Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

Two Dimensional S-matrix Bootstrap


Pedro Vieira1,2

1
Perimeter Institute for Theoretical Physics, Waterloo, Ontario N2L 2Y5, Canada
2
ICTP South American Institute for Fundamental Research, IFT-UNESP, São Paulo, SP Brazil 01440-070

Abstract

Four 75 minute lectures at TASI 2019 on S-matrix things in two space-time dimensions.
Some sections are copy/pasted from works with many amazing collaborators who I thank for amazing
times. I will add references at some point I hope! :)


pedrogvieira@gmail.com
Contents
1 Lecture 1. Kinematics. 3
1.1 Variables and a quick counting . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Higher Dimensional Parentheses I . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 2 → 2 scattering of identical particles . . . . . . . . . . . . . . . . . . . . . . 4
1.4 2 → 2 scattering of different particles . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Higher dimensional parentheses II . . . . . . . . . . . . . . . . . . . . . . . . 7
1.6 A sort of multi-Regge limit for 2 → n scattering . . . . . . . . . . . . . . . . 9

2 Lecture 2. Tree Level and One Loop Games, Integrability. 12


2.1 An important S-matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2 Finding Sine-Gordon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Bullough-Dodd and Quantum No-Particle Production . . . . . . . . . . . . . 16
2.4 3 → 3 in 2D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5 Ising With Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.6 Advanced Sine-Gordon Topics . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3 Lecture 3. Bootstrap. Complex Analysis. 20


3.1 Complex Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2 Sine-Gordon and other friends . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.3 Magnetic Ising Field Theory Parentheses . . . . . . . . . . . . . . . . . . . . 23

4 Lecture 4. More Advanced Bootstrap, Flux tubes, etc. 27


4.1 Higher Dimensional Parentheses III . . . . . . . . . . . . . . . . . . . . . . . 27
4.2 Flux Tubes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.3 General Comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

2
1 Lecture 1. Kinematics.

1.1 Variables and a quick counting


We are in two dimensions; we have time t and space x and the on-shell two-momenta of
particles of mass m can be written in several useful ways, e.g.
m 1 1 p
pµ = m(cosh(θ), sinh(θ)) = a + ,a − = ( m2 + p2 , p) . (1)
2 a a
• Show that boosts act very simply in terms of the so called hyperbolic rapidity as
θ → θ + Λ and in terms of the light-cone momenta as a → eΛ a.

So a scattering amplitude will be a function of n incoming variables (e.g. θ1 , . . . , θn ) and cor-


responding masses m1 , . . . , mn and m out-going momenta and masses. (In the all-incoming
notation, outgoing particles have negative energy.) Since we can boost all variables and
since we have momentum and momentum and energy conservation which fixes two of the
momenta in terms of the n + m − 2 other momenta, a scattering amplitude involving a total
of N = n + m particles is a function of N − 3 variables. In a two-to-two process this is pretty
obvious: Since space is just a line, the scattering process just depends on the centre of mass
energy of the two particles approaching head-on.
In more group theoretical terms we would sat that the dimension of the Lorentz group is
3 in two dimensions (boost plus time and space translations) so we go from 2N − N variables
(N on-shell two vectors) to N − 3 variables once we use the Lorentz generators.

1.2 Higher Dimensional Parentheses I


• Argue that in 3 + 1 dimensions a scattering process would be a function of 3N − 10 variables. What
is the physical meaning of those variables for a two-to-two process for which N = 4? What about in
2 + 1 dimensions?

• Lets recover these results in another Lorentz invariant way. An amplitude would be a function of the
Mandelstam invariants
sij ≡ pi · pj (2)
When i = j they simply give m2i by the on-shellness condition and since the definition is symmetric
there are a priori N (N − 1)/2 such invariants. However, they are not all independent.
– Show that by momentum conservation we can reduce N (N − 1)/2 straight away to N (N − 3)/2
variables.
– Show that this agrees with the four dimensional linear relation 3N − 10 when N = 4 and when
N = 5 and also agrees with the linear relation you derived in three dimensions when N = 4.
This is because, for such small number of particles the dimensionality constraints are now yet
important.
– Next we take into account the dimension of space-time. Explain why any of the momenta
pD+1 , . . . , pN −1 can be expressed in terms of a linear combination of p1 , . . . , pD . (We do not
worry about pN since we can get that last one by momentum conservation.) Clearly these
constraints will play a role as long at the set pD+1 , . . . , pN −1 is not empty, i.e. N ≥ D + 2 thus
explaining the previous remark.

3
– Consider the coefficients αjA in X
pj = αjA pA
A
with A = 1 . . . , D and j = D + 1, . . . N − 1. By contracting this relation with well chosen
momenta, argue that each coefficient αjA can be expressed solely in terms of s1j , . . . s(D−1)j and
sAB with A, B = 1, . . . , D − 1. Clearly any Mandelstam will now be given in terms of these so
we are left with (D − 1)(N − 1 − D) + D(D − 1)/2 variables at this point.
– Show that this is indeed what we expect from group theory and show that this nicely reproduces
all three examples (D = 2, 3, 4) considered above.

To summarize we have

 N (N − 1)/3 , N <D+1
# of independent Mandelstam variables = (3)

(D − 1)N − D(D + 1)/2 , N ≥D+1

• Plot this number for various dimensions.


• For N = 4 show that the three Mandelstam variables s = (p1 + p2 )2 , t = (p1 + p3 )2 and u = (p1 + p4 )2
are related by
X 4
s+t+u= m2j (4)
j=1
so that we have indeed only two independent variables. If all particles are identical the right hand
side is just 4m2 . As will be recalled in detail below, in two dimensions, in this equal mass case we
have also u = 0 (or t = 0 – for identical particles these are equivalent) so that we end up with a single
Mandelstam variable s with the other simply given by t = 4m2 − s.
• Google other Mandelstam invariants counting arguments or come up with your own.
• Repeat the counting of invariants for non-relativistic theories.

1.3 2 → 2 scattering of identical particles


If two particles with rapidities θ1 , θ2 and the same mass scatter against each other to pro-
duce two particles of the same mass in the final state then, for those final particles, energy
and momentum conservation fix {θ10 , θ20 } = {θ1 , θ2 } or {θ10 , θ20 } = {θ2 , θ1 }. If particles are
distinguishable these are two different possibilities of course. If they are identical they are
the same option. So when two identical particles scatter in two dimensions their momenta
in the initial and final state are the same.
In particular one of the Mandelstam invariants is thus zero in two dimensions, u =
(p02
− p2 )2 = 0. The other two thus obey s + t + 0 = 4m2 so that t = 4m2 − s and everything
can be expressed in terms of s = (p1 + p2 )2 , the centre of mass energy (squared) of the
process.
Note in particular that because of this, there is no big difference between the connected
and disconnected component in the S-matrix so in two dimensions we can actually jump
between S and T (in S = 1 + iT ) rather straightforwardly, not in a fancy distributional
sense. Indeed
0 0
out hθ1 , θ2 |θ1 , θ2 iin = δ(p1 − p01 )δ(p2 − p02 ) + iδ (2) (p1 + p2 − p01 − p02 )T (θ1 , θ2 )
 iT (θ) 
= δ(p1 − p01 )δ(p2 − p02 ) S(θ) ≡ 1 + (5)
sinh(θ)

4
since by boost invariance the S-matrix (and the T-matrix) should only depend on the differ-
ence of rapidities θ = θ1 − θ2 .

• Work out the Jacobian in the delta functions leading to the precise expression (5).

Recall that when we do Feynman diagrams we compute the amplitude iT . The probability
of getting that two particle final state given that initial final state is however given by the
full S-matrix simply as
Prob(θ) = |S(θ)|2 . (6)
This probability is constrained by |S(θ)|2 ≤ 1 for any real θ since there are in principle other
possible final states.

• How does unitarity


|S(θ)|2 ≤ 1 (7)
look like in terms of T (θ)?

• Show that
s = 4m2 cosh2 (θ/2) . (8)

• What happens under θ → iπ − θ? What is this transformation?

• Suppose we have a gapped theory with a single particle of mass m. Argue that in this
case we expect |S(θ)|2 = 1 for real θ with |θ| ≤ 2 arccosh(3/2).

