
give subleading corrections to the decay determined by H∗ . We have universality : the
N
Y YN
leading assymptotics of h φ(L xi )iH is h φ(Ln xi )iH∗ if H ∈ Ms (as n → ∞ and xi
n
i=1 i=1
are apart from each other) and the details of H are only seen in the O(L−αi n ) corrections.
We say that all H ∈ Ms have the same critical behaviour.
Also, all the exponents αi are determined by H∗ , which appear not to depend on which
H we have (in Ms ).
91
This is the Gaussian fixed point of the block spin RG. We should now like to perform
the linear analysis near C ∗ . We will do this in a slightly different model, where it is less
messy (everything can be done in the case of block spin with identical conclusions).
Lemma 14.1 Suppose µGi , i = 1, 2 are two Gaussian measures on S 0 (Rd ) and G = G1 +G2 .
Then for all F ∈ L∞ (µG ),
Z Z
F (ϕ)dµG (ϕ) = F (ϕ1 + ϕ2 )dµG1 (ϕ1 )dµG2 (ϕ2 ).
Z Z Z
1 1
e i(ϕ1 (f )+ϕ2 (f ))
dµG1 dµG2 = e iϕ1 (f )
dµG1 eiϕ2 (f ) dµG2 = e− 2 [(f,G1 f )+(f,G2 f )] = e− 2 (f,Gf ) .
d−2
Let us apply the Lemma to (14.1). We note that the covariance of L− 2 ϕ( Lx ) is L2−d G( x−y
L
)
if ϕ has covariance G. Hence :
·
Z Z
d−2
F (ϕ)dµG (ϕ) = F (L− 2 ϕ( ) + Z)dµG (ϕ)dµΓ (Z).
L
93
This gives our RG :
with
Z
1 −Ve (ϕ)
dν̃(ϕ) = e dµG (ϕ) Z = e−V dµG
e
Z
Z
− d−2 ·
e−V (ϕ) = e−VΛ (L 2 ϕ( L )+Z) dµΓ (Z)
e
− d−2
d−2
ϕ( L· )+Z)
F (L− ϕ( L· ) + Z)e−VΛ (L
R 2
2 dµΓ (Z)
Fe(ϕ) = − d−2
.
ϕ( L· )+Z)
e−VΛ (L
R 2
dµΓ (Z)
Hence, we got a renormalized measure ν̃ which is again of same form as ν and we can
define the RG as a map of V ’s by
Z
− d−2 ·
RL V (ϕ) = V (ϕ) = − log e−V (L 2 ϕ( L )+Z) dµΓ (Z).
e (14.4)
Compare with block-spin ! The latter can actually, with a suitable definition of ϕ on the
RHS, be written exactly as (14.4).
R
Remark. If V = Λ P (ϕ(x))dx then
·
Z
− d−2 d−2 x
V (L 2 ϕ( ) + Z) = P (L− 2 ϕ( ) + Z(x))dx
L Λ L
Z
d−2
= Ld P (L− 2 ϕ(y) + Z(Ly))dy
L−1 Λ
i.e. it depends on ϕ in L−1 Λ. So we may call Ve by VeL−1 Λ if we wish. Of course this will
not matter in the limit Λ % Zd .
We have now a fixed point V = 0, i.e. dµG → dµG under our RG. Call this the Gaussian
fixed point.
94
This starts to look not too bad ! We will study the spectrum of this linear map. Let us
first see some examples.
R
Example 1. Let V = Λ ϕ(x)2 dx. Then
Z Z x 2
− d−2
LV (ϕ) = L 2 ϕ + Z(x) dxdµΓ (Z) =
L
Z Λ 2 Z Z
x
= L2−d ϕ dx + dx Z(x)2 dµΓ
L
Z Λ Λ
Z
2 2
=L ϕ(x) dx + |Λ|Γ(0) since Z(x)2 dµΓ = Γ(0)
Λ/L
Z
= L2 ϕ(x)2 + Ld−2 Γ(0) dx.
Λ/L
Hence, since, from (14.1), G(0) = L2−d G(0) + Γ(0), we get Ld−2 Γ(0) = Ld−2 G(0) − G(0),
and Z Z
2 2
L (ϕ(x) − G(0))dx = L (ϕ(x)2 − G(0))dx
Λ Λ/L
R
i.e. we found an eigenfunction (as Λ → Rd ) (ϕ(x)2 − G(0))dx with eigenvalue L2 i.e.
with exponent α = 2. Note that this corresponds to a mass term. In the lattice cutoff
case we have
1 − m2 Px∈Λ ϕ(x)2
lim e 2 dµG (ϕ) = dµGm2 (ϕ)
Λ→Rd ZΛ
Z
2 1
Gm2 (x − y) = eip(x−y) e−p .
µ(p) + m2
(Can you figure out what happens in the momentum space cutoff case?).This is a relevant
eigenvector.
R 2
x
Example 2. Consider V (ϕ) = Λ ∇ϕ(x) dx. In the same way, using ∇ ϕ( L ) =
R 2
−1 x
L (∇ϕ)( L ) we get : LV = Λ/L ∇ϕ(x) dx − |Λ|∇2 Γ(0) and
Z h 2 i Z 2
2 2
L ∇ϕ(x) + ∇ G(0) dx = ∇ϕ(x) + ∇ G(0) dx
Λ Λ/L
so this is a marginal eigenvector : note that it corresponds to a change of the fixed point
R
which is corresponding to H = (∇ϕ)2 : in fact we have a one-parameter family of fixed
points dµαG α > 0.
