Real and p-Adic Physics

154
Real and p-Adic Physics Brian Raymond Trundy A Dissertation Presented to the Faculty of Princeton University in Candidacy for the Degree of Doctor of Philosophy Recommended for Acceptance by The Department of Physics Adviser: Steven S. Gubser November 2021 Copyright by Brian Trundy, 2021. All rights reserved.

Transcript of Real and p-Adic Physics

Page 1: Real and p-Adic Physics

Real and p-Adic Physics

Brian Raymond Trundy

A Dissertation

Presented to the Faculty

of Princeton University

in Candidacy for the Degree

of Doctor of Philosophy

Recommended for Acceptance by

The Department of

Physics

Adviser: Steven S. Gubser

November 2021

© Copyright by Brian Trundy, 2021.

All rights reserved.

Page 2: Real and p-Adic Physics

Abstract

This thesis examines the consequences of non-Archimedean geometry in hologra-

phy and many-body physics. In this framework real space is replaced by a non-

Archimedean field which is almost always a p-adic field or an algebraic extension of a

p-adic field. The corresponding bulk geometry is replaced by a discrete tree. The re-

sulting theories are not only useful as toy models of real theories - they help elucidate

which parts of physics do and do not depend on the choice of underlying field.

Chapter 1 sets the stage for the rest of the thesis and gives a brief review of

non-Archimedean mathematics and physics.

Chapter 2 is based on [1] coauthored with Steven S. Gubser, Christian Jepsen,

and Ziming Ji. We exhibit a sparsely coupled classical statistical mechanical lattice

model that interpolates between real and p-adic geometry when varying a spectral

exponent. Holder continuity conditions in both real and p-adic space allow us to

quantify how smooth or ragged the two-point Green’s function is as a function of the

spectral exponent. This model was motivated by proposed cold atom experiments [2]

and serves as our bridge into the non-Archimedean world.

Chapter 3 is based on [3] coauthored with Steven S. Gubser, Christian Jepsen,

and Ziming Ji. We study melonic tensor models over real and non-Archimedean

fields in tandem. Much attention is paid to the combinatorial structure of potential

interaction terms and the perturbative expansion. The Schwinger-Dyson equation is

solved exactly in the p-adic case for a subset of these field theories.

Chapter 4 is based on [4] coauthored with Steven S. Gubser and Christian Jepsen.

We examine the bulk dual p-adic field theories whose Green’s functions are non-

trivial sign characters. These theories constitute the simplest known notion of spin in

p-adic physics. The construction is achieved by the introduction of an non-dynamical

U(1) gauge field on the discrete bulk geometry and the two point functions for dual

operators on the boundary are computed explicitly.

i

Page 3: Real and p-Adic Physics

Acknowledgments

I would like to express my gratitude to all my friends and mentors at Princeton. I

will always look back fondly at my time spent on this campus. Thank you, Herman

Verlinde for seeing this project through with me to its end. I would like to thank Silviu

Pufu for serving as a reader of this thesis and for his valuable counsel. Thank you,

Lyman Page and Simone Giombi for serving as examiners on my FPO committee.

Without the following collaborators, this thesis would not exist: Sarthak Parikh,

whose work effort is rivaled only by his kindness and intellect. Christian Jepsen, one

of the only true polymaths I’ve ever met. Ziming Ji, whose unrelenting mind plumbs

the depths of every problem. Amos Yarom, a master of molding the complex into

the straightforward. Matthew Heydeman, a mind of unmatched curiosity. Bogdon

Stoica, whose creativity and breadth knows no bounds. Ingmar Saberi, a towering

intellect. Matilde Marcolli, a seeker of the most profound truths. Gregory S. Bentsen

and Monika Schleier-Smith, whose invaluable experimental insight bridges the gap

between real life and the most abstract theory.

I would like to thank my wife Brittany Holom-Trundy and our two dogs Mason

and Elvis. Our home together makes life worth living. Thank you, Mom and Dad,

for your unending support and encouragement.

Lastly, thank you Steven Scott Gubser. Your loss hit me more deeply than I

thought it could. Yet my loss pales in comparison to that felt by the whole of Prince-

ton, your family, and the scientific community. You were remarkable in every aspect

of your life and I fear that I will never be able to repay you for your guidance and

generosity.

ii

Page 4: Real and p-Adic Physics

Contents

1 Introduction 1

1.1 The p-adic numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.2 The Bruhat-Tits Tree and its symmetry group . . . . . . . . . . . . . 7

1.3 Functions of a p-adic variable . . . . . . . . . . . . . . . . . . . . . . 9

1.4 Integration and additive characters . . . . . . . . . . . . . . . . . . . 10

1.5 Field extensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.6 Multiplicative characters and Gel’fand-Graev gamma functions . . . . 14

1.7 p-Adic field theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

1.8 p-Adic CFT and the p-adic AdS/CFT correspondence . . . . . . . . . 19

2 Continuum Limits of Sparsely Coupled Ising-like Models 24

2.1 The Statistical Mechanical Framework . . . . . . . . . . . . . . . . . 27

2.2 Nearest neighbor coupling . . . . . . . . . . . . . . . . . . . . . . . . 30

2.3 p-Adic coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

2.4 Power-law coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

2.5 Sparse coupling: From the Archimedean to the non-Archimedean . . 35

2.6 Continuity conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

2.7 2-adic field theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

2.8 Archimedean field theory . . . . . . . . . . . . . . . . . . . . . . . . . 42

2.9 Numerical evidence: 2-adic Approximation of sparse results . . . . . . 46

2.10 Numerical evidence: Holder bounds in momentum space . . . . . . . 47

iii

Page 5: Real and p-Adic Physics

2.11 Numerical evidence: Holder bounds in position space . . . . . . . . . 50

2.12 The liminal region −1/2 < s < 1/2 . . . . . . . . . . . . . . . . . . . 52

2.13 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

3 Higher Melonic Field Theories 60

3.1 Structure of higher melonic theories . . . . . . . . . . . . . . . . . . . 63

3.2 Symmetry groups of interaction vertices . . . . . . . . . . . . . . . . 68

3.3 Construction of interaction vertices . . . . . . . . . . . . . . . . . . . 73

3.4 One-factorizations and equivalent interaction terms . . . . . . . . . . 77

3.5 Finding the number of isomorphism classes of one-factorizations for

q = 12 using the orderly algorithm . . . . . . . . . . . . . . . . . . . 80

3.6 Isomorphism classes of ordered one-factorizations for q = 12 . . . . . 85

3.7 The two-point function and the Schwinger-Dyson equation . . . . . . 89

3.7.1 The IR solution . . . . . . . . . . . . . . . . . . . . . . . . . . 94

3.7.2 The zoo of theories . . . . . . . . . . . . . . . . . . . . . . . . 95

3.7.3 Full solution to the Schwinger-Dyson equation for direction-

dependent theories . . . . . . . . . . . . . . . . . . . . . . . . 97

3.8 The four-point function . . . . . . . . . . . . . . . . . . . . . . . . . . 98

3.8.1 Adelic product formula for the integral eigenvalues . . . . . . 104

3.9 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

4 Holographic Duals of Nontrivial Characters in p-adic AdS/CFT 107

4.1 Nearest neighbor actions . . . . . . . . . . . . . . . . . . . . . . . . . 111

4.1.1 Bosonic actions . . . . . . . . . . . . . . . . . . . . . . . . . . 111

4.1.2 Fermionic actions . . . . . . . . . . . . . . . . . . . . . . . . . 114

4.2 The background geometries . . . . . . . . . . . . . . . . . . . . . . . 115

4.3 Bulk-to-boundary propagators . . . . . . . . . . . . . . . . . . . . . . 118

4.3.1 Scalars on Tp . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

iv

Page 6: Real and p-Adic Physics

4.3.2 Scalars and fermions on the line graph . . . . . . . . . . . . . 119

4.4 Two-point functions . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

4.4.1 Scalars on Tp . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

4.4.2 Scalars on the line graph . . . . . . . . . . . . . . . . . . . . . 125

4.4.3 Fermions on the line graph . . . . . . . . . . . . . . . . . . . . 126

4.5 Gauge field dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

4.6 Other sign characters . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

4.7 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

v

Page 7: Real and p-Adic Physics

1

Introduction

Among the remarkable properties of the crossing symmetric version of the Veneziano

amplitude [5]

A(s, t, u) = B(−α(s),−α(u)) +B(−α(s),−α(t)) +B(−α(t),−α(u)), (1.1)

is that it can be expressed [6] as an infinite product over all prime numbers p:

A(s, t, u) =∏p

Ap(s, t, u)−1, (1.2)

where Ap(s, t, u) is the four point amplitude for the open p-adic string [7, 8]. Calcula-

tions in p-adic string theory and p-adic quantum field theory are often more tractable

than their real counterparts, so at the time of its discovery (1.2) suggested an excit-

ing new way to dig up facts about strings. The above relation is an example of what

is known as an adelic identity. These are identities that treat the real line and the

p-adic numbers Qp for each prime impartially, and can have the flavor of expressing

real quantities in terms of their p-adic “building blocks”. The name is in reference to

the ring of adeles A, which is a (restricted) product of R and all of the p-adic number

fields. One can think of Qpprimes p and R as exhausting the collection of topologies

1

Page 8: Real and p-Adic Physics

that can be given to a continuum, and the adeles treat all of these topologies in tan-

dem. We will cover the basics of the p-adic number fields in this chapter, and ease

the reader into the p-adic domain by considering a model that interpolates between

Archimedean and p-adic topologies in chapter 2.

Disappointingly, the 5-point amplitude for the open string was found not obey

such an identity [9], and excitement about p-adic string theory decreased as a result.

A sober view of the above adelic formula might be that it follows from the Veneziano

amplitude being expressible as a special function - a so called Gel’fand-Graev beta

function, that can be defined naturally over the ring of adeles. While the hope of adelic

miracles has been nearly retired, interest in p-adic physics has enjoyed a renaissance

due to the recent acknowledgement of a p-adic version [10, 11] of the AdS/CFT

correspondence [12, 13, 14]. One ends up finding formulas that are extremely similar

to those found in real AdS/CFT. These results would be unsurprising if not for the

fact that most calculations in p-adic AdS/CFT are accomplished via discrete sums

over a discrete version of AdS known as the Bruhat-Tits tree. That is, it is harder

to write off the similarities as being the result of a surface level resemblance between

the real and p-adic integrals that are used to calculate the observables. Moreover,

many of the tools that one employs to study Archimedean AdS/CFT can be ported

over to the p-adic version with varying levels of modification. This includes the use of

Mellin space to calculate amplitudes [15, 16, 17, 18, 19], holographic tensor networks

[20, 21, 22, 11], entanglement entropy and a Ryu-Takanagi-like formula [23, 24], and

a gravity-like action that describes the fluctuation of edge lengths within the p-adic

bulk [25].

Studying the p-adic counterpart of a real theory gives not only adelic formulas,

but it also reveals the aspects of theories which are invariant when one changes the

topology that is given to the continuum. This point is perhaps best appreciated in the

context of O(N) theory [26]. Moreover, sometimes the solution of the p-adic version

2

Page 9: Real and p-Adic Physics

of a theory is easier to come by than for the real theory. For instance, certain p-adic

tensor models can be solved exactly [27], and in chapter 3 we will encounter exact

solutions together with adelic formulas in the context of tensor models with more

complicated interaction terms. The p-adic analogue of spin and of higher dimensions is

possibly the most mysterious of all non-Archimedean subjects. While p-adic theories

with anti-commuting fields have been studied in the context of melonic tensor models,

p-adic superstrings [28, 29, 30, 31], and Gross-Neveu analogues [32], the relationship

between p-adic representations involving non-trivial sign characters and Archimedean

spin remains somewhat elusive. In chapter 4, we describe constructions on the line

graph of the Bruhat-Tits tree that are dual to p-adic CFTs that contain operators

that transform under these signed representations.

The rest of chapter serves as both an introduction to the thesis and as a repository

for foundational material that the rest of the chapters rely on - the organization is as

follows. In section 1.1 we give a brief review of the construction and properties of the

p-adic number line Qp. In section 1.2 we introduce the Bruhat-Tits tree construction,

identify P1(Qp) with its boundary, and discuss the relationship between its symmetry

groups and those of P1(Qp). In section 1.3 we discuss the general properties of func-

tions of a p-adic variable, including the analogues of smoothness and the lack of the

usual derivative. In section 1.5 we discuss multiplicative characters on Qp and there

relationship with the rich variety of quadratic extensions of Qp. There are many high

quality references that review the material in sections 1.1-1.4 of which [33, 34, 35, 36]

are a few. In section 1.7 we lay out the general features of the p-adic statistical field

theories that we will encounter in the rest of this thesis. Following this, we give a

brief review in section 1.6 of some special functions of a p-adic variable together with

their relationship with Archimedean special functions. The original references for this

material are [37, 38], with [34] being a handy manual for calculation. In section 1.8

we discuss the p-adic version of the AdS/CFT correspondence [10].