• In which situation(s) would you expect θ for any |θ| ≤ 2 arccosh(4/2), which corre-
sponds of course a bigger energy range?

• As we will recall before amplitudes have at least two cuts associated to two particle
processes in the s-channel and in the t-channel. They are given by s > 4m2 and
t > 4m2 respectively. Show that the map (8) maps the full complex s plane minus
these two cuts into a strip in θ where Im(θ) ∈ [0, π]. Draw this strip and identify there
the top and bottom of the s- and t-channel cuts.

• [Harder] Find a map which sends the full s plane minus the two cuts into the inside
of the unit disk. We will use this map later.

1.4 2 → 2 scattering of different particles


Consider now particles of different mass m1 , m2 in the initial state producing two particles of
different mass m3 , m4 in the final state (from the s-channel point of view; for other channels
we should re-shuffle the masses). Then we have

s + t + u = m21 + m22 + m23 + m24 (9)

5
Figure 1: Checking (10)

In two dimensions, the relation


X X
0 = stu+s(m21 +m22 )(m23 +m24 )+t(m21 +m23 )(m22 +m24 )+u(m21 +m24 )(m22 +m23 )− m2a m2b m2c + m6a
a≤b≤c a
(10)
is also true, see figure 1

• What would be a more elegant way of derivation this relation?


• What happens when all masses are equal ma = m? Show that the curve (10) factorizes
into three constrains which we knew already.
• When m1 = m4 = m and m2 = m3 = m0 show that

u=0 or u = −s − (m02 − m2 )2 /s + 2(m2 + m02 ) (11)

Interpret these two relations as a reflection and transmission process. Which one is
which? What happens to crossing symmetry in this case? What happens when one
mass is much larger than the other? What do we get at very high energy? When we
solve for s(u) in the second possibility we have two solutions. What are these?

Physical processes happen when s > smax ≡ max((m1 + m2 )2 , (m3 + m4 )2 ) since we need
to have enough energy to produce both the initial and the final state. Similarly, we have a
physical t channel process when t > max((m1 + m3 )2 , (m2 + m4 )2 ).

• What is the physical range for u?

Figure 2 summarizes this in the famous Mandelstam triangle.

• Explain how the right figure in figure 2, for three generic different masses degenerate
into figures 3a and 3b represent the two special cases (all masses equal or two pairs)
discussed above.

Finally, suppose m1 is the lightest particle and suppose there is no symmetry that forbids
two particles of type m1 to show up in this scattering process. Then the value smin = (2m1 )2
is also interesting. It describes the first energy where we can have two particle states in the
s-channel. Amplitudes will have a cut there as well although energies will not be physical
yet. Regular unitarity as in (7) is imposed for s > smax only but for smax > s > smin we still
have a constraint albeit a less known one called extended unitarity. Below (2m1 )2 we have
no discontinuities. We can have bound-state poles as we will see at length below.

6
10

10

5
s
5

u t
0
0

-5

-5

-10

-5 0 5
-5 0 5

Figure 2: Left: The Mandelstam triangle is the region where all Mandelstam variables are below
their corresponding two-particle threshold: s < max((ma + mb )2 , (mc + md )2 ) ∧ t < max((ma +
mc )2 , (mb + md )2 ) ∧ u < max((ma + md )2 , (mb + mc )2 represented by the white region √
in this figure.
(The y axis is s and the x axis is given by x = (s + 2t − m2a − m2b − m2c − m2d )/ 3 as in the
next figures.) Right: The blue lines represent the two dimensional constraint (10) described in the
text. The round shape inside is clearly unphysical while the outer portions are physical. Nicely, in
higher dimensions, the outer regions bordered by the blue lines are filled in and describe the full
physical region. The boundary correspond to scattering angle θ = 0, π, the only possible values in
two dimensions.

1.5 Higher dimensional parentheses II


The Mandelstam plane provides us with a very useful depiction of the (real sections of the) interrelations
between the three Mandelstam variables u, s, t. The first important object in this plane is the Mandelstam
triangle.
Consider a two-to-two process involving particles with momenta pa , pb , pc , pd associated to particles of
mass ma , mb , mc , md . We define the three Mandelstam invariants √ s = (pa + pb )2 , t = (pa + pc )2 and
2
u = (pa + pd ) . If particles pa , pb are the two incoming particles then s is the centre of mass energy of the
scattering process. The same is true if the incoming particles are particles pc , pd . In either of these cases the
process can only be physical if we have enough energy to produce both the initial and final state that is for
s ≥ max((ma + mb )2 , (mc + md )2 ). Of course, the same √ scattering amplitudes can describe other channels. If
pa , pc (or pb , pd ) are the two incoming particles then t is the centre of mass energy of the scattering process
and similarly for u so the physical conditions in those cases would read t ≥ max((ma + mc )2 , (mb + md )2 )
and u ≥ max((ma + md )2 , (mb + mc )2 ). The three inequalities are depicted by the shaded pink regions in
figure 2. The white region is the Mandelstam triangle.
To be in a physical region we thus need to be in the pink region. This is necessary but not sufficient.
We need to have enough energy but we also need to scatter at a real angle.

• For incoming particles pa , pb show that the scattering angle is given by


(s + m2a − m2b )(s + m2c − m2d ) + 2s(t − m2a − m2c )
cos(θab ) = p . (12)
((s − m2a − m2b )2 − 4m2a m2b )((s − m2c − m2d )2 − 4m2c m2d )

For any left hand side between −1 and +1 corresponding to a real angle, and for any s in the physical range
this equation determines a physical t. (Of course, u = m2a + m2b + m2c + m2d − s − t is automatically fixed.)

7
10 10
s-channel 11 ! 22

5 5

Forward
0 0
u-channel t-channel 12 ! 12

Ba
ck
wa
-5 -5

rd
-5 0 5 -5 0 5

(a) (b)
Figure 3: (a) Physical regions (in dark blue) for a two-to-two process with all external particles
of identical mass. The boundary of these regions correspond to scattering angle 0 or π and can
thus also be identified with the physical scattering ‘regions” (or better scattering lines) in two
dimensions. (b) Physical regions for a process where the masses are equal in pairs. The darker
blue region is the process m1 m1 → m2 m2 while the two lighter regions (which are equivalent for
identical particles) correspond to the m1 m2 → m1 m2 channels. The boundary of these regions
are again the two dimensional physical lines. The boundaries of the lighter blue regions are not
identical: one corresponds to forward scattering; the other to backward scattering.

The set of physical s and t determined in this way determine the physical region in the s-channel. The other
channels are treated similarly with
cos(θac ) = RHS of (12)mb ↔mc , s↔t , cos(θad ) = RHS of (12)mb ↔md , s↔u (13)

• Show that when all masses are equal (to m) then (12) reduces to the famous relation
2t
cos(θ) = 1 − . (14)
4m2 − s
so that the physical region in the s-channel is simply s > 4m2 and 4m2 − s ≤ t ≤ 0 represented by
the top blue region in figure 3a.

In two dimensions the angle ought to be 0 or π so that we have either t = 0 or t = 4m2 − s, i.e. u = 0.
These two conditions (t = 0 and u = 0) are equivalent if the external particles are indistinguishable so in
that case we can pick either; in the main text we took u = 0. Note that these two conditions are nothing
but the boundary of the darker green region. Similarly we could study all other channels which we can
simply obtain by relabelling the Mandelstam variables in (14). The two extra physical regions are the other
two blue regions in the same figure 3a. Note that the boundary of all these green regions can be written
concisely as stu = 0 which is nothing but the constraint obtained in the main text from the two dimensional
constraint (10) and represented by the dashed lines in figure 3a.