R R
Example 3. Consider V (ϕ) = Λ ϕ(x)4 dx. Now (using dµΓ Z 2m+1 = 0)
Z x 4 x 2
2(2−d) 2−d 2 4
dµΓ (Z) L ϕ + 6L ϕ Z(x) + Z(x)
L L
x 4 x 2
= L2(2−d) ϕ + 6L2−d ϕ Γ(0) + 3Γ(0)2
L L
95
so Z Z
4−d 4 2
LV = L ϕ(x) dx + 6L Γ(0) ϕ(x)2 dx + 3Γ(0)2 |Λ|
Λ/L Λ/L
so, putting Z
V4 = (ϕ(x)4 − 6G(0)ϕ(x)2 + 3G(0)2 )dx
Λ
4−d
we get LV4 = L V4 , i.e. V4 is relevant if d < 4, marginal if d = 4 and irrelevant if d > 4.
It is obvious now that we have eigenvectors Vn , polynomial of degree n, with eigenvalues
2−d
Ld Ln 2 . A simple way to get the formula for these eigenvectors is to consider
2−d R
Z
ϕ(f ) L 2 x
ϕ( L )f (x)dx ˜ 1
Le =e dµΓ eZ(f ) = eϕ(f )+ 2 (f,Γf )
2−d 2+d
where f˜(x) = L 2 Ld f (Lx) = L 2 f (Lx). Since
Z Z x − y
(f˜, Gf˜) = L2+d
f (Lx)G(x − y)f (Ly)dxdy = L2−d
f (x)G f (y)dxdy
L
we have, see (14.1), (f, Γf ) = (f, Gf ) − (f˜, Gf˜) and so
1 1 ˜ ˜ ˜
Le− 2 (f,Gf )+ϕ(f ) = e− 2 (f ,Gf )+ϕ(f ) .
96
with eigenvalue
2−d P P
Ld L 2
(n+ nk )
L−( knk )
i.e. the more there are ϕ’s or derivatives, the more V is irrelevant. In the class of even
local V ’s the relevant ones are, for d ≥ 3 :
R 2
ϕ L2 always
R 4
ϕ L4−d d < 4
and specify the K’s. For this, generalize a bit what we found above. Define the normal
ordered product :
n
Y X |I| Y Y
: ϕ(xi ) := (−1) 2 ϕ(xj )h ϕ(xi )i.
i=1 I⊂{1,··· ,n} j6∈I i∈I
Homework.
Z n n n
Y Y Y ∂ h 1 i
: ϕ(xi ) : f (xi )dxi = exp − (f, Gf ) + ϕ(f )
∂λi λi =0 2
i=1 i=1 i=1
n
X
f= λi fi , fi ∈ S(Rd ).
i=1
This implies :
n n x
2−d
Y Y i
n
L: ϕ(xi ) := L 2 : ϕ :.
i=1 i=1
L
97
The space Kγ consists of finite sums of the form. (Let ϕ ∈ S(Rd ))
XZ N
Y N
Y
ni mi
V (ϕ) = K∼
n m (x · · · xN ) :
∼ 1
ϕ(xi ) ∇ϕ(xi ) : dxi (14.5)
n m
∼ ∼ i=1 i=1
d
Y
where n∼ = (n1 , · · · , nN ), m d
∼ = (m1 , · · · , mN ), ni ∈ N, mi ∈ N , (∇ϕ)
mi
= (∇α ϕ)miα ,
α=1
d
mi = (mi1 , · · · , mid ) and K∼n∼m(x1 · · · xN ) = K∼ n∼m(x1 + x, x2 + x, · · · , xN + x) ∀x ∈ R ,
(translation invariance) and
Z
γL(0,x2 ,··· ,xN )
|K∼ m(0, x2 , · · · , xN )|e
n∼ ≡ kK∼
n∼mkγ < ∞
where L(x1 , · · · , xN ) ≡ d(x∼) is the length of the shortest connected path Γ in Rd such
that all xi belong to Γ; here γ > 0.
Then LV has kernels K 0 :
2−d PN P
Kn0 m(x∼) = L 2 i=1 (ni +mi ) L− mi
LN d K∼m(Lx
n∼ ∼)
∼∼
and
Z
2−d PN
− N 0
P
kKn0 mkγ 0 = L 2 i=1 (ni +mi ) L i=1 mi L Nd
|K∼m(Lx
n∼ ∼)|e
γ d(x
∼) dx2 · · · dxn
∼∼
Z
2−d PN
− N 0 −1 x )
P
= L 2 i=1 (ni +mi ) L i=1 mi L d
|K∼
n∼m(x
∼
)|eγ d(L ∼ dx2 · · · dxn .
98
R
Lemma 14.2 Let K(x) be rotation invariant with kKkγ = |K(x)|eγ|x| dx < ∞. Then
for all ϕ ∈ S(Rd ))
Z Z
K(x − y)ϕ(x)ϕ(y)dxdy = α ϕ(x)2 + β(∇ϕ(x))2 + Ve (ϕ)
where |α|, |β| ≤ CkKkγ and Ve ∈ Kγ/2 is irrelevant with kernel ≤ CkKkγ .
R
Proof. Since |K(x)|eγ|x| dx < ∞ we get
Z Z
|K(p)| = | e K(x)dx| ≤ e|Imp||x| |K(x)|dx
b ipx
K(p)
b = K(0)
b + βp2 + R(p)
b (14.7)
(we used the rotation symmetry to get the second term of the form βp2 ) where
1 1 d3 b
Z
R(p) =
b dt(1 − t)2 3 K(tp)
2 dt
X0
= pα pβ pγ H
b αβγ (p)
αβγ
we have
|α| = |K(0)|
b ≤ kKkγ
and using Cauchy’s theorem
|β| = |∂ 2 K(0)|
b ≤ γ −2 kKkγ
and so
Z 1
H
b αβγ (p) = dt(1 − t)2 (∂α ∂β ∂γ K)(tp)
b
0
Z 1
3
= i dt(1 − t)2 (xα x\
β xγ K(x))(tp).