3

Page 10: Real and p-Adic Physics

1.1 The p-adic numbers

In elementary analysis we learn that the real numbers R are the completion of the

rational numbers Q. Implicit in this construction is the choice d(x, y) to define the

notion of a Cauchy sequence. What privileges the use of the usual notion of absolute

value |p/q| = |p|/|q| to define our distance d(x, y) = |x− y| on Q? In order for d(x, y)

to satisfy the usual axioms for a metric, we must have:

|x| ≥ 0 (non-negativity)

|x| = 0⇔ x = 0 (point seperating)

|x+ y| ≤ |x|+ |y| (triangle inequality)

Additionally, it is natural to require that | · | respect the multiplicative structure of

Q:

|xy| = |x||y|. (multiplicativity)

These are the axioms of a norm over the one dimensional vector space Q and it is

natural to ask whether | · | is the only such norm. The answer is given by Ostrowski’s

theorem [33]

All nontrivial norms on Q are topologically equivalent to either the real abso-

lute value | · | or to one of the p-adic norms | · |p, of which there is one for every

prime p.

The real absolute value is set apart by its respect for Q’s linear ordering and its

Archimedean property:

Given non-zero x ∈ Q there exists a natural number N such that |N · x| > 1.

4

Page 11: Real and p-Adic Physics

By contrast, any norm that doesn’t satisfy the Archimedean property must instead

satisfy the ultrametric triangle inequality:

|x+ y|p ≤ max(|x|p, |y|p). (1.3)

One can go about proving this via the binomial theorem. Suppose we have a norm

for which |k| ≤ 1 for all k ∈ Z. Take x 6= 0, then:

|1 + x|n ≤n∑k=0

∣∣∣∣(nk)∣∣∣∣ |x|k (1.4)

≤ (n+ 1) max(1, |x|n). (1.5)

After taking then n’th root and then taking the limit n→∞, we obtain the inequality

|1 + x| ≤ max(1, |x|). The ultrametric inequality (1.3) quickly follows.

The p-adic norms are defined as follows. By fiat, |0|p = 0. Given any nonzero

rational number x, consider the exponents vp(x) in its prime decomposition 1:

x = ±∏

primes p

pvp(x). (1.6)

One can check that vp(xy) = vp(x) + vp(y) and vp(x + y) ≥ min(vp(x), vp(y)). vp(x)

is known as the p-adic valuation of x, and we can use it to define the p-adic norm for

non-zero2 x:

|x|p = p−vp(x). (1.7)

It is conventional to identify the reals with p = ∞, and we will sometimes write the

real absolute value as | · |∞ to distinguish it from its p-adic counterparts. Just as

1This infinite product is well defined because co-finitely many vp(x) vanish.2One usually uses the conventions p−∞ = 0 and vp(0) = +∞ so that this formula holds for all

x ∈ Q.

5

Page 12: Real and p-Adic Physics

we can complete Q using the real absolute value to form the real numbers, we can

complete Q using the p-adic norm to form what is known as the p-adic numbers Qp.

And just as | · | extends to a norm defined on all real numbers, | · |p extends to a

norm defined on all p-adic numbers. Note that |pk|p = p−k, so 1, p, p2, . . . , pk, . . .

represents a sequence of exponentially smaller numbers with respect to the p-adic

norm. For this reason, we can make sense of p-adic series:

x = a0 + a1p+ a2p2 + a3p

3 + · · · . (1.8)

In fact, all elements x ∈ Qp have a unique representation as such a series [33]:

x = pvp(x)

∞∑n=0

anpn, (1.9)

where a0 6= 0 and an ∈ 0, . . . , p − 1. Note again the characteristic property of a

non-Archimedean absolute value: all integers N ∈ Z have |N |p ≤ 1. In fact, the

p-adic unit ball is known as the ring of p-adic integers

Zp = x ∈ Qp||x| ≤ 1. (1.10)

Any p-adic number x =∑∞

n=νp(x) anpn has a unique decomposition into an integer

part [x] =∑∞

n=0 anpn ∈ Zp and a fractional part x = x− [x]. The unit group of Zp

is identical to the unit sphere in Qp and is denoted by Up = x ∈ Qp||x| = 1.

The group of nonzero p-adic numbers (Q×p ,×) is isomorphic to the group Z× Up

via the correspondence:

Z× Up → Q×p (1.11)

(n, u) 7→ p−nu. (1.12)

6

Page 13: Real and p-Adic Physics

The related decomposition

Q×p =⊔n∈Z

p−nUp (1.13)

is sometimes useful. The collection of balls B = B(x, n)|x ∈ Qp, n ∈ Z where

B(x, n) = |x− y| ≤ pn (1.14)

= x+ p−nZp (1.15)

form a basis for the topology on Qp. Note that the balls defined above are usually what

are known as closed balls (as opposed to open balls), but |x−u| < pn = B(x, n−1)

so this distinction is a matter of convention. By definition B(x, n) are all open sets.

Since Up = tp−1i=1 (i+ pZp) is open and

Qp \ Zp =∞⊔n=1

p−nUp, (1.16)

we find that any B(x, n) is also closed. An interesting consequence of the ultrametric

inequality is that if |x− y| = r then B(x, s) = B(y, s) whenever s ≥ r.

1.2 The Bruhat-Tits Tree and its symmetry group

The uniqueness of the p-adic series expansion (1.9) gives rise to the geometric inter-

pretation of the p-adic numbers (plus infinity) as the boundary of a regular p+1-valent

tree - the Bruhat-Tits tree Tp. To see precisely how this identification comes about,

pick some vertex c ∈ Tp to serve as the tree’s center and fix a path 3 ` that is dou-

bly infinite. These assignments are shown in figure 1.1. We can identify the tree’s

boundary ∂Tp with the collection of infinite paths (or rays) that start at c. Given a

3We require that all edges in a path are distinct, so that there is no backtracking.

7

Page 14: Real and p-Adic Physics

∞· · · · · · 0······

···

· · ·

······

···

· · ·

······

···

· · ·

······

···

· · ·

······

···

· · ·

p−2Up

p−2

p−1Up

p−1

p0Up

p0

p1Up

p1

p2Up

p2c `

Figure 1.1: The Bruhat-Tits Tree.

ray r starting at c, we can construct the following p-adic expansion:

x(r) = pν(r)

∞∑n=0

xn(r)pn, (1.17)

where ν(r) is the (directed) number of steps the ray takes before leaving ` and

xn(r) ∈ 0, . . . , p− 1 denotes the n’th choice of edge (starting at 0) that the ray

takes as it makes its way up the tree. One can see by construction that we actu-

ally only have p − 1 choices at the n = 0’th step, consistent with the requirement

x0(r) = 0. There are two rays of special interest that never leave `, the one going in

the positive ν direction converges to 0 and the one in the negative −ν direction to

∞. Indeed, the correspondence x : ∂Tp → P1(Qp) is one to one.

In the context of the p-adic AdS/CFT correspondence, we will be interested in the

possible symmetries of theories defined on the tree Tp as this will give us insight into

the analogue of the conformal group on Qp. The isometry group of Tp is enormous, and

it will be helpful to have a way of expressing distances between points on the tree in

terms of boundary points. Remarkably, one can write [39] the distance d(a, b) between

two points a, b ∈ Tp in terms of the cross ratio of boundary points x, y, z, w ∈ P1(Qp),

where the intersection of the (undirected) paths x→ y and w → z is exactly a→ b:

p−d(a,b) = |x, y; z, w|p =|x− z|p|y − w|p|x− w|p|y − z|p

. (1.18)

8

Page 15: Real and p-Adic Physics

It’s a simple exercise to show that the cross ratio is invariant under the group

PGL(2,Qp) of linear fractional transformations

x→ ax+ b

cx+ dwhere ad− bc 6= 0. (1.19)

Armed with (1.18) one can identify the isometry group of Tp with the group of func-

tions f on the boundary that preserve the cross ratio [40]:

|f(x), f(y); f(z), f(w)|p = |x, y; z, w|p. (1.20)

We can characterize this group as follows: let f(x) satisfy (1.20). Set h(x) =

f(g−1(x)) where g(x) is the linear fractional transformation satisfying g(f−1(0)) = 0,

g(f−1(1)) = 1, and g(f−1(∞)) = ∞. Then h(x) fixes x = 0, 1, and ∞. Moreover

|h(x)|p = |h(x), h(1);h(0), h(∞)|p = |x, 1; 0,∞|p = |x|p, so h(x) is an isometry. We

follow [40] and identify the conformal group of Qp with cross-ratio preserving trans-

formations:

Conf(Qp) = g h| g ∈ PGL(2,Qp) and h is an isometry fixing 0 and 1. (1.21)

1.3 Functions of a p-adic variable

Since the p-adic numbers are, among other things, a topological space we can talk

about continuous functions with p-adic domain or co-domain. Indeed, if V is any

topological space, f : Qp → V is continuous at x if it takes open neighborhoods

U ⊆ V of f(x) to open pre-images f−1(U) ⊆ Qp. In the case V = R, this is

equivalent to the usual definition from analysis: given ε > 0 there exists δ > 0 such

that |f(x)− f(y)|∞ < ε whenever |x− y|p < δ.

However, the notion of a smooth function f : Qp → R does not generalize. This is

9

Page 16: Real and p-Adic Physics

mainly due to the difficulty of defining an appropriate derivative df(x)/dx. Indeed,

the derivative is meant to be a linear approximation of a function f(x) − f(0) =

f ′(0)x+O(x). But if f(x)−f(0) ∈ R and x ∈ Qp, what then is f ′(0)? An appropriate

replacement for smooth functions turns out to be locally constant functions. For

instance, locally constant functions with compact support are used to construct p-

adic Schwartz spaces in p-adic harmonic analysis [36].

We can define local constancy quite generally: a map g between a topological

space V and a set X is locally constant if every v ∈ V has an open neighborhood O

such that g(o) = g(v) whenever o ∈ O. Local constancy is rather uninteresting in the

Archimedean regime - a locally constant function f : R→ R is simply constant. This

is not so for functions of a p-adic variable. Consider the composition

f(x) = h(|x|p), (1.22)

where h : R→ R is any function. Then f is locally constant - it is constant over each

open sphere SN = x ∈ Qp||x|p = p−N.

1.4 Integration and additive characters

Since the p-adic numbers constitute a locally compact Hausdorff topological group

(under addition) we can uniquely define a Harr measure on Qp by fixing its normal-

ization. The standard choice is to take

µ(Zp) =

∫Zpdx = 1. (1.23)

Using the identity Zp = tp−1i=0 (i + pZp) together with the translation invariance and

additivity of µ, we obtain the equation µ(pZp) = p−1. More generally, multiplying by

10

Page 17: Real and p-Adic Physics

pn scales the volume of a set by p−n. In particular, since Up = Zp \ pZp, we have:

∫Updx = 1− 1

p. (1.24)

Using (1.11) and countable additivity of the measure, we arrive at:

∫Qpf(x)dx =

∑n∈Z

p−n∫Upf(pnu)du. (1.25)

The p-adic integral breaks up into an infinite sum of integrals over the p-adic unit

group.

The additive group Qp is Pontryagin self-dual - all additive characters χ : Qp → S1

are of the form:

χ(x) = exp(2πikx), (1.26)

where k ∈ Qp. Heuristically, the fractional part of kx is used since the remaining

terms in the p-adic expansion are positive integers and don’t change the value of

the exponential. Therefore, sometimes the braces are dropped and one uses the

convention exp(2πiz) = 1 for z ∈ Zp. We define the Fourier transform over Qp with

the following conventions:

f(k) =

∫Qpdx χ(kx)f(x) and f(x) =

∫Qpdk χ(−kx)f(k) (1.27)

Just as the Gaussian is a fixed point of the Fourier transform on R, we can define

a p-adic Gaussian that is a fixed point of (1.27). It turns out that the appropriate

11

Page 18: Real and p-Adic Physics

quantity is the indicator function of Zp:

γp(x) =

1 x ∈ Zp

0 otherwise

. (1.28)

We also have a p-adic delta function:

δ(x) =

∫Qpdk χ(−kx), (1.29)

satisfying the usual delta function identity, this time over Qp:

f(x) =

∫dyf(y)δ(x− y). (1.30)

1.5 Field extensions

By the fundamental theorem of algebra, the only nontrivial algebraic extension of

the real numbers is C. The story is more complicated for Qp, as it turns out to

have nontrivial degree n extensions for all n > 1. Suppose F/Qp is a degree n field

extension. Given a ∈ Q, the mapping x 7→ ax is Qp linear and we define the field

norm (not to be confused with | · |p) of a to be the absolute value of its determinant:

NF/Qp(a) = det(x 7→ ax). (1.31)

Note that NF/Qp is Qp valued. In fact, we can use this to define an absolute value on

F

|a|F = |NF/Qp(a)|p. (1.32)

12

Page 19: Real and p-Adic Physics

Note that for x ∈ Qp we have |x|F = |x|np . We can use this absolute value to define

analogues of Zp, Up and pZp for F

ZF = x ∈ F ||x|F ≤ 1 (1.33)

UF = x ∈ F ||x| = 1 (1.34)

PF = x ∈ F ||x|F < 1. (1.35)

Considered as a subset of the ring Zp, the ideal x ∈ Qp||x|p < 1 is principal -

generated by p. Similarly, one can show that PF is principal and generated by some

ω known as the uniformizing element of PF . Moreover ZF/ωZF = Fq is a finite field

with characteristic p (indeed p ∈ ωZF ) and |ω|F = 1/q. Since Fq has characteristic p,

we have q = pf for some f . The uniformizing element essentially plays the same role

in F that p did for Qp, for instance any a ∈ F has a unique representation

a = ωνF (a)∑n≥0

anωn, (1.36)

where a0 6= 0 and an ∈ ZF/ωZF .