8
Finally, we come to the more interested case where the external masses are only pairwise equal:
√ ma =
mb = m1 and mc = md = m2 . This same configuration can describe√ the 11 → 22 process (with s being
the centre of mass energy), and the 12 → 12 processes (with t being the centre of mass energy). For the
first case we use (12) to get
−2m2 − 2m22 + s + 2t
cos(θ11→22 ) = p 1 . (15)
(s − 4m21 ) (s − 4m22 )
The physical region corresponding to −1 ≤ cos θ11→22 ≤ +1 is now a more interesting curved region,
represented by the darker blue region in figure 3b. Again, in two dimensions we can only have backward
or forward scattering so t must saturate one of these inequalities. If the particles of the same mass are
indistinguishable then both solutions are equivalent as before. We can thus take θ11→22 = 0 without loss of
generality leading to q
s 1
t, u = m21 + m22 − ± (s − 4m21 )(s − 4m22 ) . (16)
2 2
and reproduce in this way the results below (??) in the main text. In the t-channel particles ma = m1 , mc =
m2 are incoming so we are scattering a odd particle against an even particle. Then we use the first relation
in (13) which now reads
2ts
cos(θ12→12 ) = 1 + (17)
−2m21 (m22 + t) + (m22 − t) 2 + m41
√ √
Note that in our convention it is t (and not s) who is the center of mass energy in this channel.1 The
physical region corresponding to | cos(θ12→12 )| ≤ 1 is now represented by the lighter blue region in figure 3b.
The two dimensional conditions that the angle is 0 or π are now quite different. The former corresponds to
forward scattering and is obtained by s = 0 (obtained by equating the RHS of (17) to +1) while the later
corresponds to backward scattering and yields the more involved relation

2 2
 m21 − m22 2
s = 2 m1 + m2 − t − (18)
t
(obtained by equating the RHS of (17) to −1.) Note that this relation is nothing but (16) if we solve for s.
In other words, these two configurations are simply related by crossing symmetry s ↔ t.
To summarize: the boundary of the physical regions are now given by the black solid line and by the black
dashed curve in figure 3b. Crossing u ↔ t at s = 0 relates the left to the right of the straight line. Crossing
symmetry also relates the top to the bottom branch of the hyperbolic looking curve. In two dimensions
these two curves (the hyperbola and the straight line) are independent while in higher dimensions they are
smoothly connected (by moving in angle space).

1.6 A sort of multi-Regge limit for 2 → n scattering


We conclude this kinematics’ analysis by considering a 2 → n − 2 process in two dimensions
involving a single type of particle of mass m. It is a high energy process where a very
energetic particle scatters against a particle at rest and produces a shower or very energetic
particles albeit with a clear hierarchy between them as depicted in figure 4.
In detail, we can introduce a convenient multi-Regge limit where one incoming particle is
at rest, with light-cone momenta p1 = m(1, 1), and n−3 outgoing particles are very energetic
with
pj = (p+ −
j , pj ) = −m(x
j−2
, 1/xj−2 ) , j = 3, . . . , n − 1 , (19)
1
Recall footnote ?? when comparing the results that follow to the main text.

9
2 3 ... n 1

p+
1 = +m ,
p+
2 ' m
p+
j = m xj 2
, j = 3, . . . , n 1
n 1 p+
n ' +m xn 3
.

Figure 4: The precise expression


P for
P the+momenta of particles 2 and n are given by solving energy-
momentum conservation p+ j = 1/pj = 0. In the multi-Regge limit with x  1 we have a highly
energetic particle hitting a particle at rest producing a particle which is almost at rest plus a shower
+ +
of very energetic particles. (More precisely, one finds the momenta p+ n = −pn−1 −· · ·−p3 +O(x ) '
−1
n−3 − − − − 2−n
mx and p2 = −p1 − p3 − · · · − pn−2 + O(x ) ' −m.)

where x is taken to be very large and positive, see figure 4. Note that light-cone coordinates
are precisely what the aj in (1) are; here aj = −xj−2 is negative for these outgoing particles.
The momenta p2 and pn of the remaining two particles are fixed by momentum conservation.
We can find them at any finite x and simplify the corresponding expressions at large x.

• Show that of of the two remaining particles (lets take it to be particle 2) is outgoing
and almost at rest while the other (that is particle n) is incoming and highly energetic.
The configuration is illustrated in Figure 4.

This is of course just a possible configuration, parametrized by a single variable x of a


multi-dimensional scattering space. Here is something nice about it. Consider a tree level
process. (Uff, dynamics!) At tree level, any propagator separates a subset α ⊂ {1, . . . , n}
of the external particles from its complement. Most such subsets make highly energetic
jets and thus vanishingly small propagators. The only propagators which survive in the
limit x → ∞ are the ones where particles {1, 2, . . . , j − 1} are on one side and particles
{j, j + 1, . . . , n − 1, n} on the other so that the momentum transfer is small.

• Specifically, show that the propagator (here rescaled by m2 )



 −1 if α = {j, j + 1, . . . , n} (or equivalently α = {1, 2, . . . , j − 1})
lim G(α) =
x→∞ 
0 otherwise .
(20)

10
2 4 . . .n 2
3 n 1

3 4 ... n 2

= 2 n 1
1 n

= v4 ⇥ ( 1) ⇥ v3 ⇥ ( 1) ⇥ v3 ⇥ ( 1) ⇥ v4
n 1

Figure 5: The only surviving diagrams in the multi-Regge limit are one-dimensional chains with
particles ordered along the chain. They evaluate to the product of involved vertices and (−1) per
propagator.

Hence, the only surviving tree-level Feynman graphs are one-dimensional chains with
all particles ordered. (Figure 5 shows an example of a surviving Feynman diagram.)
This is of course quite a major simplification which we will explore in the next lecture.

11
2 Lecture 2. Tree Level and One Loop Games, Inte-
grability.

2.1 An important S-matrix


After so many kinematics, we need some dynamics. So without further delay, here is a
beautiful exact S-matrix:
p p
s(s − 4m2 ) + m2∗ (m2∗ − 4m2 )
S(s) = p p (21)
s(s − 4m2 ) − m2∗ (m2∗ − 4m2 )
where m > 0 and 0 < m∗ < 2m. This S-matrix describes the scattering of the lightest
breathers of a theory known as super-symmetric sine-gordon theory. We will encounter this
theory – and this S-matrix – over and over in these lectures. For now, let us absorb some of
its nice analytic properties:

• What are the cuts and poles of the S-matrix? What is their physical meaning?

• Check that the S-matrix is purely real for s ∈ [0, 4m2 ]; this is known as real analyticity.

• Show that the discontinuity of T is positive (recall that T and S are related by (5)).

• What is the probability for these two lightest breathers to produce four breathers in
the final state?

• Show that, in θ,
sinh(θ) + i sin(α)
S(θ) = (22)
sinh(θ) − i sin(α)
and relate α to mb .

• Here are two important maps:


p p
4m2 s(s − 4m2 ) − A(A − 4m2 )
x(s) + 1/x(s) = , z(s) = p p (23)
s − 2m2 s(s − 4m2 ) + A(A − 4m2 )

with 0 < A < 4m2 . One maps the complex plane minus the the s- and t-cuts to the
unit disk. The other maps the left half of the complex s plane, minus the s-cut, to the
unit disk with the points s = 2m2 i + ix and s = 2m2 i − ix since they are related by
crossing. Which one is which?

• Which value of s is mapped to the origin of the unit disk in both maps? How would
you change that?

• Consider the map of the full plane minus the two cuts into the unit disk. How does
the S-matrix look like in that variable?

• Consider next the map of the left half plane. Show that the S-matrix becomes trivial
in this variable as soon as we appropriately pick the unit disk origin.

12
• Weak coupling corresponds to α ' 0 in (22). Explain why this indeed makes sense.
What happens to mb in this limit? Compute the first few terms of the small alpha
expansion of S. What is T to leading order? How could we get that in perturbation
theory from a simple Lagrangian?

2.2 Finding Sine-Gordon


Consider for simplicity a single real scalar in (1+1)D with mass m and interaction Lagrangian

X
2 vn
Linteraction = −m φn . (24)
n=3
n!

At tree level we can now compute all possible Tn→m processes. Let us try to cancel all
T2→n ones for fun. The first production amplitude we want to suppress is T2→3 . Setting all
particles as incoming, and using the light-cone coordinates pj = m(aj , 1/aj ) we have
1 X 1 X 1X
− 2 T2→3 = v3 v4 G(α) + v33 G(α)G(β) + v5 , (25)
m α
2 α,β
2 α
| {z } | {z } |{z}


where α, β run over disjoint two-element subsets of the set of external particles {1, . . . , 5}
and where the (rescaled) propagator takes the form
1
G(α) = P P −1 . (26)
( aj )( ak ) − 1
j∈α k∈α
P5 P
Total energy-momentum conservation reads j=1 aj = 5j=1 a−1 j = 0. Rather remarkably,
2
on the support of these constraints and for v4 = 3v3 the first two terms in (25) sum to a
constant and can thus be cancelled by appropriately tuning the last term.