0
99
Hence
Z 1 Z
−d
kHαβγ kγ/2 ≤ dtt dx(|x|/t)3 |K(x/t)|eγ|x|/2
Z0 1 Z
= dt dx|x|3 |K(x)|eγt|x|/2 .
0
48 γ|x|/2
Since |x|3 ≤ γ3
e we get finally
48
kHαβγ kγ/2 ≤ kKkγ .
γ3
Thus Z
3
Ṽ = i Hαβγ (x − y)∂α φ(x)∂α ∂β φ(y)dxdy
R
where a = K(0, x2 , x3 , x4 )dx2 dx3 dx4 and Ve is irrelevant. Thus every V ∈ Kγ can be
written as
Z Z Z
2 2
V =r : ϕ : +z : (∇ϕ) : +λ : ϕ4 : +Ve , Ve ∈ Kγ/2
and then Z Z Z
LV = L r 2 2
: ϕ : +z 2
: (∇ϕ) : +L 4−d
λ : ϕ4 : +Ve 0
with Ve 0 ∈ KLγ/2 , kVe 0 kLγ/2 ≤ L−2 kVe k. These are our “coordinates” in K.
15 Perturbative analysis of RV
15.1 General formalism
Consider now the full R:
·
Z h i
2−d
RV (ϕ) = − log dµΓ (Z) exp −V (L ϕ( ) + Z) .
2 (15.1)
L
Suppose V is in K as above. There are several problems :
100
2. More serious problem is the convergence of (15.1) as Z becomes large. This is the
stability problem. Say for V = λϕ4 we need at least Re λ > 0. For a V as in the
previous section, this positivity property is not easy to state. Instead of trying to
address the stability problem, we will calculate RV perturbatively in λ if V = λϕ4 .
With a lot of work, RV can actually be analyzed rigorously. One does the following :
The problem is to combine a) and b) : ϕ(x) can be small and ϕ(y) large nearby.
One uses expansion ideas from statistical mechanics to decouple these problems.
R
Thus, let us start with V (ϕ) = λ Λ ϕ(x)4 dx and see what RV looks like. Expand in
powers of λ:
X N
(RV )(ϕ) = λn (RV )n (ϕ) + Remainder
n=0
1 dn 1 dn Z 2−d
−V (L 2 ϕ( L· )+Z)
(RV )n (ϕ) = (RV )(ϕ) = − log e dµΓ (Z) .
n! dλn λ=0 n! dλn λ=0
Hence,
Z Z x 4
2−d
λ(RV )1 (ϕ) = λ L ϕ 2 + Z(x) dxdµΓ = LV (ϕ)
Λ L
Z Z Z
2 1h 2
i
λ (RV )2 (ϕ) = − V dµΓ − V dµΓ V dµΓ
2
1 2−d · 2−d ·
= − hV (L 2 ϕ( ) + Z); V (L 2 ϕ( ) + Z)iΓ
2 L L
where we use the notation of truncated expectations.
In general
1
λn (RV)n (ϕ) = (−)n−1 hV ; V ; · · · ; V iΓ .
n!
2−d
Consider (RV)2 : Denote L 2 ϕ( Lx ) ≡ Ψ(x). So
D 4 4 E D 4 E D 4 E
Ψ(x) + Z(x) Ψ(y) + Z(y) − Ψ(x) + Z(x) Ψ(y) + Z(y)
µ µ µ
2 2 2 3 3
= 36 · 2 · Ψ(x) Ψ(y) Γ(x − y) + 16 · 3 Ψ(x)Ψ(y) + Ψ(x) Ψ(y) Γ(x − y)Γ(0)
!2
4
+4 ! Γ(x − y)4 + · 2Γ(x − y)2 Γ(0)2
2
+16 · 3 ! Ψ(x)Ψ(y)Γ(x − y)3 + 6 · 4 · 3(Ψ(x)2 + Ψ(y)2 )Γ(x − y)2 Γ(0)
+16Ψ(x)3 Ψ(y)3 Γ(x − y) + 16 · 3 · 3Ψ(x)Ψ(y)Γ(x − y)Γ(0)2
or graphically
101
+ + + +
+ + +
where the sum runs through all connected graphs G with n + m vertices x1 ,. . . , xn , y1 ,. . . ,
ym . Vertices yi have four edges ({z, z} allowed) attached and vertices xj have 4 − nj edges
attached where nj > 0. The WG (ϕ) is equals
Z Z Y m
Y Yn Z n
Y
0 n
n(G) Γ(z − z ) dyj Ψ(xi ) dxi ≡ KG (x1 , . . . , xn )
i
Ψ(xi )ni dxi
{z,z 0 }∈e(G) i=1 i=1 i=1
Recalling that |Γ(x − y)| ≤ Ce−α|x−y| , for some α and the fact that the graph is connected
we get
Exercise. kKG kγ < ∞, for some γ > 0.
N
X
Thus λn (RV )n (ϕ) ∈ Kγ for some γ. It is more convenient to consider V (ϕ) =
R n=0
λ Λ
: ϕ(x)4 : dx where, recall, : ϕ4 (x) := ϕ(x)4 − 6ϕ(x)2 G(0) + 3G(0)2 . Then
Z
4−d
λ (RV)1 (ϕ) = LV = λL : ϕ(x)4 :
Λ/L
102
where Ve is irrelevant i.e. LVe contracts and Ve = O(λ2 ). Here a comes from
15.2 Iteration
R R R
Suppose now V = λ : ϕ4 : +r : ϕ2 : +z : (∇ϕ)2 : +Ve where Ve ∈ K is irrelevant,
kLVe kγ ≤ L−α kVe kγ for α > 0. Calculate perturvatively RV:
N
X 1
(RV )(ϕ) = (−1)n+1 ( )hV ; V ; · · · ; V iΓ + Remainder
n=0
n!