Note that |p|F < 1, so there must be some e ≥ 1 such that p = ωeu where u ∈ UF .

This lead us to the following identity: p−n = |p|F = |ωe|F = p−ef , so n = ef . The

integer e is known as the ramification degree of the extension. The case e = 1 is that

of an unramified extension, whose uniformizing element can be chosen to be p. The

case e = n is that of a totally ramified extension, whose uniformizing element can be

chose to be p1/n. For a general ramification index, the uniformizer can be chosen to

be p1/e. Many of the contents of this thesis have natural generalizations to extensions

of Qp, but we have chosen to stick to Qp in these situations for simplicity.

The case n = 2 will be of special interest to us, so we develop it further here.

The quadratic extensions of Qp together with the trivial extension Qp(1) = Qp are

13

Page 20: Real and p-Adic Physics

in one to one correspondence with the factor group Q×p /(Q×p )2. We can refine the

decomposition (1.13) further by noting that Up ' F×p × U1 where U1 = 1 + pZp is a

subgroup of Up. One can show [36] that for odd p, we have U21 = U1 so that

[Q×p : (Q×p )2] = 4. (p > 2). (1.37)

There are three corresponding nontrivial quadratic extensions of Qp(√τ) for p > 2.

They are generated by√τ =√p,√ε, and

√pε where ε is a (p − 1)’th root of unity

in Qp.

The case of p = 2 is trickier. To see why, suppose we are given 1 + 2z where

|z|2 < 1. Then (1 + 2z)2 = 1 + 4z + 4z2 = 1 + O(23). In fact, for p = 2 we have

U21 = 1 + 23Zp. We then have

[Q×2 : (Q×2 )2] = 8. (p = 2). (1.38)

There are seven corresponding nontrivial quadratic extensions Q2(√τ) of Q2. The

generators can be chosen to be√τ where τ = −1,±2,±3, and ± 6.

1.6 Multiplicative characters and Gel’fand-Graev

gamma functions

The goal of this section is to introduce special functions due to Tate, Gel’fand, and

Graev [37, 38]. The presentation is clearest if let K = R,C, and Qp and treat all

three cases in tandem. We gave a classification of the additive characters over Qp in

(1.26). By a multiplicative character, we mean a group homomorphism π : K× → C×.

For a given additive character χ : K → C, the Gel’fand-Graev gamma function

14

Page 21: Real and p-Adic Physics

associated with a multiplicative character π : K → C is defined by

Γ(π) =

∫K

dx

|x|Kχ(x) π(x) . (1.39)

The multiplicative characters that are most relevant to this thesis are

πs(t) ≡ |t|sK , πs,sgn(t) ≡ |t|sK sgn(t) . (1.40)

where a sign function sgn(t) is any multiplicative character taking only the values ±1.

It is straightforward to show using the definition (1.39) that the Fourier transforms

of the multiplicative characters (1.40) are given by

F [πs](ω) = Γ(πs+1) π−s−1(ω) , F [πs,sgn](ω) = Γ(πs+1,sgn) π−s−1,sgn(ω) . (1.41)

In order to write down explicit expressions for the Gel’fand-Graev gamma func-

tions associated with R and Qp and the multiplicative characters (1.40), it is expedient

to introduce the local zeta functions ζ∞, ζp : C→ C:

ζ∞(s) = π−s2 ΓE

(s2

), ζp(s) =

1

1− p−s, (1.42)

where ΓE(s) is the familiar Euler gamma function. ζp(s) = (1 − p−s)−1 is so named

because of the adelic relation ζ(s) = Πpζp(s).

For K = R, the Gel’fand-Graev gamma functions are given by

Γ(πs) =ζ∞(s)

ζ∞(1− s), Γ(πs,sgn) = i

ζ∞(1 + s)

ζ∞(2− s). (1.43)

For K = C there are no sign functions, and the Gel’fand-Graev gamma function

15

Page 22: Real and p-Adic Physics

is given by

Γ(πs) = (2π)−2s (ΓE(s))2 sin(πs) . (1.44)

As we saw in section 1.5, for K = Qp, there are multiple distinct quadratic extensions

Qp(√τ). Associated with each of these extensions, there is a sign function, which

takes the value 1 on the image of the field norm NQp(√τ)/Qp

sgnτ (x) =

1 x = a2 − τb2 for some a, b ∈ Qp

−1 otherwise

. (1.45)

We can rephrase this definition as follows: given z = a +√τb ∈ Qp(

√τ), define the

conjugate z∗ = a−√τb. Then sgn(x) = 1 if and only if x = z∗z for some z ∈ Qp.

For p = 2, there are seven distinct non-trivial sign functions corresponding to

τ = −1, ±2, ±3, and ±6, and the gamma functions evaluate to

Γ2(πs) =ζ2(s)

ζ2(1− s),

Γ2(π(−3)s ) =

ζ2(1− s)ζ2(2s)

ζ2(2− 2s)ζ2(s),

Γ2(π(−1)s ) = Γ2(π(3)

s ) = i4s

2,

Γ2(π(−2)s ) = −Γ2(π(6)

s ) = i8s√

8,

Γ2(π(2)s ) = −Γ2(π(−6)

s ) =8s√

8.

(1.46)

For p > 2 there are three distinct non-trivial sign functions, which can be labeled

by τ equal to p, ε, and εp, where ε is an integer that is not a square modulo p. The

16

Page 23: Real and p-Adic Physics

Gel’fand-Graev gamma functions are given by

Γp(πs) =ζp(s)

ζp(1− s), Γp(π

(ε)s ) =

ζp(1− s)ζp(2s)ζp(2− 2s)ζp(s)

,

Γp(π(p)s ) =

ps√p,

−i ps

√p,

Γp(π(εp)s ) =

− ps√p

for p ≡ 1 mod 4 ,

ips√p, for p ≡ 3 mod 4 .

(1.47)

1.7 p-Adic field theory

Statistical field theory over the p-adic numbers can be defined in the same way as

field theory over R. We consider an action functional S(φ) on a space of fields φ(x)

of a p-adic variable x ∈ Qp and form the partition function

Z(J) = Z(0)

∫Dφ exp(−S(φ) + φ · J), (1.48)

where φ · J =∫Qp φ(x)J(x)dx is shorthand for the inner product between functions

formed by the p-adic integral. There is no clear p-adic equivalent of the derivative on

R, so it is easier to start by writing p-adic actions in momentum space. The quadratic

part of the action for a scalar field takes the form:

Squad(φ) =1

2

∫Qpφ(−ω)(|ω|s + r)φ(ω). (1.49)

Introducing the Vladimirov derivative of weight s

Dsφ(x) =

∫Qpdyφ(x)− φ(y)

|x− y|s+1, (1.50)

17

Page 24: Real and p-Adic Physics

one can write the action in position space

Squad(φ) =1

2

∫Qpdx(φ(x)Dsφ(x) + rφ(x)2). (1.51)

The free two-point function associated with the action (1.49) is simply

〈φ(ω1)φ(ω2)〉0 = G(ω1)δ(ω1 + ω2), (1.52)

where

G(ω) =1

|ω|s + r. (1.53)

We will also have occasion to consider actions involving N ≥ 1 real commuting or

anti-commuting fields ψi. To that end, we take (1.49), generalize to K = R,C, or Qp,

drop the mass term, replace |k|s with an arbitrary multiplicative character π, and add

index structure:

Sπ(ψ) =1

2

∫dωψi(−ω)Ωijπ(ω)ψj(ω). (1.54)

where π : K× → C× is a multiplicative character.

We will choose Ωij to be either symmetric or anti-symmetric. For the symmetric

case we define σΩ = 1 and set Ω = δij. For the anti-symmetric case we let σΩ = −1

and set Ω = σ2 ⊗ 1N/2 to be a standard symplectic form. Note that N must even in

order for the index structure to be non-singular since det(ΩT ) = (−1)N det(Ω). We

encode the statistics of ψ in the following identity ψiψj = σψψjψi. So σψ = 1 for

commuting fields, and σψ = −1 for anti-commuting fields. Then making the change

18

Page 25: Real and p-Adic Physics

of variable ω → −ω, we write

Sπ(ψ) =1

2

∫dωψi(−ω)Ωijπ(ω)ψj(ω) (1.55)

=1

2

∫dωψi(ω)Ωijπ(−ω)ψj(ω) (1.56)

= π(−1)σΩσψSπ(ψ). (1.57)

Specializing to characters of the form πs,sgn(ω) = |ω|s sgn(ω), we obtain the constraint

σΩσψ sgn(−1) = 1. (1.58)

The free two-point function associated with (1.54) is:

G(ω) =Ωij

πs,sgn(ω)= Ωij sgn(ω)|ω|−s, (1.59)

where ΩijΩjk = δij.

1.8 p-Adic CFT and the p-adic AdS/CFT corre-

spondence

Not long after Freund and Olson’s discovery of the p-adic string [7], an axiomatic

framework for p-adic conformal field theory was suggested in [41]. In Melzer’s formu-

lation, PGL(2,Qp) plays the role of the global conformal group. Anticipating p-adic

AdS/CFT, we note here that the symmetry group of the Bruhat-Tits tree is larger

(1.21) than just PGL(2,Qp). However, we will only find need to consider representa-

tions that are trivial on the subgroup of isometries h(x) with h(0) = 0 and h(1) = 1.

Melzer’s framework consists of locally constant fields φ : Qp → C. Because of a

lack of a derivative on these fields, one cannot use the standard techniques of quantum

theory that make use of the Lie algebra of ones symmetry group. In other words,

19

Page 26: Real and p-Adic Physics

we cannot make use of infinitesimal symmetry transformations. Thus Melzer calls

primary operators those φ(x) that transform under PGL(2,Qp) as

φ′(x′) =

∣∣∣∣ ad− bc(cx+ d)2

∣∣∣∣−∆φ

p

φ(x) for x′ =ax+ b

cx+ d. (1.60)

That is, in p-adic CFT we have no descendants and only have pseudo-primary oper-

ators. We simply call these operators primary in the p-adic context. As standard in

conformal field theory, using the representation (1.60) one can easily determine the

form of the (normalized) two-point function:

〈φi(x)φj(0)〉 = δ∆i,∆j

1

|x|2∆ip

. (1.61)

In p-adic CFT without descendants we have the following OPE:

φi(x)φj(0) =∑k

Cijk|x|−∆i−∆j+∆kp φk(0), (1.62)

where the sum runs over p-adic primaries. The three point function therefore takes

the form, for |x2 − x3|p < |x1 − x2|p:

〈φi(x1)φj(x2)φk(x3)〉 =∑`

Cij`|x23|−∆j−∆k+∆`p 〈φi(x1)φ`(x3)〉 (1.63)

= Cijk|x12|−∆ij,kp |x13|

−∆ik,jp |x23|

−∆jk,ip , (1.64)

where the last equality follows from the identity |x12| = |x13| and we use the shorthand

∆i1...in,j1...jm = ∆i1 + · · · + ∆in − ∆j1 − · · · − ∆jm together with xij = xi − xj. This

result can also be obtained by performing an appropriate PGL(2,Qp) transformation

on 〈φi(∞)φj(0)φk(1)〉. In p-adic CFT, the conformal blocks take a very simple form.

20

Page 27: Real and p-Adic Physics

Consider the following four-point function for |x| ≤ |1− x| ≤ 1. We have:

〈φi(∞)φj(1)φk(0)φ`(x)〉 =∑m

CijmCk`m|x|−∆k`,mp . (1.65)

Similar formulas result from taking |1 − x|p ≤ |x|p = 1 and 1 ≤ |x|p = |1 − x|p, and

taking x = 1 we can derive the consistency condition [41]

∑m

CijmCklm =∑m

CilmCkjm =∑m

CikmCjlm. (1.66)

The OPE associativity condition boils down to an associativity condition on the

structure constants of a commutative algebra.

We will will have occasion to consider more general projective representations of

PGL(2,Qp), that satisfy

φ′(x′) =

√sgn

(ad− bc

(cx+ d)2

) ∣∣∣∣ ad− bc(cx+ d)2

∣∣∣∣−∆φ

p

φ(x) for x′ =ax+ b

cx+ d, (1.67)

and lead to the two point function

〈φ(x)φ(0)〉 = sgn(x)1

|x|−2∆φ. (1.68)

Whether φ must be commuting or anti-commuting in this context depends on sgn(−1)

and its behavior under Hermitian conjugation.

Looking back, the roots of p-adic AdS/CFT lie in a paper by Zabrodin [42]. He

showed that the non-local p-adic string action

S(O) ∼∫dxdy

(O(x)−O(y))2

|x− y|2p(1.69)

21

Page 28: Real and p-Adic Physics

could be obtained by considering a lattice field theory

S(ϕ) =1

2

∑A∼B

(ϕA − ϕB)2, (1.70)

where A ∼ B indicates that the sum ranges over all neighboring vertices on Tp.

One integrates out the interior of Tp and obtains (1.69). In hindsight, this can be

interpreted as a holographic computation of the two-point function 〈O(x)O(0)〉 ∼

|x|−2.

In [10] a full framework for holographic calculations on Tp was developed. Many

of the results are well expressed by adopting the following coordinate system on the

tree. We note that any vertex in the tree can be uniquely identified by a pair (x, z)

where x ∈ Qp is a boundary point and z = pn indicated the “depth” of the vertex

along the tree. Formally, we have the vertex equality A = B when zA = zB and

|xA − xB| ≤ |zA|.