• Check this statement (in Mathematica of course).

Having cancelled three-particle production by setting v4 = 3v32 , we can now move on to


T2→4 where we get

1 1 1
2
M2!4 = + + +
m 2 2

1 1
+
2
+ + +
6

.
(27)

13
Again, we find that on the support of energy-momentum conservation, the first six terms
sum to a constant and can thus be cancelled by an appropriate choice of v6 in the last term.

• Check this statement (in Mathematica of course!).

It is possible to proceed in this way and find that one can always cancel T2→n−2 by appro-
priately fixing vn . After some tedious calculations, this leads to
 
2 v3 3 3v32 4 5v33 5 11v34 6 21v35 7
Linteraction = −m φ + φ + φ + φ + φ + ... (28)
3! 4! 5! 6! 7!

At this point we could try to guess the result. We assume that particle production can be
exactly cancelled and let us try to fix the coupling constants that guarantee it. The key
idea that allows us to fix the couplings uniquely is to use the multi-Regge limit introduced
in the first lecture, see figure (4). Because of (20), the only surviving tree-level Feynman
graphs are one-dimensional chains with all particles ordered. Figure 5 shows an example of a
surviving Feynman diagram. For example, the 2 → 4 amplitude (27) immediately simplifies
to −T2→4 /m2 = −v34 + 2v32 v4 + v32 v4 − v42 + 2v3 v5 + 0 + v6 .
It is now easy to find the general Lagrangian by induction. We consider T2→n assuming
T2→3 , . . . , T2→n−1 were already tuned to vanish by fixing the vertices up to vn+1 . The am-
plitude T2→n in the multi-Regge limit is given by a sum of one-dimensional ordered chains.
Particle 1 must therefore be at an end-point of such chains and can interact through a vertex
of any valency, as illustrated in Figure 6. The only surviving graphs are those where the
vertex is an n-, (n + 1)- or (n + 2)-particle vertex since those are respectively dressed by 4, 3
and 2 total particle amplitudes which are the only non-zero amplitudes (since 5, 6, . . . , n + 1
were already constrained to vanish, by assumption). We thus find

0 0 0
3 3 . . .(n − 2)
...

...
...

2 + 2 + ... + 2
1 1 1

n (n + 1)
3 . . . (n − 1) 3 ... n 3
...
(n + 1)
+ 2 (n + 1) + 2 + 2 (n + 2)
1 1 1
(n + 2) (n + 2)

Figure 6: We organize all ordered diagrams so that particle 1 is always on the left. The sum of
diagrams where particle 1 is attached to a k-point vertex with k < n evaluates to zero since it is
attached to a total amplitude with n − k + 4 > 4 external particles. As a result, only the three
contributions in the last row survive.

14
1 
− T2→n = vn+2 + (−vn+1 v3 ) + vn v32 − v4 (29)
m2 |{z} | {z } | {z }
1 1 1 + 1

Recalling that v4 = 3v32 and requiring this amplitude to vanish, we obtain the desired recur-
sion relation which one can readily solve.

• Show that
2 + (−2)n n−2
0 = vn+2 − vn+1 v3 − 2vn v32 ⇒ vn = λ , (30)
6
and show that this leads to the famous Bullough-Dodd model,
1 m2  λφ 
LBD 2
= (∂φ) − 2 2e + e−2λφ − 3 . (31)
2 6λ
We see that this is the only theory with a single massive scalar particle, a cubic coupling, a
perturbative expansion with no derivative couplings and no particle production at tree-level.
It is a pleasure to check that the Taylor expansion of this potential does match with the
painfully obtained data in (28).
We can also study Z2 -symmetric scalar theories, i.e. those where all odd-point interaction
vertices vanish. In this case, we can set v3 = 0 in (29) to obtain the simpler Z2 recursion
relation vn+2 − vn v4 = 0.

• Show that this leads to vn = β n−2 for n even so that the potential resums to the
sinh-Gordon theory2
1 m2
LsG = (∂φ)2 − 2 [cosh(βφ) − 1] . (32)
2 β

• We constructed theories with T2→4 = 0 but we can not have T3→3 = 0 obviously! Aren’t
they related by crossing though? What is going on? Hint: See Patrick Dorey’s ”Exact
S-matrices” review.
• Compute the 2 → 2 S-matrix in sine-Gordon for for small coupling. How does it
compare with the S-matrix in the previous section?
• If you feel brave, compute the S-matrix to one loop and note that it beautifully agrees
with the sub-leading expansion of the S-matrix of the previous section.

Given the trivial weak coupling expansion of (22) there should be a trivial way of computing
this S-matrix to all loops. Perhaps one should try to bring all the recent amplitude technology
to this problem and see if there is indeed some missing overlooked simplifications one could
exploit.
2
Or sine-Gordon if we take β to be purely imaginary.

15
2.3 Bullough-Dodd and Quantum No-Particle Production
• For Bullough-Dodd

sinh(θ) + i sin(α)
S(θ) = f2/3 (θ)f−2B/3 (θ)f(B−1)2/3 (θ) , fα ≡ (33)
sinh(θ) − i sin(α)

Each factor fα in (33) is called a CDD factor. Recall that the sine-Gordon breather
S-matrix was given by a single such factor. What do you think perturbation theory
corresponds to in terms of B? Why?

• Expand the S-matrix in the regime of the previous point and compare with 2 → 2
perturbation theory with the Lagrangian (31). You will find

Bπ Bπ Bπ
T (s) = + + + 3Bπ (34)
s − 1 (4 − s) − 1 0 − 1

which we like very much! We identify Bπ = c23 so that the last term is nicely c4 = 3c23 .

• Show that
S(θ + iπ/3)S(θ − iπ/3) = S(θ) (35)

This is the so-called fusion condition. It comes from the cubic property of the theory: The
bound-state in the 2 → 2 S-matrix is the fundamental particle itself. This condition can be
thought in two ways:

1. As the condition that the scattering of a particle with a boundstate being the same
after or before the bound-state forms since in an integrable theory we can shift wave
packets at will, sere figure 7.

2. As the condition that the Landau diagrams in 2 → 3 cancel which is required to have
such a vanishing amplitude. The figure would be the same with A − B = 0 instead of
A + B = 0.

Both are equivalent of course. We could discuss YB and regular vs singular parts of 3 → 3
and 2 → 4 here (see next section). That explains the minus sign also in A − B cancelation
in fusion.
If the bound-state were not the fundamental particle then the right hand side of (35)
would be the S-matrix of the bound-state with the fundamental particle. We can also see
that nicely from the 3 → 3 wave function. To see that we would:

• Write down the two-particle wave function with in-coming and out-going waves and
explain why a pole in the S-matrix is a bound-state if in the proper place in the complex
plane (upper vs lower HP)

• From a factorized three-particle wave function, with its six terms, explain why the
bound-state S-matrix is given by the generalization of (35)

16
Figure 7: Suppose we take two (to be) constituents of a bound-state and throw them very slowly
at each other so that they travel (almost) parallel to each other in space-time until they are close
enough to feel each other and thus form the bound-state. Now suppose we want to scatter a
fundamental particle with this bound-state as indicated on the left in this figure. This is how we
would compute the left hand side of (51). In an integrable theory we can shift at will the position
of the wave packet of this fundamental particle. So we can shift it far into the past such that it
scatters instead with the constituents of the bound-state well before they were bound together as
represented on the right. This leads to the right hand side of (51).