(15.2)
d−2
where V = V (L− 2 ϕ( L· ) + Z)
N
X
≡ (RV)n (ϕ) + Remainder
n=0
The point is now that RVn ∈ Kγ again. Indeed, we get graphs as before and also from
r, z and Ve . Pictorially, the Ve ones are
103
4
Z Y 4
x Z Y
2−d i
= dxi L 2 ϕ dyi K(x1 , x2 , y1 , y2 )Γ(y1 − y3 )Γ(y2 − y4 )K(y3 , y4 , x3 , x4 )
i=1
L i=1
Z Y
≡ e 1 , · · · , x4 )
K(x ϕ(xi )dxi
where
4
Z Y
e ) = L4−2d L4d
K(x dyi K(Lx1 , Lx2 , y1 , y2 )Γ(y1 − y3 )Γ(y2 − y4 )K(y3 , y4 , Lx3 , Lx4 )
∼
i=1
and
Z
eγL(0,x2 ,x3 ,x4 ) |K(0,
e · · · , x4 )|dx2 · · · dx4
Z
−3d −1
=L e L−1 x2 , L−1 x3 , L−1 x4 )|dx2 · · · dx4
eL γL(0,x2 ,x3 ,x4 ) |K(0, (15.3)
−1 −1
with K̄ = eL γL K, Γ̄(x) = eL γ|x| Γ(x).
Use |Γ̄(y2 − y4 )| ≤ C, integrate over y4 , x3 , x4 :
3
Z Y
4−d
(15.3) ≤ L dyi dx2 |K̄(0, x2 , y1 , y2 )Γ̄(y1 − y3 )|kKkL−1 γ .
i=1
104
where Ven ∈ K and kLVen k ≤ L−α kVn k, α > 0.
We get the recursion (denote gn = (zn , rn , Ven ))
zn+1 = zn + O(Vn2 )
rn+1 = L2 rn + O(Vn2 )
λn+1 = L4−d λn + aλ2n + O(λ3n , λn gn , gn2 )
Ven+1 = LVen + O(Vn2 ).
15.3 d>4
We want to determine the critical point, i.e. take
Z Z Z
V = z : (∇φ) : +r : ϕ : +λ : ϕ4 :
2 2
and find r(λ) such that V ∈ Ms . Since 4 − d < 0 we expect the relevant fixed point to
R
be the Gaussian V = z̃ : (∇ϕ)2 :, for some constant z̃ 6= z, noted z∞ below.
Hence, the problem is to find r0 = r(λ) such that
rn , λn , Ven → 0 as n → ∞.
I1 ⊂ I0
such that
r1 (I1 ) = [−Aλ21 , Aλ21 ].
Now, keep on iterating. We find intervals In ⊂ In−1 ⊂ · · · I0 such that rn as a function of
r0 satisfies
rn (In ) = [−Aλ2n , Aλ2n ]
and λn ≤ L(4−d−)n λ, for > 0, kVen k → 0.
∞
\ ∞
\
Thus, In 6= ∅ (since In closed ⊂ In−1 ) and, for r0 ∈ In , we have
n=0 n=0
Ven , λn , rn → 0 as n → ∞.
105
∞
\
Actually it is readily seen that |In | n→∞
−→ 0 so In = one point ≡ r(λ), the critical r value.
n=0
Clearly r(λ) = O(λ2 ).
n−1
X
How about zn ? Since zn+1 = zn + O(λ2n ) and z0 = 0 we have zn = O(λ2m ) =
m=0
n−1
X
O((L(4−d)m λ)2 ) n→∞
−→ z∞ = O(λ2 ). So
m=0
Z
n
−→ z∞
R V n→∞ : (∇ϕ)2 :
2
R
Hence, we have the one-parameter family of Gaussian fixed points e−z : (∇ϕ) : dµG and
2 4 −z∞ : (∇ϕ)2 :
R R R
e−r(λ) : ϕ : −λ : ϕ : dµG −→
R n e dµG .
The Gaussian line has one unstable direction : ϕ2 : and, once that is fixed to the critical
value, we flow to the line :
r
z
Ms
(r(λ), λ)
15.4 d=4
Since λn+1 = λn to leading order we need to go to O(λ2 ). The sign of a in λn+1 =
λn + aλ2n + · · · is very important so let us calculate it.
We will prove that λn → 0 and rn , zn , Ven = O(λ2n ). Thus a gets contribution from the
term Z
1 D 2−d x 4 2−d y 4 E
− : L ϕ2 + Z(x) : : L ϕ 2 + Z(y) : dxdy.
2 L L
This gives graph a) which contributes directly and b) which
contributes indirectly.
106
a) Equals
Z
72 4−2d x 2 y 2
− L dxdyϕ ϕ Γ(x − y)2 dxdy
2 L L
Z x 4 Z
72 4−2d
= − L dxϕ dyΓ(y)2 + irrelevant (goes into Ve ).
2 L
b) We have to normal order since Ven+1 is normal ordered. This graph equals
Z
1 2−d
x 3 y 3
− 42 L−( 2 )6 :ϕ : Γ(x − y) : ϕ :.
2 L L
Use
x 3 y 3 x 3 y 3
:ϕ ::ϕ :=: ϕ ϕ :
L L L L
x − y x 2 y 2
+ 9G :ϕ ϕ : + quadratic term + constant
L L L
and write
x − y
Z x 2 y 2
G Γ(x − y) : ϕ ϕ : dxdy
L L L
Z x 4 Z y
= dx : ϕ : dy G Γ(y) + irrelevant.