In the simplest variation of p-adic AdS/CFT we consider the action (1.70) and

add a mass term since we anticipate it to be related to the scaling dimension of the

dual boundary operator:

S(ϕ) =1

2

∑A∼B

(ϕA − ϕB)2 +∑A∈Tp

1

2m2φ2

A. (1.71)

The Green’s function for A,B ∈ Tp is

GAB = ζp(2∆)p−∆d(A,B), (1.72)

where m2∆ = (−ζp(−∆)ζp(∆− 1))−1. Compare this to the real case m2

∆ = ∆(∆− n).

The bulk to boundary propagator K(A, x) satisfying the normalization condition

∫QpK(A, x)dx = |zA|1−∆

p , (1.73)

22

Page 29: Real and p-Adic Physics

can be written

K(A, x) =ζp(2∆)

ζp(2∆− 1)

|zA|∆pmax(|zA|p, |xA − x|p)2∆

. (1.74)

In Fourier space:

K(A, k) =

(|zA|1−∆ + |zA|∆|k|2∆−1

p

ζp(1− 2∆)

ζp(−1 + 2∆)

)γp(kzA). (1.75)

The correlators of the boundary theory can then be computed using the standard

AdS/CFT dictionary:

⟨exp

(∫Qpdx φ0(x)O(x)

)⟩= e−Son-shell(φ), (1.76)

where the on-shell action is evaluated on φ satisfying

limzA→0

|zA|∆−1φ(zA, x) = φ0(x). (1.77)

In practice, the correlation functions are a sum of diagrams. To compute the dia-

grams, one sums the relevant bulk-to-boundary propagators over all of Tp [10]. More

advanced methods for computing these diagrams can be found in [43] together with

the Mellin space approach taken in [18, 19].

23

Page 30: Real and p-Adic Physics

2

Continuum Limits of Sparsely Coupled Ising-like

Models

This chapter is based on [1] coauthored with Steven S. Gubser, Christian Jepsen, and

Ziming Ji. We thank S. Hartnoll for getting us started on this project by putting us

in touch with M. Schleier-Smith’s group, and we particularly thank G. Bentsen and

M. Schleier-Smith for extensive discussions.

The study of statistical field theory over the p-adic numbers arguably began with

Dyson’s hierarchical model [44]. A rigorous study of Dyson’s phase transition was

carried out in [45] and the identification of p-adic field theory with the continuum

limit of Dyson’s model was made in [46]. The ideas found in this chapter can be

understood, for p = 2, in terms of rephrasing of Dyson’s original work.

Consider the “furthest neighbor” Ising model. By this we mean start with 2N

Ising spins labeled by sites 0, . . . , 2N − 1 and strongly couple each spin σi to the spin

that is sequentially furthest from it σi+2N−1 , where the addition takes place in Z/2NZ.

This produces 2N−1 pairs of strongly coupled spins where each pair is decoupled from

all other pairs. In the interest of seeking a more interesting thermodynamic limit, we

proceed by coupling each pair of spins with the pair of spins that is furthest away

from it. We then couple pairs of pairs, and so on. At each stage of this process we

24

Page 31: Real and p-Adic Physics

0

1

2

3

4

5

6

7

0 4 2 6 1 5 3 7

000 100 010 110 001 101 011 111

000 001 010 011 100 101 110 111

Monna map M

h

M(h)

Figure 2.1: Left: A furthest neighbor coupling pattern among eight spins. The

thickness of lines indicates the strength of the coupling between spin 0

and the other spins. The coupling pattern is invariant under shifting

by a lattice spacing, so for example spins 1 and 5 are as strongly

coupled as spins 0 and 4. The blue circle is to guide the eye and does

not indicate additional couplings.

Right: A hierarchical representation of the couplings between spins.

Above each spin’s label we have given the base 2 presentation of the

spin number, and we have shown how the Monna map acts on these

numbers by reversing digits in the base 2 presentation.

reduce the coupling strength by a fixed factor 21+s, where s ∈ R is what we will call

the spectral parameter. The overall picture is illustrated in figure 2.1.

It is natural to view the furthest neighbor model as a hierarchy of spin clusters,

also shown in figure 2.1. This tree of clusters gives a particularly clear understanding

of the 2-adic distance. Indeed, if we define d(i, j) to be the number of steps required

to go from i to j via a path through the tree, then we have the following identity:

|i− j|2 = 2−N+d(i,j)/2 (2.1)

for distinct i and j. A quick comment on how to interpret this formula is now

appropriate. For i ∈ Z/2NZ we let |i|2 = |i∗|2 where i∗ is the representative of i in

0, . . . , 2N − 1. Note that this convention sets |0 + 2NZ|2 = 0 consistent with the

usual |0|2 = 0 for 0 ∈ Q but not with (2.1) for coincident points. It’s also important

to note that while | · |2 is not an absolute value on Z/2NZ (e.g. multiplicativity

25

Page 32: Real and p-Adic Physics

fails), the statement |u|2 = 1 is actually independent of the representative chosen for

u ∈ Z/2NZ. Green’s functions in the furthest neighbor Ising model depend only on

the 2 -adic distance |i− j|2, and this is a consequence of the partition function being

invariant under a relabeling of the lattice sites i→ ui+ b where |u|2 = 1. Intuitively,

|u|2 is a rotation and rotational invariance implies that physical quantities depend

only on the norm |i− j|2.

The sparse coupling pattern that we study in this chapter couples spins i, j ∈

Z/2NZ only if i−j is a power of 2 (modulo 2N). With such a coupling, the number of

spins coupling to spin 0 grows as the logarithm of the system size N = log2(2N). We

will see that when the spectral parameter of this model is made large and negative

that, to a good approximation, only nearest neighbor couplings survive.

Less obvious is that when the spectral parameter is large and positive, the model

behaves to a good approximation like the furthest neighbor model. Here’s an expla-

nation. When s 0, the coupling between spins 0 and 2N−1 produces very tightly

coupled pairs. When we proceed to the next layer of the tree pairs of pairs are also

tightly coupled. Proceeding to the next layer, we couple 0 weakly to ±2N−3 but not

±3× 2N−3. However, what matters is that we are coupling the already tightly bound

quartets 0, 2N−1,±2N−2 and ±2N−3,±3× 2N−3. This logic works at general lay-

ers of the hierarchy - while we aren’t coupling all constituents of one spin cluster to

all of the constituents of another spin cluster, what matters is that the clusters are

already tightly bound relative to the the strength of the inter-cluster coupling.

The Green’s function of nearest neighbor model with 2N spins is well-approximated

at large N by a continuum Green’s function that can be extracted from statistical

field theory over R. This two-point function is smooth away from the origin, and

this smoothness gives a good way of understanding how the continuous topology of

R emerges from a discrete lattice. As reviewed in section 1.3, the accepted analogue

of a smooth function f : Qp → R is a locally constant function. Indeed, the Green’s

26

Page 33: Real and p-Adic Physics

functions arising from the p-adic field theories in section 1.7 are locally constant.

When we study sparse coupling patterns, we will find that the Green’s functions

are well approximated by locally constant functions in the limit of large positive s.

However, we will have to use a more subtle measure of strong continuity - Holder

continuity - to characterize the transition from large positive s to large negative

s. The long and short of it is that as s is varied from positive to negative, the

Green’s function transitions from being 2-adically continuous to being continuous in

the Archimedean sense.

The organization of the rest of this chapter is as follows. In section 2.1 we describe

the class of one-dimensional statistical mechanical models that we will study and give

a general account of the form of the two-point function in these models. In sections

2.2-2.5 we treat four models within this class. In order: Nearest neighbor interactions,

p-adic interactions, power-law interactions, and finally the sparse coupling model

which interpolates between the nearest neighbor and 2-adic models. After a brief

discussion of Holder continuity in section 2.6 we derive Holder continuity bounds for

the continuum limit of the sparsely coupled Green’s function in sections 2.7-2.8. In

sections 2.9-2.11 we show through numerical studies that to a good approximation, the

continuity of the discrete momentum space Green’s function is characterized by these

Holder bounds. The continuity of the position space Green’s function is characterized

partly by the continuum Holder bounds. Indeed, a stronger form of continuity than

one might expect from field theoretical results emerges in particular regimes.

2.1 The Statistical Mechanical Framework

We wish to study the general properties of non-local couplings. Hence, we consider

an elementary class of statistical mechanical models consisting of L real degrees of

27

Page 34: Real and p-Adic Physics

freedom φi together with a Hamiltonian

H = −1

2

∑ij

Jijφiφj −∑j

hjφj, (2.2)

which is a sum of a quadratic form in φi together with an external source. We view

the degrees of freedom as sitting on L lattice sites and impose translation invariance

of the coupling, i.e. Jij = Ji−j where i − j is understood modulo L. As with any

quadratic form, we can rearrange the coefficients to be symmetric: Jij = Jji. In the

single subscript notation Jh = J−h.

The translation invariance and symmetry of J guarantee that we can unitarily

diagonalize J via a discrete Fourier transform. Specifically, we define an orthonormal

basis for CL:

(uk)j =1√L

exp

(2πikj

L

)for k = 0, 1, 2, . . . , L− 1. (2.3)

We then use the Fourier transform convention Xk = u†kX for X ∈ CL. The uk are

eigenvectors of J :

Juk =√LJkuk, (2.4)

where Jk is the Fourier transform of the coupling strength vector (J0, . . . , JL−1). The

eigenvalue equation (2.4) allows us to write the Hamiltonian in momentum space

H = −√L

2

L−1∑k=0

Jkφ−kφk −L−1∑k=0

h−kφk. (2.5)

In order for the integral defining the partition function

Z(h) =

∫dLφ exp(−βH(φ, h)) (2.6)

=

∫dLφ exp

[1

2φTJφ+ hTφ

])(2.7)

28

Page 35: Real and p-Adic Physics

to converge, J must be negative definite. However, we want to study the massless case

in which the Hamiltonian depends only on the squared differences of the parameter

between lattice sites (φi−φj)2. This is equivalent to invariance under the target space

translation φ → φ + cu0 where u0 = (1, . . . , 1) and c is any constant. Therefore, we

require a vanishing eigenvalue J0 = 0 of J . To overcome this infrared divergence, we

will regulate by setting J = −µ where µ > 0 and take the limit µ→ 0 after discarding

uninteresting constants in our formulas.

Having established it’s convergence, the partition function takes the standard

form:

Z(h) = Z(0) exp

(−β

2hTJ−1h

)(2.8)

= Z(0) exp

(− β

2√L

L−1∑k=0

1

Jkh−khk

). (2.9)

The two-point function Gij = 〈φiφj〉 is also translation invariant Gij = Gi− j and

can be calculated directly:

Gh = − 1

βL3/2

L−1∑k=1

1

Jkexp

(2πikh

L

)− 1

µβL3/2. (2.10)

Dropping the constant divergent as µ → 0, we arrive at the formula for the Green’s

function that we will use for the rest of this chapter:

Gh = − 1

βL3/2

L−1∑k=1

1

Jk

(2πikh

L

). (2.11)

It will also be convenient to have a condition on the position space couplings Jh that

ensures Jk < 0 for k 6= 0 and J0 = 0. One can check it is enough to have∑L−1

h=0 Jh = 0

and Jh ≥ 0 for h 6= 0 with the strict inequality holding for at least one h.

As mentioned at the begging of the chapter, we will be studying a sparsely cou-

29

Page 36: Real and p-Adic Physics

pled model that interpolates between Archimedean and p-adic coupling. Before in-

troducing it, we will introduce the models it interpolates between, starting with the

Archimedean model.

2.2 Nearest neighbor coupling

Dyson showed in [44] that phase transitions could occur in a one dimensional Ising

model given the right coupling strength. While there is still a lack of a phase transition

in the one-dimensional Ising model whose sites are coupled along a line, Dyson’s model

describes sites that are coupled hierarchically. If we take the view that the coupling

determines the topology of model instead of the indexing of the fields, then the nearest

neighbor coupling

JNNH = J∗(δh+1 + δh−1 − 2δh), (2.12)

describes the (Archimedean) topology of degrees of freedom distributed on a circle.

The corresponding two point function is:

GNNh =

1

4βJ∗L

L−1∑k=1

e2πikh/L

sin2 πkL

. (2.13)

Taylor expanding the 1/ sin2 term, we arrive at:

1

LGNNh =

1

4βJ∗L2

L−1∑k=1

e2πikh/L

(L

πk

)2∑n≥0

ζ(2n)(2n− 1)

(k

L

)n(2.14)

=1

βJ∗

L−1∑k=1

e2πikh/L

8π2k2+

1

β4J∗

∑n≥0

ζ(2n)(2n− 1)1

Ln

L−1∑k=1

kn−2e2πikh/L. (2.15)

30

Page 37: Real and p-Adic Physics

The last term is O(1/L), so we can take the L→∞ limit and write:

GNNh ∼ L

βJ∗G(h/L), (2.16)

where the continuum Green’s function on the circle R/Z for a representative x ∈ [0, 1]

is:

G(x) =1

2

(x− 1

2

)2

− 1

24. (2.17)

We have the Green’s function identity G′′(x)− 1 = −δ(x). We also have∫R/ZG = 0.