2.4 3 → 3 in 2D
We discuss some results from Iagolnitzer’s short paper https://journals.aps.org/prd/
pdf/10.1103/PhysRevD.18.1275. We would like to understand what replaces his equation
(4) in a general 2D QFT where particles have indices. It becomes3 is (in the vicinity of
(P1 , P2 , P3 ) = (P6 , P5 , P4 )):
" ! #
Ŝ(~
p ,
1 2p
~ )Ŝ(~
p ,
1 3p
~ )Ŝ(~
p ,
2 3p
~ ) Ŝ(~
p ,
2 3p
~ )Ŝ(~
p ,
1 3p
~ )Ŝ(~
p ,
1 2p
~ )
Ŝ3→3 = δ (2) (Pin − Pout ) − + r̂3,3
p~2 − p~5 − i0 p~2 − p~5 + i0
(36)
where r̂ is regular. If the product of the S-matrices is the same we can factor it out and use
1 1
− = 2iπ δ(~q)
~q − i0 ~q + i0
to obtain the missing delta function. (In the guessed formula there should be a simple
Jacobian missing (at least) since we want this new delta function to combine with the overall
energy-momenta conservation delta function to yield the I3,3 in Iagolnitzer’s equation (4).)
As a check consider convoluting the above S-matrix with wave packets
Z Z Z
−τ (p1 −P1 )2 −τ (p3 −P3 )2 2
dp~1 e , dp~3 e and dp~2 e−τ (p2 −P2 ) −iτ p~2 β (37)

3
This guess is inspired in the discussion following figure 5 in Iagolnitzer’s paper. With indices the diagrams
in the first and second line there do not cancel. Also, his equation (15) would not longer hold (being replaced
by a specific order for each D(i) ); probably the discussion after (15) could also be redone to yield this guess.
Finally, it is more or less derived in the book by the same author, see equations 1.99 and 1.100 therein.

17
were τ is a large positive number. Assume furthermore that p1 is moving to the right; p3
to the left and p2 is almost quiet. (So that the picture in space-time is the one we always
draw.)
If β = 0 the three wave packets will meet at the same point (the origin of space-time).
If β > 0 we shift particle p2 to the left hence particles 12 scatter first while if β < 0 we shift
the particle p2 to the right hence particles 23 scatter first.
Importantly, when integrating over p~2 we can move the contour up and down. We want
to move it down when β is positive so that e−iτ β p~2  1 and similarly we want to move it
up when β is negative. Now, if the S-matrix is regular – i.e. if it has no poles or cuts close
to the real axis – we will then get an exponential suppression from the exponential in the
wave packet. In particular we see that r̂3,3 which is regular by assumption will lead to an
exponentially small contribution in this limit both for positive or negative β. This is the
statement of macro-causality. On the other hand, the first two terms have poles in the real
axis ±i0’s. So the first term in (36) will contribute when we only when we need to move
the contour upwards (i.e. for β < 0) and will force us in that case to set p~2 = p~5 plus small
correction from poles further from the real axis. Similarly, the second one will contribute
when we only when we need to move the contour downwards (i.e. for β > 0) and will also
force us to set p~2 = p~5 . Furthermore, when p~2 = p~5 automatically we get p~1 = p~6 and p~3 = p~4
from the overall delta function hence landing on the usual factorization. Note also that the
fact that the two terms in (36) come with a relative sign is perfect since when we close the
contour up or down we pick a relative minus sign so in both cases we will get a product SSS.
Finally, note that in either case we indeed pick up an order of S-matrices compatible with
what we expect from the preceding paragraph. All seems consistent.

2.5 Ising With Magnetic Field


Pure magnetic field deformation of the critical Ising model leads to a beautiful massive theory
with the S-matrix of the lightest fundamental excitation given by

SMagnetic Ising (θ) = f2/3 f2/5 f1/15 (38)

It has eight particles with masses m1 , . . . , m8 . How would you go about finding their masses
from this fundamental S-matrix?

2.6 Advanced Sine-Gordon Topics


Sine-Gordon has breathers and Kinks.

Breathers

• For example, we have,

sinh(θ) + i sin(α)
S11 (θ) = , S12 (θ) = S11 (θ + iα/2)S11 (θ − iα/2) (39)
sinh(θ) − i sin(α)

18
Write again S12 as a simple product of CDD factors.

• What expression would you write for S22 ?

• What is the mass of the second breather measured in units of the lightest breather?

• Where are the poles of S12 and of S22 ? How do you interpret them?

Kinks

The breathers themselves are bound-states of kinks ψ (and anti-kinks ψ̄) which interpolate
between φ ∼ φ + period vacua.

• For example, we expect

S11 (θ) = S1ψ (θ + iβ)S1ψ̄ (θ − iβ) (40)

Explain why. What picture would you draw for this process?

• We also expect
S12 (θ) = S1ψ (θ + iβ 0 )S1ψ̄ (θ − iβ 0 ) (41)
What is the relation between β 0 and β? Hint: You should use the second breather
mass computed above.

• We also expect
S1ψ (θ) = S11 (θ + iµ)S1ψ̄ (θ − iη 0 ) (42)
Explain why and draw a corresponding figure again.

• If we assume the S1ψ (θ) to be given by a single CDD factor then the above equations
restrict it up to a few discrete options. In the end we conclude that

sinh(θ) + i cos(α)
S1ψ = S1ψ̄ = (43)
sinh(θ) − i cos(α)

Check that this indeed obeys all equations above and fix all constants in (40), (41) and
(42). Use it to relate the mass of the kinks to the mass of the breather bound-states.

• Finally, we expect the kinks themselves to fuse into S1ψ (θ + iβ) so that in the end the
CD
fundamental S-matrices would be SAB where A, B, C, D = ψ, ψ̄ for a genuine matrix.

• How would crossing symmetry constrain all these elements?

• Google for the kink S-matrix and learn that it satisfies YB.

19
3 Lecture 3. Bootstrap. Complex Analysis.
Here we forget Lagrangians. What is the theory with a single bound-state of mass mb in
the lightest particle S-matrix and with biggest residue at the corresponding pole, i.e. with
largest coupling?
Physically, it makes sense that such bound exist. A stronger coupling would lead to more
bound-states or lower mass for mb . Mathematically also, see first subsection below. To warm
up for our S-matrix bootstrap explorations, we will do some complex anaysis stretching first.

3.1 Complex Analysis


Maximum Modulus Principle

• Consider an holomorphic function f (z) with no singularities inside the unit disk. As-
sume it is bounded on the unit disk as |f (z)| ≤ 1 at |z| = 1. Show that the function is
bounded as |f (z)| ≤ 1 for any point inside the unit disk. This is the famous maximum
modulus principle

• What is the maximum value such function could take at the origin? What function
yields that maximum value?

• Consider a function with a single simple pole at the origin and no other singularities
inside the unit disk. Assume, again, that it is bounded on the unit disk as |f (z)| ≤ 1
at |z| = 1. What is the maximum value for the residue of the single pole and which
function saturates that bound?

• Consider now an even function, real for z real, bounded by 1 again at the boundary of
the unit disk and with two symmetric poles at z = ±z∗ ∈ [0, 1]. What it the maximum
value of its residue at those simple poles?
Spoiler: This will be directly relevant for S-matrix bootstrap once we map the full cut
s-plane to the unit disk; the two poles will them be the s- and t- channel poles and the
even symmetry of the function will be nothing but crossing symmetry.

• How do the first two points change if we replace the unit disk by any other connected
region?

Dispersion Relations

• Consider a function with a pole at z = 0, a cut going from z = z∗ > 0 to infinity and
large z behavior given by f (z) ∝ 1/z at large z. No other singularities. Show that
Z ∞
g ρ(w)
f (z) = + (44)
z z∗ z − w

Hint: Start with a contour integral and blow the contour.

20
• Relate ρ to the discontinuity of the function f along the cut, with all factors of π
carefully worked out.
• Consider a function with a pole at z = 0, a cut going from z = z∗ > 0 to infinity
and large z decay given by f (z) ∝ 1/log(z) or faster at large z. No other singularities
again. Note that this includes the previous case. Show that
z g Z ∞ ρ(w) 
0
f (z) = (z + z0 ) + (45)
z z ∗ (z − w)(w + z0 )

where z0 is an arbitrary point (not on the cut say). Is the ρ here the same as the ρ
in (44)? Relations (44) and (45) are so called dispersion relations. The second one
has one subtraction; the first one has no subtractions. When would you need more
subtractions?
• Consider next a function with two cuts, a few poles and bounded at infinity. No other
singularities. What dispersion relation would you write in this case?
Spoiler: Obviously, very relevant for S-matrix physics! The two cuts will be the s- and
t- channel cuts later.