L L
Thus a) and b) together give
Z Z y Z
4−d 4 2 2−d
−36L dx : ϕ(x) : Γ(y) + 2L G Γ(y) = −b(L) : ϕ4 : dx
L
zn+1 = zn + O(λ2n )
rn+1 = L2 rn + O(λ2n )
λn+1 = λn − b(L)λ2n + O(λ3n ) (15.5)
Ven+1 = LVen + O(λ2n )
kLVen k ≤ L−2 kVen k
But
d4 p eip1 r −p2
Z
G(r) = e (15.6)
(2π)4 p2
√
Z 4 ip1 Z 3 − p2
dp e 2 2 dp e
= r−2 4 2
e−p /r r→∞
−→ r−2 p .
(2π) p (2π)3 2 p2
So 2 2
Z ∞
4
−2s −s
2 dρ −ρ 1
lim r e G(e r) = ρe = .
r→∞ 0 2π 2 2π 2
Hence
b(L) = 9Aπ −4 log L ≡ β2 log L
A = area of unit sphere in R4 .
2
Remark. β2 is universal, it did not depend on our cutoff e−p . We could have used any
cutoff χ(p) ! This is a deep fact, see below.
108
Summary. For d = 4, again we find a r0 (λ) such that
Z Z
R (r0 : ϕ : +λ : ϕ4 :) = Vn
n 2
15.5 d<4
Now λn+1 = L4−d λn + O(λ2n ) increases, so sooner or later λn is big enough so we can not
trust our perturbative analysis that holds only if λn is small enough.
It is very instructive to pretend that d is a continuous variable and take d = 4 − ,
small. Then
λn+1 = L λn − b(L)λ2n + O(λ3n ).
Let us look for a fixed point λ∗ = a + O(2 ). We get 1 = L − b(L)a + O(2 ) or
a = log L
= β2−1 above (we used b(L) = b(L) +O() = β2 log L + O()). Thus a
b(L)
d=4− d=4
fixed point would be
λ∗ = β2−1 + O(2 ).
However, now zn+1 = zn + O(λ2n ) = zn + O(λ∗2 ) increases. What went wrong ? We have
indeed a new fixed point λ∗ 6= 0 and we need a different scaling of ϕ in our RG :
d−2 d−2+η
L− 2 → L− 2 .
Thus, 1 + z0 = 1 (z0 = 0)
⇒ η = O(2 ) and
109
so λ∗ is still λ∗ = β2−1 + O(2 ).
We got also rn+1 = L2−η rn + O(λ2n ), kLVen k ≤ L−2−η kVen k.
Summary d = 4 −
Now we have a new fixed point
Z Z Z
∗ 1 ∗ ∗
H = 2
(∇ϕ) + r 2
: ϕ : +λ : ϕ4 : +Ve ∗
2
with λ∗ = β2 + O(2 ), r∗ = O(2 ), Ve ∗ = O(2 ).
Also D E 1
ϕ(x)ϕ(y) ∼
H∗ |x − y|d−2+η
η = O(2 ).
Remark d = 3 is = 1 i.e. is non-perturbative. We expect, say for block-spin RG, that
for all λ there is r(λ) such that H = (ϕ(x) − ϕ(y))2 + r(λ) ϕ(x)2 + λ ϕ(x)4 is
P P P
has only one relevant direction, namely the correlation length and no marginal ones (for
small , one sees that the marginal (∇ϕ)2 becomes irrelevant).
110
It is convenient to relate all of these effective theories to a fixed lattice Zd . For this, define
a Zd field φn (i) = L−an ϕn (L−n i) where a is a parameter to be fixed later, and let µn be
the distribution of φn , i.e.
DY E DY E
−an −n
φn (iα ) = L ϕn (L iα ) . (16.2)
µn νn
α α
µn−1 = C ∗ µn .
Hn−m = Rm Hn ∀n, m.
There exists a ∈ R and H∗ such that RH∗ = H∗ and Hn n→∞ −→ H∗ . Then all Hn are
on the unstable manifold of H∗ . Hence we got a nice picture: critical phenomena are on
MS , QFT on Mu !
1
Example ν = Gaussian measure, covariance G(p) b = p2 +m 2 (free field, mass m). Then
νn has covariance
Z Z 2 dd p
−n −n ipL−n (i−j) b nd
Gn (L i, L j) = e G(p)L eipx dx .
R d n
σ0 (2π)d
Z 2
nd
Let ρn (p) = L eipx dx and, for a = d−2 , µn has covariance Cn :
2
σ0n
dd p
Z
(2−d)n −n 1
Cn (i, j) = L eipL (i−j) 2 ρn (p)
p + m2 (2π)d
dd p
Z
ip(i−j) 1 n
= e ρn (L p)
p2 + L−2n m2 (2π)d
111
where
d
2 sin pµ/2 2
Z 2 Y
n ipx
ρn (L p) = e dx = ≡ ρ(p)
σ00 µ=1
p µ
is n-independent. Here Cn are covariances that parametrize the Gaussian part of the
unstable manifold of the Gaussian fixed point Cn=∞ of the block spin RG. They have
correlation length Ln m−1 .
DY E
G(n) (x1 , · · · , xN ) = LN an φav (L(Ln−1 xi )) (1 + O(L−n ))
H∗
i
= G (n−1)
(x1 , · · · , xN )(1 + O(L−n ))
since (Cφ)(x) = La φav (Lx) and RH∗ = H∗ . Thus, one expects the limit in (16.4) to exist.
The distributions G(x1 , · · · , xN ) are correlations of a QFT (xi ∈ Rd now). What are the
effective field theories νm or µm or Hm of this QFT ?