2.3 p-Adic coupling

We now introduce the p-adic coupling. Fix a prime number p and consider a model

with L = pN lattice sites. We can then define the following all-to-all coupling for

h 6= 0:

Jp−adich = J∗|h|−s−1

p . (2.18)

This coupling endows the model with the hierarchical topology that we saw in section

2. We will write this as Jph for the rest of the section so as to not encumber the reader

with verbose notation. Using the constraint Jp0 = 0 one can derive the following

recursion relation on N : Jp0 |N+1 = p(J0|N − J∗(1− p−1)ps(N+1)). Solving this with a

geometric sum, we find for L = pN

Jp0 = −J∗Lζp(−s)

ζp(1)ζp(−sN). (2.19)

Let us turn our attention to calculating the Fourier transform of the coupling. It

will be helpful to first compute the transform of f(N)h =

√L|h|−s−1

p (1− δh) over CpN

since it will reappear in our calculations for the Fourier transform of the Green’s

function. It’s easy to show that f(N)k depends only on |k|p, so it is enough to find

31

Page 38: Real and p-Adic Physics

it’s value for k = 0, 1, p, . . . , pN−1. We first derive a recursion relation for f(N)h . Let

0 ≤ −logp|k| < N , then kp < pN and:

f(N)kp = f

(N−1)k +

pN−1∑h=pN−1

|h|−s−1p e2πihk/pN−1

(2.20)

= pf(N−1)k + pN+(N−1)s. (2.21)

This simplifies if we express it using the following normalized quantities g(N)k =

p−N(s+1)f(N)k . We then have:

gNkp = p−s(gN−1k +

1

ζp(1)

). (2.22)

We now make the ansatz g(N)k = AN |k|sp + BN . Note that this conforms with our

intuition from Archimedean physics, where the Fourier transform of |x|−s−1 scales

under k → λk like |λ|s+1−d where d is the dimension of the space. The recursion

relation (2.22) implies that AN = A is independent of N and gives the relation

BN = p−s(BN−1 + 1/ζp(1)). This has constant (B = BN) solution

B = −ζp(−s)ζp(1)

. (2.23)

Lastly, we use A = B − g(1)1 to find

A =ζp(−s)ζp(1 + s)

. (2.24)

Putting this all together, we have for k 6= 0:

f(N)k = Ls+1ζp(−s)

( |k|spζp(1 + s)

− 1

ζp(1)

). (2.25)

32

Page 39: Real and p-Adic Physics

For k = 0 one has immediately f(N)0 = −Jp0/J∗. We can use these results to easily

compute the Fourier transform of Jph = J∗(f(N)h /√L− f (N)

0 δh):

Jpk =J∗√L

(f(N)k − f (N)

0 ). (2.26)

From this equation we manifestly have J0 = 0. For k 6= 0 our result simplifies to

Jpk = J∗√Lζp(−s)ζp(1 + s)

(∣∣∣∣ kL∣∣∣∣s − ζp(1 + s)

ζp(1)

). (2.27)

We will be interested in the regime s > 0 where the first term in (2.27) is larger than

the second. We can therefore write for k 6= 0:

Gpk = − 1

βLJpk= − 1

βL3/2J∗

ζp(1)

ζp(−s)∑n≥1

(ζp(1 + s)

ζp(1)L−s

)n|k|−nsp . (2.28)

We can now make use of the results derived in this section to apply the inverse Fourier

transform and obtain

Gph = − 1

βL2J∗

ζp(1)

ζp(−s)∑n≥1

ζp(1 + s)n

ζp(1)n

[ζp(−ns+ 1)

ζp(ns)

(|h|ns−1

p − ζp(ns)

ζp(1)

)(1− δh)

(2.29)

− ζp(−ns+ 1)

ζp(1)ζp(N(−ns+ 1))δh

]. (2.30)

Note that L2Gph is independent of L for h 6= 0. Indeed, only the allowed range of h

changes as we increase L = pN . Taking L→∞ corresponds to the range of h becom-

ing dense in the of p-adic integers Zp = x ∈ Qp||x|p ≤ 1. It is worth noting that

since L2Gph depends on h only through its norm |h|p, the continuum Green’s function

is locally constant. This local constancy guarantees a locally constant extension of

L2Gph to all of Zp. More explicitly, given x ∈ Zp we can simply evaluate the Green’s

function at |x|p.

33

Page 40: Real and p-Adic Physics

2.4 Power-law coupling

The p-adic coupling introduced in the last section and the sparse coupling to be intro-

duced in the section following this one both include an adjustable spectral exponent

s. Since the nearest neighbors coupling has no such exponent, we modify it to include

one:

Jpowerk = − J∗

2s√L

sin−s(πk

L

). (2.31)

This reduces to the nearest neighbor model when s = −2. Let us justify this coupling’s

name. For s < 1, and L 1 we can approximate

Jh = −J∗2s

1

L

L−1∑k=0

sin−s(πk

L

)e−2πikh/L (2.32)

' −J∗2s

∫ 1

0

sin−s(πx) exp(−2πihx)dx (2.33)

= −J∗π

sin(πs/2)Γ(1− s)Γ(h+ s/2)

Γ(1 + h− s/2). (2.34)

Using Stirling’s approximation for 1 h L, we verify that the coupling obeys a

power law:

Jpowerh ∼ −J∗

πsin(πs/2)Γ(1− s)hs−1. (2.35)

Given s < −1 and in the limit L 1, the position space Green’s function can be

expressed in terms of the polylogarithm Lin(x) =∑∞

k=1 xk/kn:

Gpowerh =

21+sπs

βJ∗L1+sRe Li−s(e

2πih/L). (2.36)

34

Page 41: Real and p-Adic Physics

2.5 Sparse coupling: From the Archimedean to

the non-Archimedean

We now exhibit the primary model that we are interested in studying. Let L = 2N

for some integer N ≥ 1 and consider the following coupling:

J sparseh = J∗

N−1∑n=0

2ns(δh−2n + δh+2n − 2δh). (2.37)

We refer to this coupling as sparse because any particular degree of freedom will be

coupled to at most O(logL) other degrees of freedom. It is immediately apparent that

for s sufficiently negative this sparsely coupled model will become well approximated

by a nearest neighbor model. In particular, with suitable normalization, we expect

Gsparseh to closely approximate GNN

h . What is not so immediately apparent is the

following:

1. For s sufficiently positive, G2-adich closely approximates Gsparse

h .

2. As s is varied from negative to positive values, Gsparseh transitions from exhibiting

Archimedean continuity to exhibiting 2-adic continuity. A precise statement

about the form that this continuity takes will be made in the next few sections

when we discuss Holder bounds and study the continuum Green’s function of

the sparsely coupled model.

We are interested in the notion of closeness not only in the Archimedean regime, but

in the p-adic regime as well. We therefore find it useful to employ a discrete variation

of the Monna map. We can express any h ∈ Z/pNZ in a p-adic expansion:

h =N−1∑n=0

hnpn where hn ∈ 0, 1, . . . , p− 1. (2.38)

35

Page 42: Real and p-Adic Physics

The hn in the above expression are unique, so we can define a single valued function:

M(h) =N−1∑n=0

hN−1−npn. (2.39)

Our discrete Monna map is an involutionM−1 =M. Suppose h and g are p-adically

close, i.e. hn = gn for n < M where 1 M < N . Then M(h) and M(g) are

sequentially close:

|M(h)−M(g)|∞ =

∣∣∣∣∣N−1∑n=M

(hn − gn)pN−1−n

∣∣∣∣∣ ∼ pN−M−1 (2.40)

The converse is not true: sequentially close numbers can be mapped to numbers that

aren’t p-adically close.

2.6 Continuity conditions

In the following sections we will investigate continuum field theories with sparse cou-

pling patterns. In section 2.7 we study 2-adic field theories and derive Holder con-

tinuity bounds for the two-point function in both position and momentum space.

We follow this up in section 2.8 with similar bounds in a sparsely coupled bilocal

Archimedean field theory.

Since the bounds we will encounter in the following sections are all Holder bounds,

it will be useful to review the appropriate definitions. Let X and Y be metric spaces.

In this chapter, we will generally have Y = R and X ⊆ Qp or X ⊆ R. The metric

on X is simply dp(n,m) = |n−m|p for subspaces of Qp and d∞(x, y) = |x− y|∞ for

subspaces of R. Let α > 0 be given. A function f : X → Y is said to be α-Holder

continuous on X if there exists some K > 0 such that given any x1, x2 ∈ X, the

36

Page 43: Real and p-Adic Physics

s = -1Gh

sparseGh

power

10 20 30 40 50 60h

-0.2

0.2

0.4

0.6

Gh/G0

s = -1Gh

sparseGh

2-adic

1 2 3 4 5 6log2(h)

-0.2

0.2

0.4

0.6

Gh/G0

s = -0.3Gh

sparseGh

power

10 20 30 40 50 60h

-0.1

0.1

0.2

0.3

Gh/G0

s = -0.3Gh

sparseGh

2-adic

1 2 3 4 5 6log2(h)

-0.1

0.1

0.2

0.3

Gh/G0

s = 0.3Gh

sparseGh

power

10 20 30 40 50 60h

-0.1

0.1

0.2

0.3

0.4

Gh/G0

s = 0.3Gh

sparseGh

2-adic

1 2 3 4 5 6log2(h)

-0.1

0.1

0.2

0.3

0.4

Gh/G0

s = 1Gh

sparseGh

power

10 20 30 40 50 60h

-0.2

0.2

0.4

0.6

0.8

1.0

Gh/G0

s = 1Gh

sparseGh

2-adic

1 2 3 4 5 6log2(h)

-0.2

0.2

0.4

0.6

0.8

1.0

Gh/G0

Figure 2.2: Left: Gsparseh and Gpower

h versus h. This column shows how close

Gsparseh is to a smooth function in the usual Archimedean sense, and

confirms that Gsparseh ≈ Gpower

h when s is sufficiently negative.

Right: Gsparseh and G2−adic

h versus log2 M(h). This column shows how

close Gsparseh is to a smooth function in the 2-adic sense, and confirms

that Gsparseh ≈ G2−adic

h when s is sufficiently positive.

37

Page 44: Real and p-Adic Physics

following inequality is satisfied:

dY (f(x1), f(x2)) ≤ KdX(x1, x2)α. (2.41)

f is said to be α-Holder continuous over some subset O ⊂ X if f |O is α-Holder

continuous. f is said to be locally α-Holder continuous at x ∈ X if f is α-Holder

continuous over some open neighborhood of x. Lastly, the function f as a whole is

said to be locally α-Holder continuous if it is locally α-Holder continuous at every

x ∈ X.

The strength of Holder continuity depends on the exponent α: the larger the ex-

ponent the more restrictive the condition. Holder continuity is stronger than uniform

continuity for any positive α. The case α = 1 is equivalent to Lipschitz continuity.

It’s unusual to have α > 1 in the real case. Indeed, if α > 1 and X ⊆ R then any

α-Holder continuous function f must be constant on each connected component of X.

However, for X ⊆ R there is a rich variety of Holder continuous functions that have

α > 1. One such example is any linear combination of functions∑

i ai|x − xi|α for

constants xi ∈ Qp. We will make use of local Holder continuity. Indeed, we will find

instances in which the two point function is globally Holder with a smaller exponent

than it is locally Holder away from the origin.

2.7 2-adic field theory

We begin with the 2-adic field theory. The basics of p-adic field theory were given in

section 1.7. We will specialize to p = 2 and consider the bilocal action

S(φ) = −∫Q2

dxdy1

2φ(x)J(x− y)φ(y), (2.42)

38

Page 45: Real and p-Adic Physics

where

J(x) = J∗∑n∈Z

2ns(δ(x− 2n) + δ(x+ 2n)− 2δ(x)). (2.43)

As in section 2.1, the action can be diagonalized in Fourier space:

S(φ) = −1

2

∫Q2

dkφ(−k)J(k)φ(k). (2.44)

We can immediately write down the Fourier space two point function G(k) = −1/J(k).

To avoid having to carry J∗ around everywhere, we will use the convenient value

J∗ = 1/4. The Fourier space coupling may be written

J(k) = J∗∑n∈Z

2ns(χ(2nk) + χ(−2nk)− 2) (2.45)

= −∑n∈Z

2ns sin2(π2nk). (2.46)

We can restrict the infinite sum to n < ν(k)− log2 |k|2, since k2n = 0 otherwise.

An amusing connection to population dynamics can be observed at this point.

Recall the logistical map, x→ rx(1−x). If xnn∈Z is a solution to this iterated map,

then we can think of xn as (proportional to) the population of a species at generation

number n. For r = 4, a solution is xn = sin2(π2nk) where k is a real number. However,

this is not the most general solution, because it has the property xn → 0 as n→ −∞.

Consider instead xn = sin2(π2nk) where k is a 2-adic number. Then xn = 0 for all

n ≥ −v(k), but we need not have xn → 0 as n → −∞. Thus we see that the 2-adic

number k parametrizes the routes to extinction under the r = 4 logistical map, and

the 2-adic norm of k predicts the moment of extinction: n∗ = log2 |k|2. To make

the discussion simple, suppose now that k is a 2-adic integer, so that extinction has

occurred by the time n = 0. Further suppose that each generation leaves an imprint

on its environment proportional to xn, and that this imprint dissipates over time with

a half life of 1/s generations. So the environmental imprint at time 0 of generation n

39

Page 46: Real and p-Adic Physics

(with n < 0 since extinction occurs no later than time 0) is In = α2nsxn, where α is

the constant of proportionality. Then I = −αJ(k) as computed in (2.45) is the total

environmental imprint of the species, summed across all generations and measured

at time 0.