Schwarz-Pick Lemas

Schwarz-Pick Lemas are powerful generalizations of the maximum modulus principle. Con-
sider a function in the upper half place which is bounded as |f (z)| ≤ 1 for real z and also at
infinity. Suppose furthermore that the function has no singularities in the upper half plane.
Then by the maximum modulus principle we have |f (z)| ≤ 1 all over the upper half place.

• Show that
f (z) − f (w) . z − w
g(z) ≡ , (46)
1 − f (z)f (w) z − w
is bounded as |g(z)| ≤ 1 everywhere in the upper half plane as well for any parameter
w in the upper half plane. This is the so called Schwarz-Pick lemma.
Note that here w is a parameter so g(z) is a family of functions parametrized by w; it is
on the other hand an holormorphic function of z only; w appears non-holomorphically
which is totally fine of course.
• Show that
g(z) − g(w0 ) . z − w0
h(z) ≡ 0 , (47)
1 − g(z)g(w0 ) z − w
is bounded as |g(z)| ≤ 1 everywhere in the upper half plane as well for any parameter
w0 in the upper half plane. This is the so called second Schwarz-Pick lemma.
2n−1
P
• Consider a function f (z) = ei n γn z which is manifestly a phase for real z. Assume
it has no singularities in the upper half plane as in the previous points. Further set
γ1 = 1/4 since we can always re-scale z. Then show that the space of allowed expansion
parameters γ3 , γ5 , . . . can be powerfully constrained by the maximum modulus principle
as well as its Schwarz-Pick lemmas.

21
<latexit sha1_base64="ML55oyr37N+DTbkwO41v8m9Dr08=">AAACB3icbVDLTgJBEJz1ifhCOXqZSEw8kV010SPRi0dM5JEAIb1DgxNmdjczvRpC+AC/wquevBmvfoYH/8VZ5KBgnSpV3enqChMlLfn+p7e0vLK6tp7byG9ube/sFvb26zZOjcCaiFVsmiFYVDLCGklS2EwMgg4VNsLhVeY37tFYGUe3NEqwo2EQyb4UQE7qFortB9lDkqqHvD0AraF72i2U/LI/BV8kwYyU2AzVbuGr3YtFqjEiocDaVuAn1BmDISkUTvLt1GICYggDbDkagUbbGU/DT/hRaoFinqDhUvGpiL83xqCtHenQTWqgOzvvZeJ/Xiul/kVnLKMkJYxEdsi9idNDVhjpWkHekwaJIEuOXEZcgAEiNJKDEE5MXU1510cw//0iqZ+UA8dvzkqVy1kzOXbADtkxC9g5q7BrVmU1JtiIPbFn9uI9eq/em/f+M7rkzXaK7A+8j29KXJkW</latexit>
e3

<latexit sha1_base64="2EK+LaKywrE52gBPNbJX0R3P9dM=">AAACB3icbVDLTgJBEJz1ifhCOXqZSEw8kV1jgkeiF4+YyCMBQnqHBifM7G5mejWE8AF+hVc9eTNe/QwP/ovDykHBOlWqutPVFSZKWvL9T29ldW19YzO3ld/e2d3bLxwcNmycGoF1EavYtEKwqGSEdZKksJUYBB0qbIajq5nfvEdjZRzd0jjBroZhJAdSADmpVyh2HmQfSao+8s4QtIZepVco+WU/A18mwZyU2By1XuGr049FqjEiocDaduAn1J2AISkUTvOd1GICYgRDbDsagUbbnWThp/wktUAxT9BwqXgm4u+NCWhrxzp0kxrozi56M/E/r53S4KI7kVGSEkZidsi9idkhK4x0rSDvS4NEMEuOXEZcgAEiNJKDEE5MXU1510ew+P0yaZyVA8dvzkvVy3kzOXbEjtkpC1iFVdk1q7E6E2zMntgze/EevVfvzXv/GV3x5jtF9gfexzdQmJka</latexit>
e7

<latexit sha1_base64="YQ5j/3E8VlF03U1E2ZIKjeEV9R4=">AAACB3icbVDLTgJBEJz1ifhCOXqZSEw8kV2j0SPRi0dM5JEAIb1DgxNmdjczvRpC+AC/wquevBmvfoYH/8VZ5KBgnSpV3enqChMlLfn+p7e0vLK6tp7byG9ube/sFvb26zZOjcCaiFVsmiFYVDLCGklS2EwMgg4VNsLhVeY37tFYGUe3NEqwo2EQyb4UQE7qFortB9lDkqqHvD0AraF71i2U/LI/BV8kwYyU2AzVbuGr3YtFqjEiocDaVuAn1BmDISkUTvLt1GICYggDbDkagUbbGU/DT/hRaoFinqDhUvGpiL83xqCtHenQTWqgOzvvZeJ/Xiul/kVnLKMkJYxEdsi9idNDVhjpWkHekwaJIEuOXEZcgAEiNJKDEE5MXU1510cw//0iqZ+UA8dvTkuVy1kzOXbADtkxC9g5q7BrVmU1JtiIPbFn9uI9eq/em/f+M7rkzXaK7A+8j29NepkY</latexit>
e5

1
Figure 8: Allowed region in the {γ̃3 , γ̃5 , γ̃7 } with γ̃n =γn + (−1)(n+1)/2 n23n−1 . If we interpret f (z) as
an S-matrix we will see that these parameters constrain the space of low energy Wilson coefficients.

– For example, you should find that


1
γ3 > − (48)
768
– More generally, for the first three non-trivial constants γ3 , γ5 , γ7 you should find
an allowed space with the shape of figure (8).

Two bonus problems

Easy Consider a regular function f (z) which is bounded on the boundary of the unit disk
as |f (z)| ≤ 1 in the arc where z = eiφ and −α < φ < α and |f (z)| ≤ B > 1 in the
complement arc. Show that the maximum modulus of f (0) is given by B α/2π . Solve
this problem analytically and numerically.

Hard Consider a set of regular functions fn (z) bounded on the left side of the unit disk (that
is, for |z| = 1 with re(z) > 0) and with reflection symmetry fn (−z) = Cnm fm (z) where
the constant matrix C obeys C 2 = I as required by consistency of z → −z → z. What
is the maximum modulus of f1 (0)? How would you go about solving this problem given
a constant matrix C.
Spoiler: I do not know how to solve this problem analytically. If you do let me know!
This problem would be very relevant for the S-matrix bootstrap where we have various
type of particles with flavour symmetry for example. In that context C would be the
crossing matrix relating the various scattering channels.

22
Figure 9: Maximum cubic coupling g1max between the two external particles of mass m and the
exchanged particle of mass m1 . Here we consider the simplest possible spectrum where a single
particle of mass m1 shows up in the elastic S-matrix element describing the scattering process of
two mass m particles. The red dots are the numerical results. The blue (white) region corresponds
to allowed (excluded) QFT’s for this simple spectrum.

3.2 Sine-Gordon and other friends


• Answer the question we started this section with, namely: What is the theory with a
single bound-state of mass mb in the lightest particle S-matrix and with biggest residue
at the corresponding pole, i.e. with largest coupling?
• Plot the corresponding residue as a function of the coupling. Normalize it as to be
the residue of T . Then you should get a plot as in figure 9. This is our first S-matrix
bootstrap bound. Note that s-channel poles have positive residue and t-channel have
negative residue. That is quite important!

• What theory lies on boundary for mb > 2?

• What theory lies on boundary for mb < 2? Hint: We don’t know. Not a physical
theory probably?
• We can now generalize for more masses. For example, for three masses with m1 = m,
the same as the external particle, and m3 > m2 > m1 , we get figure 10. Why all these
chambers with so many discontinuities? Hint: Residues must have good signs...
• What question could have the Bullough-Dodd S-matrix (33) as its solution?