112
Proof This is just bookkeeping. By definition (16.1), (16.2),
N
DY E Z
φ(iα ) = dx1 · · · dxN χ(xα ∈ σimα )G(x1 , · · · , xN )LN (d−a)m
µm
α=1
where σimα = L−m cube at L−m iα . Write the integral as a Riemann sum and use the
definition (16.4) of G:
Z Y X N
DY E
−N dn
dxα χG = lim L χ(xα ∈ σimα )LN (d−a)m LN an φ(Ln xα ) .
n→∞ H∗
{xα ∈L−n Zd } α=1
1 X 2 L−2n m2 X
Hn = φ(x) − φ(y) + φ(x)2
2 2 x
hxyi
i.e.
113
on Zd . We know that
2n( d−2 )
D
n n
E 1
L 2 φ(L x)φ(L y) −→
n→∞ (x, y).
Cn −∆Rd + m2
From RG we find
1
Rn−` Hn n→∞−→ (φ, G−1 ` φ) = H`
∗
2
dd p
Z
1
G` (x, y) = eip(x−y) 2 ρ(p) (x, y ∈ Zd )
Rd p + L−2` m2 (2π)d
which are on the unstable manifold : RH`∗ = H`−1
∗
. Pictorially,
So, in general, we should look for a 1-parameter family of Hamiltonians HLn such that
Rn−` HLn converge as n → ∞ for all m. It is easy to see that HLn must tend to MS as
n → ∞ and Rn−` HLn to Mu .
As an example we consider the ϕ4 QFT in various dimensions.
16.3 ϕ4d
We consider a momentum cutoff for simplicity and let
1 −z R :(∇ϕ)2 :−ρ R :ϕ2 :−g R :ϕ4 :
dν (ϕ) =
e Λ Λ Λ dµG .
Z
b (p) = 2 1 2 χ(p) and
R R R
Let z Λ : (∇ϕ)2 : +ρ Λ : ϕ2 : +g Λ : ϕ4 := W (ϕ); here8 G p +m
−p2
let, e.g., χ = e . We ask : are there functions of z , ρ , g such that ν converges to a
8
We put m2 into the covariance for simplicity : we are concerned with the UV-problem, not the IR.
114
measure ν on S 0 as → 0 ? We know that λ = 0 is a solution, and ask if a non-Gaussian
ν is possible. The effective measures are defined as above, only easier : Write, for ` >
1 1 1
Gb (p) = χ(p) = χ(`p) + χ(p) − χ(`p)
p2 + m 2 p2 + m 2 p2 + m2
≡ G
b` (p) + Γ
b` (p)
and define Z
() −W (ϕ+z)
dν` (ϕ) = e dµΓ` (z) dµG` (ϕ)
which has cutoff `. To connect to Statistical Mechanics and RG, we need to scale as in
(16.2) :
(d−2)
φ (x) = 2 ϕ(x)
d−2
(we take a = 2
to anticipate what comes later). Then ν becomes µ :
2−d
2 φ( · ))
dµ (φ) = e−W ( dµC (φ) ≡ e−V (φ) dµC
1
where C
b (p) =
p2 +2 m2
χ(p), has UV-cutoff 1, and is Gaussian. (Note the IR cutoff 2 m2 ).
Here Z Z Z
2 2 2 4−d
V (φ) = z : (∇ϕ) : + ρ : ϕ : + g : ϕ4 :
note the powers of (so called canonical dimensions of mass, coupling). Then we have
()
V` ≡ R`/ V
where RL is as before :
Z 2−d
2 φ( · )+Z)
RL V = − log e−V (L L dµΓ (Z)
1
Γ(p)
b = p2
(χ(p) − χ(Lp)). Let us put
r = 2 ρ , λ = 4−d g .
As before, we could study the iteration rn , zn , λn → rn+1 , zn+1 , λn+1 but here we can
also do it infinitesimally, i.e. as a differential equation. We put Ln = es and get, for
Res V = z (s), r (s), λ (s), Ve (s) ≡ V (s):
dz
= O(V 2 )
ds
dr
= 2r + O(V 2 )
ds
dλ
= (4 − d)λ − β2 λ2 + O λ3 , λ(z + r + Ve ), (z + r + Ve )2
ds
dVe
= MVe + O(V 2 )
ds
115
where Z
d 2−d
MV (φ) =
e Ve (e 2 s φ(e−s ·) + Z)dµΓ (Z).
ds 0
We want to choose r (0), z (0), λ (0) such that lim Ve−t (t + s) ≡ V eff (s) exists, ∀s ∈ R.
t→∞
d=2. Take r = 0, λ = 2 g, z = 0.
Put
λ (s) = 2 e2s gt (s), r (s) = 2 e2s ρt (s), = e−t .
We get
dg dρ
= O(e−2(t−s) g 2 ), = O(e−2(t−s) g 2 ).
ds ds
Hence the limits lim gt (t + s), limt→∞ rt (t + s) exist. This is a super renormalizable the-
t→∞
ory : the “bare” coupling g can be taken -independent and z , ρ also.
d=3. Now λ (s) = es gt (s) and we have a “resonance” in the ρ-equation:
dg
= O(e−(t−s) g 2 )
ds
dρ
= O(g 2 ).
ds
dr
More precisely, write ds
= 2r + γλ2 + O(λ3 ), then,
dρ
= γg 2 + O(e−(t−s) g 3 ).
ds
Since g stays almost constant, we get
g = g, ρ = γg 2 log
this is a “mass counterterm” of O(g 2 ), coming from the graph that di-
verges logarithmically in UV.
dλ
= −β2 λ2 + O(λ3 )
ds
λ(0)
λ(s) ∼ .