Since n reaches arbitrarily negative numbers, the coupling weights 2ns will give a

well defined J(k) if and only if s > 0. This is encouraging since s > 0 is the regime

in which we expect 2-adic continuity to emerge. Given s > 0 we can thus write

J(k) = −|k|s2Ψ(k) where k = |k|2k is a 2-adic unit and

Ψ(k) =∑m≥1

2−ms sin2(π2−mk). (2.47)

For clarity, we have used the following change of summation index: m = −(n+v(k)).

We will now show that Ψ is bounded and s-Holder continuous over the 2 adic units

U2. First boundedness:

|Ψ(k)| ≤∑m≥1

2−ms (2.48)

= −ζ2(−s). (2.49)

In fact, we can obtain a stronger bound from below. Note that since it is a 2-adic

unit, k = 1 + k12 + · · · where k1 = 0 or 1. Therefore the first two terms of the

sum (2.47) are non-vanishing and give the bound 2−s + 2−2s−1 ≤ Ψ(k). To show

s-Holder continuity note that fractional parts of 2−mk1 and 2−mk2 are exactly when

m ≤ ν(k1 − k2). We therefore have:

|Ψ(k1)−Ψ(k2)| ≤ |k1 − k2|s∞∑j=0

2−js| sin2(π2−j k1)− sin2(π2−j k2)| (2.50)

≤ Ks|k1 − k2|s, (2.51)

40

Page 47: Real and p-Adic Physics

where Ks = −2ζ2(−s). Since

|1/Ψ(k1)− 1/Ψ(k2)| = |Ψ(k1)|−1|Ψ(k2)|−1|Ψ(k1)−Ψ(k2)|, (2.52)

one can show using the fact that Ψ is bounded below by 2−s+2−2s−1 > 0 that 1/Ψ(k)

is s-Holder continuous. Therefore given k 6= 0, the momentum space Green’s function

G(k) =1

|k|s2Ψ(k)(2.53)

is the product of a locally constant function and an s-Holder continuous function.

Hence G(k) is locally s-Holder continuous away from the origin.

Let’s turn our attention to the position space two-point function G(x). As in

analysis over R, we expect that if G(k) is integrable at large |k| that G(x) will be

continuous. We therefore focus on the regime s > 1 for which G ∈ L1(Qp). We run

into an IR divergence if we try to calculate G(x) outright, therefore we introduce an

IR regulator by integrating over the region |k| ≥ |kIR| = 2−νIR . We find

G(x) = ζ2(s− 1)|kIR|1−s2

∫U2

dk

Ψ(k)+ |x|s−1

2 g(x), (2.54)

where

g(x) = ζ2(s− 1)

∫U2

dk

Ψ(k)+∞∑n=1

2(1−s)n∫U2

dkχ(2−nkx)

Ψ(k). (2.55)

Dropping the divergent term in (2.54) is equivalent to working with

G(x) =

∫Q2

χ(kx)− 1

|k|s2Ψ(k)= |x|s−1

2 g(x). (2.56)

We now derive a local (2s−1)-Holder bound for g(x). Note that by the same logic we

used to derive the Holder bound for Ψ(k), when calculating the difference g(x1)−g(x2)

41

Page 48: Real and p-Adic Physics

we can restrict the sum to n > ν(x1 − x2). Therefore:

g(x1)− g(x2) =∑

n>ν(x1−x2)

2(1−s)n∫U2

dk

(χ(2−nkx1)

Ψ(k)− χ(2−nkx2)

Ψ(k)

)(2.57)

=∑

n>ν(x1−x2)

2(1−s)n∫U2

dq χ(2−nq)

(1

Ψ(q/x1)− 1

Ψ(q/x2)

)(2.58)

In the second line we’ve used changes of variable q = kxi. Note that we don’t pick

up any Jacobian factor since |xi|2 = 1. Also note that |1/x1 − 1/x2|2 = |x1 − x2|2.

Now we use the fact that Ψ is s-Holder together with |q|2 = 1:

|g(x1)− g(x2)| ≤ 2(1−s)ν(x1−x2)

∞∑n=1

2(1−s)n∫U2

dq

∣∣∣∣ 1

x1

− 1

x2

∣∣∣∣s2

(2.59)

= Ms|x1 − x2|2s−12 , (2.60)

where Ms = −ζ2(1−s)/ζ2(1). Now G(x) = |x|s−12 g(x) where |x|s−1

2 is locally constant

and globally s− 1-Holder continuous and g(x) is (2s− 1)-Holder continuous. We can

therefore conclude that G(x) is globally Holder continuous with exponent s− 1 and

locally Holder continuous with exponent 2s− 1.

2.8 Archimedean field theory

We now carry out the analogous analysis in real field theory. We have the action

S = −1

2

∫Rφ(x)J(x− y)φ(x) = −1

2

∫Rφ(−k)J(k)φ(k), (2.61)

where we define J(x) using the same formula as in (2.43). Our Fourier transform

convention is f(x) =∫R exp(2πikx)f(k). The two point function satisfies G(k) =

42

Page 49: Real and p-Adic Physics

−1/J(k). Again using J∗ = 1/4 we have

J(k) = −∑n∈Z

2ns sin2(π2nk). (2.62)

For s ≤ −2, the sum diverges as n → −∞. But looking back at the definition

(2.43), this means that the coupling is concentrated at x = 0. Indeed, if we regulate

the sum by imposing a cutoff and re-scale J∗ while removing it, then we can attain

J(x) → −δ′′(x). The theory becomes local. If s ≥ 0 - the region of s that Holder

bounds were proved on the 2-adic side - then the sum diverges as n→∞. Thus s ≥ 0

is a region where the coupling is dominated by arbitrarily long-ranged interactions.

We therefore restrict our attention to the regime −2 < s < 0 where the sum

(2.62) converges and the theory is not perfectly local. We can ask what properties

the function J(k) satisfies analogous to the ones found for Ψ(k) in the 2-adic case. In

fact, we claim

1. −J(k) ≈ |k|−s∞ , meaning that there exist positive constants K1 and K2, inde-

pendent of k, such that K1|k|−s∞ < −J < K2|k|−s∞ for all k ∈ R\0.

2. For −1 < s < 0, J(k) is globally Holder continuous with exponent −s.

3. For −2 < s < −1, the derivative J ′(k) = dJ(k)/dk is globally Holder continuous

with exponent −s − 1. (Note that J(k) itself cannot have Holder continuity

exponent greater than 1 without being constant. So the derivative condition

we claim here is the best that can be expected.) It follows that J(k) is globally

1-Holder continuous on any bounded domain.

To arrive at the estimate −J(k) ≈ |k|−s∞ , we define

nk ≡ − log2(π|k|∞) . (2.63)

43

Page 50: Real and p-Adic Physics

Then we have

−J(k) =∑n<nk

2ns sin2(π2nk) +∑n≥nk

2ns sin2(π2nk) ≈∑n<nk

2ns(π2nk)2 +∑n≥nk

2ns

≈ (π|k|∞)−s [−ζ2(−s− 2) + ζ2(−s)] .(2.64)

To derive the Holder condition on J(k) for −1 < s < 0, set δ = |k1 − k2|∞ and note

that

| sin2(π2nk1)− sin2(π2nk2)|∞ ≤ min1, π2nδ . (2.65)

Defining

nδ = − log2(πδ) , (2.66)

we see that

|J(k1)− J(k2)|∞ ≤∑n∈Z

2ns min1, π2nδ =∑n<nδ

2n(s+1)πδ +∑n≥nδ

2ns

≈ (πδ)−s [−ζ2(−s− 1) + ζ2(−s)] ,(2.67)

where again ≈ means equality to within fixed multiplicative factors, independent in

this case of δ. The last expression in (2.67) is the desired Holder bound, valid when

−1 < s < 0. If instead −2 < s < −1, then we may calculate

J ′(k) = −π∑n∈Z

2n(s+1) sin(π2n+1k) . (2.68)

By the same method as in (2.67) we arrive at the Holder continuity condition for

J ′(k) with exponent −s− 1.

By combining the property −J(k) ≈ |k|−s∞ with the Holder bounds, we see that

G(k) is locally Holder with exponent −s for −1 < s < 0. Also, G′(k) is locally Holder

away from k = 0 with exponent −s − 1 for −2 < s < −1, implying that G(k) is

locally 1-Holder away from k = 0.

44

Page 51: Real and p-Adic Physics

Now let’s investigate smoothness of the Green’s function in position space. We

naively define

G(x) =

∫Rdk

e2πikx

−J(k). (2.69)

As in the 2-adic case, our intuitive understanding is that G(x) will be continuous

everywhere iff G(k) is integrable at large k, which is the case iff s < −1. Because

−J(k) ≈ |k|s∞, the UV-integrable regime is −2 < s < −1, and in this regime the

integral (2.69) has an IR divergence. Again as in the 2-adic case, the infrared diver-

gence results in an overall additive constant in G. It does not matter much how this

constant is removed; one option is to alter (2.69) to

G(x) ≡∫Rdke2πikx − 1

(−J(k)). (2.70)

For the purposes of a Holder continuity condition we must estimate

G(x1)−G(x2) =

∫R

dk

(−J(k))(e2πikx1 − e2πikx2) . (2.71)

Setting δ = |x1 − x2|∞, we have

|G(x1)−G(x2)|∞ ≤∫R

dk

(−J(k))min2, 2π|k|∞δ

= 2π

∫|k|<1/δ

dk

(−J(k))|k|∞δ + 2

∫|k|>1/δ

dk

(−J(k))

≈∫ 1/δ

0

dk ks+1δ +

∫ ∞1/δ

ks

≈ δ−s−1

[1

s+ 2− 1

s+ 1

].

(2.72)

In short, for −2 < s < −1, we have obtained a global Holder bound with exponent

−s− 1.

45

Page 52: Real and p-Adic Physics

log maxG˜κsparse

/G˜κ2-adic

log minG˜κsparse

/G˜κ2-adic

0.5 1.0 1.5 2.0 s

-0.5

0.5

1.0

N

6

7

8

9

10

11

12

13

14

15

16

log maxG˜κsparse

/G˜κ2-adic

log minG˜κsparse

/G˜κ2-adic

6 8 10 12 14 16 N

-0.5

0.5

1.0

1.5

s

0.1

0.2

0.3

0.4

0.5

Figure 2.3: Left: Optimal values of the constants K1 and K2 appearing in (2.73)

as functions of s for fixed N .

Right: Optimal values of the constants K1 and K2 as function of

N for fixed s. The expectation is that provided s > 0, K1 and K2

asymptote to constants at sufficiently large N .

2.9 Numerical evidence: 2-adic Approximation of

sparse results

We begin our numerical discussion by considering the question of how well the two-

point function of the 2-adic model approximates the two-point function of the sparse

model. From the results in section (2.7), we expect to find s dependent constants

K1(s), K2(s) > 0 such that

K1(s) <Gsparsek

G2−adick

< K2(s) (2.73)

in the region s > 0. Numerical evidence for these bounds is show in figure 2.3, where

the optimal values of Ki(s) have been computed for various values of s and N . As

s → 0, it is not apparent whether the K2(s) remains bounded as N → ∞. The

algorithm used to find Ki(s) is O(N2N), so barring the discovery of a better way

to calculate K2(s), we are limited by our ability to calculate using sufficiently high

values of N . Based on empirical analysis on the curves on the left side of figure 2.3,

we find Ki(s) ≈ 1 + 2−2sκi(s) where κi(s) must vary slowly with s. For reference, the

46

Page 53: Real and p-Adic Physics

steps used to compute Ki(s) for given values of s and N are listed below:

1. Compute Gsparsek and G2−adic using the methods of sections 2.5 and 2.3.

2. Normalize each Green’s function in position space by adjusting J∗ to satisfy

G0 = L−1/2

L−1∑k=1

Gk = 1. (2.74)

3. Compute K1(s) and K2(s) via

K1(s) = mink 6=0

Gsparsek

G2−adick

and K1(s) = maxk 6=0

Gsparsek

G2−adick

. (2.75)

2.10 Numerical evidence: Holder bounds in mo-

mentum space

Next we would like to numerically investigate the local Holder continuity bounds that

we found in sections 2.7 and 2.8. We start with the momentum space bounds, which

we summarize here:

For 0 < s, G : Q2 → R is locally s-Holder continuous away from the origin.

For −1 < s < 0, G : R → R is locally (−s)-Holder continuous away from the

origin.

For s < −1, G : R → R is locally 1-Holder continuous (i.e. locally Lipschitz)

away from the origin.