3.3 Magnetic Ising Field Theory Parentheses


Focus on one of the chambers in figure 10, the one with
√ √
2m1 < m2 < 3m1 < m3 < 2m1 . (49)

23
ore generally, for m1 = 1 and two other masses m2 , m3 < 2 we get the optimal bound
ure 6. We have
8
>
> [↵1 ] [↵2 ] region A
>
>
>
> [↵ 1 ] [↵ 2 ] [↵ 3 ] region B
>
> m1 = m
< [↵1 ] [↵3 ] region C
Smax m1 residue = [↵1 ] [↵3 ] region D (25)
>
>
>
> [↵1 ] [↵2 ] [↵3 ] region E
>
>
>
> [↵ ] [↵ ] region F
: 1 3
[↵1 ] region G
max 2
log(gthe
mi = 2 cos(↵i /2) and )
1 short-hand notation

sinh(✓) + i sin(↵)
[↵] ⌘ ,
sinh(✓) i sin(↵)

CDD factor. The analysis is in the CDDs m2m3 plane.nb notebook. DSome interesting
ns of the general three dimensional plotAcan also be found there [to add]. E G
F

erences B

Zamolodchikov, A. B., Zamolodchikov, A. B. (1979). Factorized S-matrices in two di-


(m2 /m)2 (m3 /m)2
mensions as the exact solutions of certain relativistic quantum field theory models.
Annals of physics, 120(2), 253-291.
C
http://www.sns.ias.edu/~malda/Zamolodchikov.pdf

Dorey, Patrick. ”Exact S-matrices.” Conformal field theories and integrable models.
Springer Berlin Heidelberg, 1997. 85-125. http://arxiv.org/abs/hep-th/9810026

Basso, Benjamin, Amit Sever, and Pedro Vieira. ”Space-time S-matrix and flux tube
S-matrix II. Extracting
Figure 10:and matching
Maximal data.”g max
coupling Journal of mHigh
(m2 /m, Energy Physics 2014.1
3 /m) for the spectrum m1 = m (i.e. a cubic coupling
1
2014): 1-72. http://arxiv.org/abs/1306.2058
m + m → m) and generic m < m2 < m3 < 2m.
Coleman, Sidney, and H. J. Thun. ”On the prosaic origin of the double poles in the
ine-GordonS-matrix.” Communications
In this region in Mathematical
the S-matrix which Physics
maximizes g1 is a simple61.1 (1978):
product 31-39.
of three CDD factors,
http://users.physik.fu-berlin.de/~kamecke/ps/coleman-thun.pdf
sinh(θ) + i sin(2π/3) sinh(θ) + i sin(α2 ) sinh(θ) + i sin(α3 )
S(θ) = × × , mj = 2 cos(αj ) . (50)
sinh(θ)
Smirnov, F. A. ”Reductions − isine-Gordon
of the model− as
sin(2π/3) sinh(θ) i sin(α2) sinh(θ) − of
a perturbation i sin(α3)
minimal
models of conformal field theory.” Nuclear Physics B 337.1 (1990): 156-180. http:
We will now argue that in the region (49) of parameter space our bound should not be the strongest possible
//users.physik.fu-berlin.de/~kamecke/ps/smirnov2.pdf
bound except at a single isolated point which we will identify with a well known and very interesting field
theory.4 We will do this by observing some simple pathologies with (50) which are resolved once α2 and α3
S. Dubovsky, V.take
Gorbenko and M.values
some particular Mirbabayi,
which we“Natural Tuning: Towards A Proof of
identify below.
Concept,” JHEP 1309 (2013) 045 doi:10.1007/JHEP09(2013)045 [arXiv:1305.6939 [hep-
To proceed we need to make three natural assumptions about a putative theory living in the boundary
h]]. of our bounds for a fixed mass spectrum M:

A1 The theory is integrable.5


4
The reader fond of section titles probably guessed which one.
5
This is of course very natural since the S-matrices we found saturate unitarity and thus admit no particle
12 production is of course a necessary condition for integrability. In most cases
production. Absence of particle
it is also a sufficient condition.

24
m1 = m01 = 1 m3 = m03 m04

m2 m02

m2 = m02

log(g1max )2
m02 m2

1.0 1.5 2.0 2.5

poles of S = S11
poles of S12

m3 /m

m2 /m S(✓ + i⇡/3)S(✓ i⇡/3) = S(✓)

Figure 11: Blow up of region B from figure ??. The thick black line is where the cubic fusion
property (52) holds (i.e. assumption (1) in the discussion). In the upper right corner we plot the
s-channel poles of S12 versus those of S. We see that, following the thick black line, only at the
blue dot does S12 have poles at the same locations as S indicating that assumption (3) also holds.

A2 The exchanged particle with mass m1 = m is really the same as the external particle, i.e. it is not
just another particle in the theory with the same mass as the external particle.
A3 There are no other stable particles below the two particle threshold 2m1 other than those in M.

In an integrable theory we can construct bound-state S-matrix elements from the fundamental S-matrix
by fusion as discussed above. If the stable particle shows up as a pole at θ = iαj in S(θ) then the S-matrix
of this bound-state with the fundamental particle of mass m can be built by scattering both its constituents,
Sjth bs, fund (θ) = S(θ + iαj /2)S(θ − iαj /2) . (51)

With the fusion property (51) following from assumption A1 we will now show that powerful constraints
on the spectrum follow from assumptions A2 and A3.
If a theory has a cubic coupling and m1 = m shows up as a pole in the S-matrix then it can itself be
thought of as a bound-state. That is, under the assumptions (1) and (2) above we conclude that we must
have
S(θ) = S(θ + iπ/3)S(θ − iπ/3) . (52)

25
This is an important self-consistency constraint. We can now plug the solution (50) in this relation. We
observe that it is generically not satisfied. However, there is a line α3 (α2 ) or equivalently m3 (m2 ) where it
holds. This is the thick black line in figure 11. Away from this black line we can already conclude that our
bound is either not the optimal bound or some of the assumptions A1 or A2 (or both) should not hold.
Sticking to the black line and continuing with assumption (3) we can do even better. We can now
construct the S-matrix element S12 (θ) = S(θ + iα2 /2)S(θ − iα2 /2) for the scattering m1 + m2 → m1 + m2
involving the lightest and the next-to-lightest particles. We can then look at the poles of this S-matrix which
will correspond to asymptotic particles of the theory. There is a point in the black line, marked with the blue
dot in figure 11 where these poles correspond perfectly to the spectrum M = {m1 (= m), m2 , m3 }. Namely
we find precisely three s-channel poles at s = m21 , m22 , m23 < (2m1 )2 which are the very same locations in the
fundamental S-matrix S(θ). However, as we move away from this blue point something bad happens. We
see that the poles at s = m21 and s = m23 are as expected however the pole at m22 shifts to a nearby position
m02
2 . This would indicate the presence of a new particle not in M with a mass close to that of m2 . This
violates assumption A3.
Ultimately, only the blue dot in figure 11 which is located at

m2 = 2 cos(π/5)m1 , m3 = 2 cos(π/3)m1 , (53)

survives! We conclude that under the assumptions A1–A3 the maximal coupling in region B of figure 10
(which corresponds to masses satisfying (49)) should be lower than the one we found everywhere except
perhaps at the blue point.6
What about this blue dot? Is there a special integrable theory with these masses and an S-matrix given
by (50)? Yes, it is the Scaling Ising model field theory with magnetic field! This is a very interesting strongly
coupled integrable theory with E8 symmetry which describes the massive flow away from the critical Ising
model when perturbed by magnetic field (holding the temperature fixed at its critical value).7 Thus the
CDD solution provides a sharp (i.e. as strong as possible) upper bound on g1 for this value of the masses.
In what follows we shall refer to the blue dot in figure 11 as the magnetic point.
Away from the magnetic point, the bound in figure 11 is not optimal. The obvious question is then how
to improve it? Multiple correlators, more amplitudes,... These would affect unitarity and hence improve the
bounds.

6
Note that we can not exclude having other integrable theories living in the black line provided we accept
more stable particles below threshold showing up in other S-matrix elements. We could also drop assumption
A2 and conceive integrable theories where m1 is not the same particle as the external one (despite having
the same mass). If we keep assumption A3, the conclusion leading to the blue dot as a special isolated theory
still holds.
7
This is perhaps not that surprising. After all, many of the conditions we just imposed are simple recast
of standard integrable bootstrap logic as used, for instance, in [?].

26
4 Lecture 4. More Advanced Bootstrap, Flux tubes,
etc.