1 + β2 sλ (0)
116
If λ (0) stays bounded in , as we have to assume in our perturbative analysis, then
−→ 0. Thus the only perturbative theory (= the one where the coupling constant
λ(s) s→∞
remains small on all scales) is the Gaussian. Same holds for
dλ
d > 4. ds
= −(d − 4)λ + O(λ2 ) i.e. λ(s) → 0.
One might hope that, by taking λ (0) large enough, things might change. E.g. there
might be another fixed point λ∗ for the RG, i.e. the flow in λ would be
λ=0 λ∗
Unfortunately, this is unlikely to happen : in d > 4 it is rigorously proven that for all
z(), r(), λ() the only → 0 limits are Gaussian. For d = 4 there are strong theoretical
and numerical arguments supporting the same conjecture.
Consider the d = 3 QFT constructed above. This has mass ∼ m2 , coupling λ. What
about m → 0 ? Above we could have taken m = 0 and started with r , λ to produce an
effective Hamiltonian on scale 1 of the form
Z Z Z
z : ∇ϕ : +r : ϕ : +λ : ϕ4 : +Ve
2 2
with various values of r. To solve for the IR we would need to keep iterating the RG. Now
λ increases and eventually, if we really choose r = r(λ) = critical point, we would tend
to the nontrivial fixed point H∗ . Thus, UV is controlled by the Gaussian fixed point, IR
by the non-Gaussian.
We could construct a scaling limit at H∗ : this would be a scale-invariant QFT with non-
Gaussian IR and UV. H∗ has (at least) one relevant direction : the correlation length.
We could construct, as above, a massive QFT corresponding to the unstable manifold of
H∗ . We expect the unstable manifold of the Gaussian fixed point H0∗ to be 2-dimensional
(r, λ), and the unstable manifold of H∗ to be one dimensional. So the flow is
117
r
Mu (H0∗ )
H0∗ λ
H∗
Mu (H∗ ) Ms (H∗ )
Thus, a “typical” QFT is a typical point of the r, λ plane with Gaussian UV and finite
correlation length. The ones on Ms (H∗ ) have non-Gaussian IR, Gaussian UV and the
ones on Mu (H∗ ) have non-Gaussian UV, ξ < ∞. (We should really include also the
r = ∞ (ξ = 0) “high-temperature” fixed point).
dλ λ (s)
= |β2 |λ2 + O(λ3 ) ⇒ λ (0) =
ds 1 + |β2 |sλ (s)
i.e. if = e−t , s = t, and we want to fix λe−t (t) = g, then the “bare coupling” λ is
g
λ = −→ 0
1 + |β2 | log −1 g →0
dλ
= β(λ) = β2 λ2 + β3 λ3 + β̃(λ), β̃ = O(λ4 ).
ds
118
So
Z g
dλ
= log −1
λ β(λ)
Z g Z g
1 1 1 β3
= dλ = dλ − + O(1)
λ β2 λ2 1 + β3
β2
λ + O(λ2)
λ β2 λ2 β22 λ
1 1 β3 g
= − − 2 log + O(g) = log −1
β2 λ β2 g β2 λ
g
⇒ λ = gβ3
.
1 + β2 g log − β2 log log −1 + O(1)
−1
This is how the bare coupling should be chosen to have a continuum limit.
Remark. We have seen that β2 is universal (independent on regularization). It can be
shown that β3 is also. (β4 is not). They are intrinsic, depend only on the continuum limit
(which does not depend on the regularization). An example of asympotically free theory
is the Ginzburg-Landau model with λ negative. This is not stable, but can be rigorously
defined by analytic continuation. Since λ = −g, g > 0 we get dg ds
= β2 g 2 + · · · β2 > 0.
More interesting one is the SU (N ) gauge theory. There one has two couplings z and λ,
both marginal, with
dz
= γλ2 + · · ·
ds
dλ
= β3 λ3 + · · ·
ds
2 Rt
so λ(s)2 ≈ 1+β3 gg2 (t−s) where = e−t , λ2 (t) = g 2 and z (t) − z (0) = γ 0
λ(s)2 ∼ γ
β3
log t so
z (0) must diverge as − βγ3 log log −1 .
Remarks.
1. Thus, we expect to find QFT’s at fixed points and on their unstable manifolds. For
Ginzburg-Landau model, d < 4 we find several non-Gaussian ones. For d ≥ 4 the
fixed points and unstable manifolds are Gaussian.
2. Asymptotically free theories have Gaussian fixed points with marginally unstable
non-Gaussian direction. Here we have a non-Gaussian continuum limit.
Examples : Non Abelian gauge theories with not too many fermions (QCD is one).
Non-linear σ-models in d = 2 (e.g. O(N ), N > 2 spin systems we discussed earlier).
Certain d = 2 fermionic models (Gross-Neveu model).
119
variable like r in Ginzburg-Landau model that would provide a mass (correlation
length). Nevertheless, they are massive (ξ < ∞) : ξ with be ∼ e−const/λ (so called
2
∼ e−const/λ .
Hence, the dependence on cutoff is ∼ e−const/λ , very small for λ small ( λ1 = 137 for
QED). We say that all the above theories are good effective theories for distances
>> (in practice, energies << 1015 GeV). For smaller distances one needs to find
a new theory (say string theory).
References
[1] Aizenman M., Translation invariance and instability of phase coexistence in the
two-dimensional Ising system, Commun. Math. Phys. 73 (1) (1980) 83–94.
[2] Aizenman M. and Wehr J., Rounding effects of quenched randomness on first-order
phase transitions, Commun. Math. Phys. 130 (3) (1990) 489–528.
[3] Aizenman M., Barsky D., and Fernández R., The Phase Transition in a General
Class of Ising-Type Models is Sharp. J. Stat. Phys. 47 (1987) 343-374.