After normalizing Gsparse0 = 1 and G2−adic

0 = 1, we expect that Gsparsek /G2−adic

k to obey

a local s-Holder bound. It’s helpful to write the condition for a local Holder bound

47

Page 54: Real and p-Adic Physics

explicitly for each pair of 2-adic nearest neighbors (k, k + L/2):

∣∣∣∣∣ Gsparsek

G2−adick

−Gsparsek+L/2

G2−adick+L/2

∣∣∣∣∣∞

≤ K2−adic(s)2−α2−adic(N,s)(N−1). (2.76)

For the moment we are being agnostic with regard to the value of α2−adic. Our goal

given the data is to approximate α2−adic(N, s) and compare it with the theoretical

result α2−adic = s. We expect the quantity

A2−adic(N, s) = log2 maxk odd

∣∣∣∣∣ Gsparsek

G2−adick

−Gsparsek+L/2

G2−adick+L/2

∣∣∣∣∣∞

(2.77)

to be approximately linear in N :

A2−adic(N, s) ≈ log2 K2−adic(s)− α2−adic(N, s)(N − 1). (2.78)

We now assume that α2−adic(N, s) is convergent as N → ∞, so that for sufficiently

large N we have estimate

α2−adic(N, s) ≈ −A2−adic(N, s) + A2−adic(N − 1, s). (2.79)

The same numerical analysis can be carried out on the Archimedean side, making only

a few adjustments. First, our local Holder bound will now be sensitive to sequential

nearest neighbors (k, k + 1). Second, we now expect Gsparsek /Gpower

k to obey a local

Holder bound. This is because Gpowerk is smooth away from the origin so forming

the ratio Gsparsek /Gpower

k doesn’t affect the local Holder exponent of Gsparsek . Moreover,

dividing by Gpowerk cancels out part of the trend of Gsparse

k to grow larger near k = 0 =

L, which makes it easier to accurately distill the local Holder exponent with a finite

48

Page 55: Real and p-Adic Physics

4 6 8 10 12 N

-10

-5

5

A˜2-adic s

-0.75

-0.50

-0.25

0

0.25

0.50

0.75

4 6 8 10 12 14 16 N

-10

-5

5

10

A˜power s

-0.75

-0.50

-0.25

0

0.25

0.50

0.75

Figure 2.4: Left: 2-adic smoothness in momentum space. The dots are evalua-

tions of A2−adic(N, s) in (2.77), and the lines are plots of the linear

trajectories indicated in (2.78), with K(s) chosen so that the line goes

through the last data point.

Right: Archimedean smoothness in momentum space. The dots are

evaluations of Apower(N, s) in (2.80), and the lines are plots of the

linear trajectories indicated in (2.81), with K(s) chosen so that the

line goes through the last data point.

sample of points. We therefore use

Apower(N, s) = log2 maxL4<k< 3L

4

∣∣∣∣∣Gsparsek

Gpowerk

−Gsparsek+1

Gpowerk+1

∣∣∣∣∣∞

(2.80)

instead of (2.77). We expect this to follow linear trajectories:

Apower(N, s) ≈ log2 Kpower(s)− αpower(N, s)(N − 1). (2.81)

Our estimate for the exponent is

αpower(N, s) ≈ −Apower(N, s) + Apower(N − 1, s). (2.82)

In figure 2.4, the linear trajectories (2.78) and (2.81) are shown for a variety of values

of s. Interestingly, the trajectories persist in regions for which the Holder exponent,

which can be roughly identified with the slope of the lines, becomes negative.

49

Page 56: Real and p-Adic Physics

2.11 Numerical evidence: Holder bounds in posi-

tion space

We can carry out the exact same analysis in position space. Specifically, given the

quantities

A2−adic(N, s) = log2 maxh odd

∣∣∣∣∣ Gsparseh

G2−adich

−Gsparseh+L/2

G2−adich+L/2

∣∣∣∣∣∞

(2.83)

Apower(N, s) = log2 maxL4<h< 3L

4

∣∣∣∣Gsparseh

Gpowerh

−Gsparseh+1

Gpowerh+1

∣∣∣∣∞, (2.84)

we have the large N estimates

α2−adic(N, s) = −A2−adic(N, s) + A2−adic(N − 1, s) (2.85)

αpower(N, s) = −Apower(N, s) + Apower(N − 1, s). (2.86)

We find good evidence that the estimates α2−adic(N, s), αpower(N, s) for the local

momentum space exponents and the estimates α2−adic(N, s), αpower(N, s) for the local

position space exponents all have finite large N limits. Our numerical results are

shown in figure 2.5 and are well described by the following piecewise linear functions:

For s < −1, αpower = αpower = 1.

For −1 < s < 0, αpower = −s and αpower = −2s− 1.

For 0 < s, α2−adic = s and α2−adic = 2s− 1.

The following caveats apply to these estimates:

For |s| > 1 it becomes difficult to obtain high quality numerical results, partic-

ularly for αpower and αpower. In this regime, the Green’s functions become very

smooth and one has to compute very small differences with high accuracy. Even

50

Page 57: Real and p-Adic Physics

2-adicfield theory

real field theory

• α2-adic

• αpower

-2 -1 1 2s

-1

1

2

3

α

2-adicfieldtheory

real field theory

• α2-adic

• αpower

-2 -1 1 2s

-0.5

0.5

1.0

1.5

2.0

α

Figure 2.5: α versus s and α versus s in the 2-adic and Archimedean settings.

Field theory bounds derived in section 2.6 are shown in dashed black

and dashed blue. Dotted black and dotted blue show the naive ex-

trapolations of these bounds to negative α and α. Red and green

dots are numerical evaluations of α and α as defined in sections 2.10

and 2.11, respectively, with N = 20. Solid red and green lines show

the obvious piecewise linear trends which approximately match the

numerical evaluations. Open circles denote evaluations in which we

restricted 7L16 ≤ h <

9L16 ; otherwise we use half the available points as

explained in the main text. For s ≤ −2, convergence of the sparse

model to the nearest neighbor model implies that α = α = 1, but our

numerical scheme for picking out α and α becomes less reliable in this

region due to difficulty normalizing Gsparse and Gpower in a mutually

consistent way.

51

Page 58: Real and p-Adic Physics

without the numerical challenges, it becomes a difficult to distinguish between

rapid smooth variation and the non-smooth behavior that characterizes Holder

exponents.

Numerical calculations diverge a bit from our piecewise linear description at

s = 0 and s = −1. This isn’t too surprising since the pre-factors K(s) in

the Holder inequalities diverge at these values. See [references to divergences].

More extensive numerical investigations are needed to establish reliable results

at these values of s. It’s possible that we need logarithmic corrections to the

Holder conditions at these values, or that the exponents are only approximately

described by piecewise linear functions.

2.12 The liminal region −1/2 < s < 1/2

Of special interest is the regime −1/2 < s < 1/2. In momentum space, our numerics

are consistent with there being a single exponent on the 2-adic side, α2−adic = s, which

describes both the global Holder continuity condition over all k and the local continu-

ity at each possible value of k. In other words, as far as we can tell, the function G(k)

is equally ragged everywhere. A similarly uniform story applies on the Archimedean

side, with αpower = −s. Numerical results are fully in accord with expectations from

field theory, where we were able to compute α2−adic and αpower analytically. The up-

shot is that the transition from Archimedean to ultrametric continuity happens rather

simply, with ordinary continuity failing just as 2-adic continuity emerges: i.e. αpower

becomes negative just as α2−adic becomes positive, at s = 0.

The numerically estimated position space exponents behave differently. Both the

2-adic and Archimedean exponents become negative in the region s ∈ (−1/2, 1/2),

indicated a loss of local Holder continuity. Indeed, the piecewise linear description

found in section 2.11 suggests that Gsparseh loses Archimedean continuity at s = −1/2

52

Page 59: Real and p-Adic Physics

s = 0.3

0.6 0.7 0.8 0.9 1.0(h)/L

-1.0

-0.5

0.5

1.0

1.5

Ghsparse

/Gh2-adic

s = 0.3

0.6 0.7 0.8 0.9 1.0(κ)/L

0.5

1.0

1.5

G˜κsparse

/G˜κ2-adic

s = -0.3

0.6 0.7 0.8 0.9 1.0(h)/L

-60

-40

-20

20

Ghsparse

/Gh2-adic

s = -0.3

0.6 0.7 0.8 0.9 1.0(κ)/L

1

2

3

4

5

6

7

G˜κsparse

/G˜κ2-adic

Figure 2.6: Plots of Gsparseh /G2−adic

h and Gsparseκ /G2−adic

κ over the Monna map of

the odd integers. As s becomes more positive, the numerical data is

closer to a 2-adically continuous curve when N is large. Blue points

are for N = 6, while the red curves are for N = 10.

53

Page 60: Real and p-Adic Physics

s = 0.3

0.3 0.4 0.5 0.6 0.7h/L

-1500

-1000

-500

Ghsparse

/Ghpower

s = 0.3

0.3 0.4 0.5 0.6 0.7κ/L

2

4

6

8

10

12

G˜κsparse

/G˜κpower

s = -0.3

0.3 0.4 0.5 0.6 0.7h/L

-8

-6

-4

-2

2

4

Ghsparse

/Ghpower

s = -0.3

0.3 0.4 0.5 0.6 0.7κ/L

0.5

1.0

1.5

2.0

2.5

3.0

G˜κsparse

/G˜κpower

Figure 2.7: Plots of Gsparseh /Gpower

h and Gsparseκ /Gpower

κ over the middle half of

points. As s becomes more negative, the numerical data is closer to

a continuous curve when N is large. Blue points are for N = 6, while

the red curves are for N = 10.

54

Page 61: Real and p-Adic Physics

and does not gain 2-adic continuity until s = +1/2.

Based on these figures and related studies, the scenario we regard as most likely is

that for −1/2 < s < 0, the continuum limit of Gpowerh defines an absolutely continuous

measure, G(x)dx, with respect to ordinary Lebesgue measure dx, but for s > 0 any

such continuum limit would necessarily have a singular term in its Radon-Nikodym

decomposition. Similarly, we suggest that for 0 < s < 1/2, the continuum limit of

G2−adich defines an absolutely continuous measure with respect to the standard Haar

measure on Q2 while for s < 0 any such continuum limit would have a singular term

(with respect to the Haar measure on Q2) in its Radon-Nikodym decomposition. We

find support for the claim of absolutely continuous measures in the above-mentioned

regimes when we study the scaling of the height of the spikes in figures 2.6 and 2.7

as a function of N : the weight of each spike (meaning the integral over a small region

including the spike) distinctly appears to tend to zero with increasing N . When

singular terms in Radon-Nikodym decompositions do exist, we conjecture that they

have as their support sets which are dense in position space.

One way in which singular terms in Radon-Nikodym decompositions could arise

is for the continuum limit G(x) to include delta functions. Inspection of figure 2.6 is

consistent with there being a dense set of delta function spikes in G(x) as a function

of 2-adic x when s = −0.3, but none when s = 0.3. Similarly, figure 2.7 is consistent

with there being a dense set of delta function spikes in G(x) as a function of real x

when s = 0.3, but none with s = −0.3. The discerning reader may note, however,

that the spikes on the Archimedean side are stronger at s = 0.3 than the ones on the

2-adic side at s = −0.3. This asymmetry manifests itself in the scaling of the height of

these spikes with N , for the weight of each spike grows with N on the Archimedean

side for s = 0.3, but may be trending very slowly toward 0 on the 2-adic side at

s = −0.3. A related effect appears in figure 2.5: α2−adic ≈ −1 for s < 0, while

αpower ≈ −1− s for s > 0.

55

Page 62: Real and p-Adic Physics

Inspection of figures 2.6 and 2.7 reveals some self-similarity in the Green’s func-

tions both before and after the Monna map is applied. We have not investigated

this fractal behavior in detail; however, we note that similar behavior has been found

independently in band structure calculations in connection with proposed cold atom

experiments [47].

2.13 Outlook

For decades, p-adic numbers have been considered as an alternative to real numbers

as a notion of continuum which could underlie fundamental physics at a microscopic

scale; see for example [35]. This chapter has shown how the large system size limit

of an underlying discrete system naturally interpolates between a one-dimensional

Archimedean continuum and a 2-adic continuum as we vary a spectral exponent. By

focusing on a free field example, we are able to solve the model through essentially

trivial Fourier space manipulations. The correlators of the theories we study are all

determined in terms of the two-point function through application of Wick’s theorem.

The two-point function is continuous in an Archimedean sense when s is sufficiently

negative, and in a 2-adic sense when s is sufficiently positive. The transition from

these two incompatible notions of continuity can be precisely characterized in terms of

Holder exponents characterizing the smoothness of the two-point function. We have

found the dependence of these exponents on s through a combination of analytical

field theory arguments and numerics on finite but large systems.

Quite a wide range of generalizations of our basic construction can be contem-

plated:

1. We can generalize to primes p > 2. One significant subtlety arises when doing so,

namely the structure within Z/pZ of sparse couplings. The simplest alternative

is for spin 0 to couple to spins ±θpn with a strength pns, where θ runs over all

56

Page 63: Real and p-Adic Physics

elements of 1, 2, 3, . . . , p−1. This coupling pattern is featureless within Z/pZ

because it treats all values of θ the same. One can however contemplate other

possibilities. For example, if p = 5, an interesting alternative is to introduce

couplings only for θ = 1 and θ = 4 (the quadratic residues). More generally,

one could expand the dependence of couplings on θ in a sum of multiplicative

characters over Z/pZ.

2. This chapter has focused entirely on bosonic spins φi, but there is no reason

not to consider fermions ci instead. Then the coupling matrix Jij would have to

be anti-symmetric, and likewise the two-point Green’s function would be odd.

Within this framework one could consider a variety of sparse coupling patterns.