4.1 Higher Dimensional Parentheses III


We found analytically that a single CDD solution S-matrix is the solution with the largest possible coupling
and a single bound state. We did it analytically. In figure 12 we have a few numerical implementations using
the unit disk maps we discussed above.
In higher dimensions we can not solve things analytically but these numerical implementations can be
generalized nicely. The main complication is how to preserve both crossing and unitarity. We can make
the first manifest and impose the second one by decomposing in partial waves. In detail (here goes more
shameless copy/paste):
Consider again the elastic scattering process of two identical real scalar particles of mass m. In our
conventions the S-matrix element is

hp3 , p4 |S|p1 , p2 i = 1 + i(2π)d+1 δ (d+1) (p1 + p2 − p3 − p4 )M (s, t, u) (54)

with normalization such that


 
1 = (2π)2d 4Ep1 Ep2 δ (d) (p1 − p3 )δ (d) (p2 − p4 ) + (3 ↔ 4) (55)
p
where Ep = m2 + p2 . The Mandelstam invariants are given by

s = (p1 + p2 )2 t = (p1 − p3 )2 u = (p1 − p4 )2 (56)

which of course obey s + t + u = 4m2 , and we henceforth work in units such that m2 = 1. We often write
M (s, t) ≡ M (s, t, 4 − s − t). In the channel under consideration s is the squared center-of-mass energy and
the scattering angle is given by
2t 2u
x = cos(θ) = 1 + = −1 − (57)
s−4 s−4

Physical values of the Mandelstam invariants are therefore 4 ≤ s and 4 − s ≤ t ≤ 0. We can project onto
channels with definite angular momentum by introducing the partial amplitudes:

d−2 Z1
(s − 4) 2 d−3 (d)
S` (s) = 1 + i √ dx (1 − x2 ) 2 P` (x) M (s, t)| 1 (58)
s t→ 2 (s−4)(x−1)
−1

(d)
where P` (x) is proportional8 to the Gegenbauer polynomials. In our conventions,

(3) 1 (2) 1
P` (x) = P` (x) , P` (x) = cos(`θ) , (59)
32π 8π
with P` (x) the usual Legendre polynomials, normalized such that P` (1) = 1. We note that S` (s) = 1 for odd
` because Bose symmetry implies invariance under the reflection θ → π − θ.
Although the S-matrix element (54) has all kind of distributional properties, the amplitude M (s, t, u) is
a regular function (see e.g. [?, section 4.3]). We will assume that M (s, t, u) obeys three further constraints:
8
In general spacetime dimension, we have

(d) l! Γ( d−2
2 ) (d−2)/2
P` (x) = d C` (x) .
4(4π) 2 Γ(d + l − 2)

27
Figure 12: To find optimal S-matrices we parametrizes regular functions inside the unit disk (plus
a few poles) and simply find the maximum within such variational ansatz. The last plot is a
comparison with the analytical solution (in this case a single CDD); there are two curves there so
clearly it is working super well.

• Crossing Symmetry: M (s, t, u) is completely symmetric in its arguments. The symmetry u ↔ t


follows from the aforementioned Bose symmetry, but the other generator of the crossing symmetry
group can only be found from a more sophisticated analysis and requires the LSZ prescription.
• Analyticity: M (s, t, 4 − s − t) is analytic for arbitrary complex s and t, except for potential bound-
state poles at s = m2b with 0 < m2b < 4, a cut along the real axis starting at s = 4, and the images
of these singularities under the crossing symmetry transformations. It further obeys the usual reality
condition M (s∗ , t∗ 4−s∗ −t∗ ) = M ∗ (s, t, 4−s−t). We note that the analyticity assumption is actually
rather optimistic, since this ‘maximal’ analyticity has not been proven from axiomatic field theory.9
9
Certain analyticity properties are known to be valid very generally, derived either to all orders in per-
turbation theory or from axiomatic field theory; the latter case sometimes requires the Wightman axioms
and other times merely requires the validity of the LSZ prescription and causality. Typically one can prove
two-variable analyticity for all s (modulo the known poles and cuts) but only for some finite range of values
of t or of x which in particular includes the physical values. A standard result is that the proven analyticity
is sufficient to analytically continue the amplitude from the s-channel to the t or u channels, establishing

28
• Unitarity: From S † S = 1 we find that the unitarity constraint for elastic scattering takes the form

|S` (s)| ≤ 1 (60)

for all s ≥ 4 and ` ∈ {0, 2, 4, . . .}. Generically no other channels are available for a finite window of
values of s, starting at 4 and ending at a higher threshold (like s = 9 for three-particle scattering). In
such a window the above inequality should in fact be saturated. In this work we will not impose such
saturation, but our numerics in principle allows for it.

Next we explore the consequences of our analyticity assumption in some detail. As a toy model we can start
with a single-variable function f (z) which is analytic in a simple domain D ⊂ C. If we define ρ : D → ∆ as
a biholomorphic map between D and the unit disk ∆ = {ρ ∈ C : |ρ| < 1}, then any such f (z) has a Taylor
series expansion of the form
X∞
f (z) = cn ρ(z)n (61)
n=0

which converges as long as |ρ(z)| < 1. Our multi-variable problem is unfortunately not so easy, since for
M (s, t) the moving cuts imply that the domain of analyticity in one variable, say s, depends on the other
variable t. We will remedy this as follows. First we relax the constraint s + t + u = 4 and consider three-
variable functions M (s, t, u). Then we transform the variables (s, t, u) → (ρs , ρt , ρu ) using the map
√ √
4 − s0 − 4 − s s0 (1 − ρs )2 + 16ρs
s 7→ ρs = √ √ , s= . (62)
4 − s0 + 4 − s (1 + ρs )2

In this case, it is convenient to choose s0 = 43 so that ρs = ρt = ρu = 0 corresponds to the crossing symmetric


point s = t = u = 34 . Now, since the transformation ρs maps the s-plane minus the right cut starting at
s = 4 to the unit disk, we see that in the ρ variables all the cuts lie outside the polydisk ∆3 defined by
|ρs | < 1, |ρt | < 1 and |ρu | < 1. The only remaining singularities are then the poles and it is natural to write

g2 g2 g2 X
M (s, t, u) = − 2 − 2 − 2 + αabc ρas ρbt ρcu (63)
s − mb t − mb u − mb
a,b,c=0

where the triple ρ series converges inside ∆3 , and for definiteness we have put in the poles for a single
scalar bound state of mass mb . The demands of crossing symmetry are implemented by demanding that the
coefficients αabc are totally symmetric in their indices. When restricted to the surface defined by s+t+u = 4
the ansatz (63) obeys the analyticity and crossing symmetry constraints. It is perhaps more surprising that
the converse is also true: any function obeying the analyticity constraints on the surface s + t + u = 4 can
be extended to a function on ∆3 , analytic modulo the poles, and therefore can be written in the form (63).
This follows from a mathematical theorem known as Cartan’s theorem B, which is a statement about the
vanishing of higher cohomologies of coherent analytic sheaves on Stein manifolds – in the case at hand this
implies that there is no obstruction to an extension away from the surface s + t + u = 4.
The triple ρ expansion in equation (63) is the starting point for our numerical work. Our approach is to
restrict the expansion to a finite sum by imposing

a + b + c ≤ Nmax (64)

and then further restricting to the constraint surface s + t + u = 4 which is given by a polynomial equation

ρ2s ρ2t ρu + ρ2s ρ2u ρt + ρ2t ρ2u ρs + (lower degree terms) = 0 (65)

and which in practice allows us to eliminate many terms in (63). The remaining freedom in our ansatz
then consists of the finitely many remaining αabc together with the bound state parameters; since this is a
finite-dimensional space we can use a computer to numerically explore the space of scattering amplitudes. Of

crossing symmetry.

29
course we want to keep Nmax as large as possible. In fortunate cases the numerical results stabilize already
for feasible values of Nmax , while in other cases we can extrapolate.
It will be the job of the computer to impose the unitarity constraints, which are quadratic constraints
in the parameters g 2 and αabc . Rather than checking the infinity of constraints for all s and `, we impose
a cutoff and check that unitarity constraints are obeyed only for ` ≤ `max and along a grid of values for s.
Experimentally we observe that our results remain meaningful if `max is not much smaller than Nmax and if
the grid is sufficiently refined.
The paper from where this section was copy/pasted has nice plots.

4.2 Flux Tubes


Here we basically review our last flux tube paper with Aditya, Andrea, Joan and Joao as a
nice application of the complex analysis technology we developed in lecture 3.

4.3 General Comments


Mention open problems.

30

You might also like