[4] Baxter Rodney J., Exactly solved models in statistical mechanics, London, Academic
Press Inc. [Harcourt Brace Jovanovich Publishers] (1982).
[5] Bricmont J. and Kupiainen A., Phase Transition in the 3d Random Field Ising
model, Commun. Math. Phys. 116 (1988) 539–572.
[6] Bricmont J. and Fontaine J.R., Perturbation about the mean field critical point,
Commun. Math. Phys. 86 (1982) 337–362.
120
[7] Bricmont J. and Kupiainen A., High temperature expansions and dynamical systems,
Commun. Math. Phys. 178 (1996) 703–732.
[9] Dobrushin R.L., Existence of a phase transition in the two-dimensional and three-
dimensional Ising models, Soviet Physics Doklady 10 (1965) 111–113.
[10] Dobrushin R.L., The description of a random field by means of conditional proba-
bilities and conditions of its regularity, Theory Prob. Appl. 13 (1968) 197–224.
[12] Dobrushin R.L., Gibbs state describing coexistence of phases for a three-dimensional
Ising model, Theory Prob. Appl. 17 (4) (1972) 582–600.
[13] Dobrushin R.L., “ Gaussian Random Fields – Gibbsian Point of View, Multicom-
ponent Random Systems”, in “Advances in Probability and Related Topics”, 6, New
York, Dekker (1980) 119–51.
[14] Dobrushin R.L. and Shlosman S.B., “Absence of breakdown of continuous symmetry
in two-dimensional models of statistical physics”, Commun. Math. Phys. 42 (1975)
31.
[15] Fernández R., Fröhlich J. and Sokal A.D., Random Walks, Critical Phenomena, and
Triviality in Quantum Field Theory, New York, Springer Verlag (1992).
[16] Fisher M.E., Critical temperatures of anisotropic Ising lattices, II. General upper
bounds, Phys. Rev. 162 (3) (1967) 480–85.
[17] Fröhlich J. and Pfister C.E., On the absence of spontaneous symmetry breaking
and of crystalline ordering in two-dimensional systems, Commun. Math. Phys. 81
(2) (1981) 277–98.
[18] Fröhlich J., Simon B. and Spencer T., Infrared bounds, phase transitions and con-
tinuous symmetry breaking, Commun. Math. Phys. 50 (1) (1976) 79–95.
121
[20] Gawedzki K., Kotecký R. and Kupiainen A., Coarse-Graining Approach to First-
OrderPhase Transitions, J. Stat. Phys. 47 (1988) 701–724.
[21] Glimm J. and Jaffe A., Quantum Physics, A Functional Integral Point of View, New
York, Springer (1981).
[24] Itzykson C. and Drouffe J.-M., Statistical field theory, Volume 1 : From Brownian
motion to renormalization and lattice gauge theory, Cambridge University Press,
(1989) ISBN 0521408059.
[25] Kelley J.L., General Topology, Graduate Texts in Mathematics, New York, Springer
Verlag (1955).
[26] Klein A., Landau L.J. and Shucker D.S., On the absence of spontaneous breakdown
of continuous symmetry for equilibrium states in two dimensions, J. Stat. Phys. 26
(3) (1981) 505–12.
[27] Kosterlitz J.M. and Thouless D.J., “Ordering, metastability and phase transitions
in two-dimensional systems”, Journal of Physics C : Solid State Physics 6 (1973)
1181–1203.
[28] Kotecky R. and Shlosman R.B., First-order transitions in large entropy lattice mo-
dels, Commun. Math. Phys. 83 (1982) 493–515.
[29] Lanford III O.E. and Ruelle D., Observables at infinity and states with short range
correlations in statistical mechanics, Commun. Math. Phys. 13 (1969) 194–215.
[30] McCoy Barry M., Advanced Statistics Mechanics, International series of monographs
on physics 146, Oxford University Press (2010).
[33] Peierls R., On the Ising model of ferromagnetism, Proceedings of the Cambridge
Philosophical Society 32 (1936) 477–481.
122
[34] Pirogov S. and Sinai Ya. G., Phase diagrams of classical lattice systems, Theoretical
and Mathematical Physics 25 (1975) 1185–1192 ; 26 (1976) 39–49.
[35] Reed M. and Simon B., Methods of Modern Mathematical Physics 1, Functional
Analysis, New York (N.Y.), Academic Press (1980).
[36] Reed M. and Simon B., Methods of Modern Mathematical Physics 2, Fourier Ana-
lysis, Self-adjointness, New York, Academic Press (1975).
[37] Rudin W., Real and Complex Analyis, New York, McGraw-Hill (1987).
[38] Seiler E., Gauge Theories as a Problem of Constructive Quantum Field Theory and
Statistical Mechanics, Lecture Notes in Physics 159, Berlin, Springer (1982).
[39] Seneta E., Non-negative Matrices and Markov Chains, 2nd ed., New York, Springer
Verlag (1981).
[40] Sinai Y.G., Theory of Phase Transitions : Rigorous results, Pergamon Press, Oxford
(1982).
[41] Simon B., Functional Integration and Quantum Physics, New York, Academic Press
(1979).
[42] Simon B., The Statistical Mechanics of Lattice Gases, Volume 1, Princeton, Prince-
ton University Press, (1993).
[43] Van Enter A., Fernández R., Sokal A.D., Regularity properties and pathologies
of position-space renormalization-group transformations : Scope and limitations of
Gibbsian theory, J. Stat. Phys. 72 (1993) 879–1167.
[44] Wilson, K. (1975). The renormalization group: Critical phenomena and the Kondo
problem, Reviews of Modern Physics 47 (1975) 773-840.
123