3. Higher-dimensional examples are not hard to come by. Consider real bosonic

spins φ~ı labeled by a two-dimensional vector ~ı = i11 + i22, where i1 and i2 are

in Z/3NZ. Suppose we establish a coupling matrix J~ı~ = J~ı−~ where

J~h =

3minn1,n2s if h1 = ±3n1 and h2 = ±3n2

3n1s if h1 = 0 and h2 = ±3n2

3n2s if h2 = 0 and h1 = ±3n1 ,

(2.87)

with all other entries vanishing except J0, whose value we choose in order to

have the Fourier coefficient J~κ vanish when ~κ = 0. Then for sufficiently negative

s we have effectively a nearest neighbor model which approximates the massless

field theory S =∫d2x 1

2(∇φ)2. For s sufficiently positive, one obtains instead a

continuum theory over Z3×Z3, which can be understood as the ring of integers

in the unramified quadratic extension of Q3, see section 1.5.

All the examples above remain within the paradigm of free field theory. Still easy to

formulate, but obviously much harder to solve, are interacting theories with sparse

57

Page 64: Real and p-Adic Physics

s = 0 112

23

−1 −12

−23

φ4re

leva

ntφ

6re

leva

nt

φ 4relevant

φ 6relevant No relevant deformations

ULTRAMETRICARCHIMEDEAN

Gaussian Gaussians

Figure 2.8: Conjectured pattern of fixed points of the renormalization group for

interacting field theories of a single bosonic scalar field with φ→ −φsymmetry.

couplings. For example, one could start with any of the models introduced in sec-

tions 2.2-2.5 and add a term∑

i V (φi) to the Hamiltonian describing arbitrary on-site

interactions. To get some first hints of what to expect these interactions to do, recall

in 2-adic field theory that G(x) ≈ |x|s−12 at small x. Comparing this to the standard

expectation G(x) ≈ |x|2∆φ

2 , we arrive at ∆φ = (1− s)/2 as the ultraviolet dimension

of φ. When describing perturbations of the Gaussian theory, we can use normal UV

power counting: [φn] = n∆φ. Thus φn is relevant when s > 1− 2/n. If we impose Z2

symmetry, φ→ −φ, then in the region s < 1/2, the Gaussian theory has no relevant

local perturbations, but as s increases from 1/2 to 1, first φ4 and then higher powers

of φ2 become relevant. It is reasonable to expect some analog of Wilson-Fisher fixed

points to appear. Possibly as s → 1 these fixed points extrapolate to analogs of

minimal models. An analogous story presumably applies on the Archimedean side to

power-law field theories controlled by s in the range (−1, 0), with G(x) ≈ |x|−s−1∞ and

therefore ∆φ = (1 + s)/2. See figure 2.8.

The sparse coupling theories are sufficiently similar to 2-adic field theories for

s > 0 and to power-law field theories for s < 0 that it is reasonable to conjecture that

the same pattern of renormalization group fixed points arises. However, this line of

reasoning is incomplete. In particular, one would require an understanding of how

58

Page 65: Real and p-Adic Physics

the improved local Holder smoothness arises and how it might affect renormalization

group flows. A Monte Carlo study of the phases of the sparsely coupling Ising model

might help shed light on the renormalization group flows available to interacting

models, particularly in the range −2/3 < s < 2/3 where no powers of φ higher than

φ4 are relevant—according at least to naive power counting.

59

Page 66: Real and p-Adic Physics

3

Higher Melonic Field Theories

This chapter is based on [3] coauthored with Steven S. Gubser, Christian Jepsen, and

Ziming Ji. We thank Igor Klebanov for useful discussions and suggestions.

In the last chapter, we developed a sparsely coupled statistical model that interpo-

lated between Archimedean and p-adic continuity. In this chapter we will explore the

diversity that arises when we consider field theory over R and Qp in tandem. In this

context, we will classify a large collection of melonic theories with q-fold interactions,

demonstrating that the interaction vertices exhibit a range of symmetries of the form

Zn2 for some n ∈ 0, 1, . . .. The number of different theories (over both Archimedean

and non-Archimedean spaces) proliferates quickly as q increases above 8 and is related

to the problem of counting one-factorizations of complete graphs. The symmetries

of the interaction vertex lead to an effective interaction strength that enters into the

Schwinger-Dyson equation for the two-point function as well as the kernel used for

constructing higher-point functions.

Melonic theories [48, 49, 50, 51, 52] are an interesting class of quantum field the-

ories whose essential property is that in an appropriate large N limit, the dominant

Feynman diagrams can be generated by iterating on the replacement of a propagator

by a melonic insertion, as shown in figure 3.1 for a melonic version of scalar φ4 theory.

Melonic theories are interesting for two related reasons: 1) The melonic large N limit

60

Page 67: Real and p-Adic Physics

is relatively tractable because its Green’s functions can be determined through func-

tional techniques including Schwinger-Dyson equations; and 2) The simplest Green’s

functions are the same as for the Sachdev-Ye-Kitaev (SYK) model [53, 54], widely

studied because of its proposed relationship to AdS2.

It is well recognized that melonic theories exist not just with quartic interactions,

but also higher order interactions [55, 56, 57, 58]. The aim of this chapter is to take

some steps toward an understanding of what sorts of higher order interactions are

possible. We focus on the Klebanov-Tarnopolsky model [52] where the interaction is of

order q—meaning that q propagators meet at each interaction vertex. Inquiries in this

direction were initiated in [59]. Starting at q = 8, the number of different interaction

vertices proliferates quickly. Some of them are symmetrical under a subgroup of

permutations of the propagators leading into them; others have no such symmetry.

Interaction vertices do not mix with one another in the leading melonic limit: The

diagrams that would permit this are subleading. We are therefore content to restrict

to theories with only one type of interaction vertex—and each different interaction

vertex gives a different theory. An interaction vertex of order q can be constructed

starting from a coloring of the complete graph with q vertices such that each of

q − 1 colors is incident once on each vertex. The questions addressed in this chapter

regarding these interaction vertices are:

1. What are the possible symmetry groups of these interaction vertices?

2. How does the number of distinct interaction vertices grow with q?

We claim in sections 3.2 and 3.3 to completely settle question #1: The possible

3 1

5 7

2

6

4 0

3 1

5 7

2

6

4 0

S4

φ4

tr Φ4

Z4

Z2 Z2

ψa0a1a2ψa0b1b2ψb0a1b2ψb0b1a2

1

Figure 3.1: A melonic insertion.

61

Page 68: Real and p-Adic Physics

symmetry groups are Zn2 , where 0 ≤ n ≤ v if q = 2v and 0 ≤ n < v if q = u2v

where u is odd and larger than 1. (It is easy to show that q must be even.) This

demonstration will be constructive, in that we produce vertices with each possible

symmetry. Question #2 turns out to be difficult, and it is essentially the problem

of counting so-called one-factorizations of complete graphs, where results are gen-

erally available up to q = 14 [60, 61, 62, 63, 64, 65, 66]: results spanning over a

century! In section 3.4, we summarize how these results can be combined and mod-

estly extended to give complete results on the number of distinct interaction vertices

up to q = 14. The symmetry analysis undertaken in this chapter suggests a new

twist on the counting problem: In addition to counting all one-factorizations, one

can count one-factorizations with a given symmetry group. An explicit example of

this symmetry-constrained counting is presented in section 3.5, and some additional

conceptual points are discussed in section 3.6.

The symmetry group of the interaction vertex leads to an effective coupling that

enters into the Schwinger-Dyson equation for the two-point function and the ladder

operator used for computing the four-point function, which will be exhibited explicitly

in all cases in sections 3.7 and 3.8. We note here that we assume every interaction

vertex we construct leads to a theory with a melonic limit, and all indications are

consistent with this. This being said, one will not find a fully rigorous proof of this

claim here. Over the p-adic numbers Qp it has already been shown in [27] that there

is quite a variety of melonic theories, depending on what sign function one chooses

over Qp. For approximately half of these melonic theories over Qp, the Schwinger-

Dyson equation can be solved exactly, not just in the infrared, but at all scales, in

terms of the solution to a q-th order polynomial equation. Remarkably, as shown in

section 3.8.1, there is an adelic product formula relating the eigenvalues of the ladder

operator integral equation across real and p-adic theories.

62

Page 69: Real and p-Adic Physics

3.1 Structure of higher melonic theories

The action of the simplest Klebanov-Tarnopolsky model [52] is

S =

∫dt

(i

2ψa0a1a2∂tψ

a0a1a2 +g

4ψa0a1a2ψa0b1b2ψc0a1b2ψc0b1a2

), (3.1)

where each index, which we think of as a color index, runs independently from 1 to

N and each ψa0a1a2 is a Majorana fermion. Each index is separately O(N) invariant

and ψa0a2a3 transforms under O(N)3 as a triple tensor product of the fundamental

representation of O(N). We seek to understand the color structure of higher rank

models with higher degree interactions. We will simultaneously generalize to models

defined over p-adic numbers and to models with either O(N) or Sp(N) indices. The

models we consider will take the form

S =σψ2

∫K

dωψA(−ω)ΩAB|ω|s sgn(ω)ψB(ω) (3.2)

+ (σψ)q4g

|G|

∫K

dtΩA(0)A(1)...A(q−1) , (3.3)

where K is either R or Qp, G is the automorphism group of the interaction vertex, to

be discussed further in section 3.2. The spectral parameter s would usually be chosen

to be 1 for fermionic theories over R or 2 for bosonic theories over R, but for theories

over Qp it is more natural to let it vary continuously over positive real values. We

set σψ = −1 for fermionic theories, and σψ = +1 for bosonic theories. By sgnω we

mean a sign character, which is to say a multiplicative homomorphism of non-zero

elements of K to 1,−1. Capital indices are really groups of q− 1 lowercase indices,

each N -valued. For example, to recover (3.1) as a special case, we would set q = 4,

63

Page 70: Real and p-Adic Physics

so that A = a0a1a2; we would set ΩAB = δa0b0δa1b1δa2b2 ; we would set

ΩABCD = δa0b0δa1c1δa2d2δb1d1δb2c2δc0d0 ; (3.4)

and of course we would set K = R and s = 1. It is well recognized (see e.g. [49,

51, 52]) that the structure ΩABCD in (3.4) corresponds to a coloring of the edges of

the tetrahedron so that only three colors are used, and opposite edges have the same

color.1

When considering theories over Qp, as explained in section 1.7 and in [27], we

must allow ΩAB to be symmetric (σΩ = 1) or anti-symmetric (σΩ = −1); and we

must choose sgnω to be one of the several multiplicative sign characters over Qp,

which are in one-to-one correspondence with the quadratic extensions of Qp. To get

a real, non-vanishing kinetic term, ΩAB must be Hermitian, and we must have

σψσΩ = sgn(−1) . (3.5)

(Surprisingly, non-trivial sign characters over Qp can have either sgn(−1) = −1 or

+1.)

For q = 4, still following [27], the obvious adaptation of the Klebanov-Tarnopolsky

model to a theory over Qp is to set ΩAB = Ωa0b0Ωa1b1Ωa2b2 where

Ωab =

1N×N for σΩ = 1

σ2 ⊗ 1N2×N

2for σΩ = −1 ,

(3.6)

1It is useful to clarify here one point of terminology: We use the terms “vertex” and “edge” todescribe the inner structure of an interaction vertex like ψa0a1a2ψa0b1b2ψc0a1b2ψc0b1a2 . From thispoint of view, an interaction vertex is a graph unto itself, with q vertices when the interaction termhas q powers of ψ. A full Feynman diagram consists of propagators connecting interaction vertices,and as is familiar from earlier work including [49], the inner structure of a propagator is q−1 threadswhich flow into the edges inside an interaction vertex.

64

Page 71: Real and p-Adic Physics

and to set

ΩABCD = Ωa0b0Ωa1c1Ωa2d2Ωb1d1Ωb2c2Ωc0d0 . (3.7)

(Note that if σΩ = −1, then because of (3.6), N must be even.) If indeed Ωab is

antisymmetric, then one needs a direction on all the edges of the tetrahedral graph in

order to decide the order of the indices in each factor on the right hand side of (3.7).

However, flipping the direction on any one edge flips the sign of ΩABCD, and so can

be compensated for by changing the sign of g.

Generalizing the kinetic term to q > 4 is easy: We need only set

ΩAB =

q−2∏i=0

Ωaibi . (3.8)

Generalizing the interaction tensor turns out to be more subtle, and laying the ground-

work for finding suitable generalizations is the focus of the rest of this section.

Up to the minor issue of directedness, constructing a rank q interaction tensor

ΩA(0)A(1)...A(q−1) as a product of q(q − 1)/2 factors Ωa

(i)r a

(j)r

corresponds to a coloring

problem on the complete graph of q points (and therefore q(q−1)/2 edges), where we

use q − 1 colors (each one labeled by a value of r) and require that each of the q − 1

links incident on a given vertex (each one labeled by a value of i) must be a different

color. A special case of Baranyai’s theorem guarantees that this can always be done

provided q is even. It is impossible when q is odd. One can map the problem onto the

scheduling of a round-robin tournament, where each link is one game, each vertex is a

contestant, and each color is a round, during which each contestant plays exactly one

game. This phrasing makes it obvious that no coloring is possible for q odd, because

in a given round one must pair up all q contestants in two-person games. For q even,

65

Page 72: Real and p-Adic Physics

there is a canonical solution, which is

rij ≡ i+ jmod q − 1 if 0 ≤ i < q − 1, 0 ≤ j < q − 1, and i 6= j

ri,q−1 ≡ 2imod q − 1 for 0 ≤ i < q − 1 .

(3.9)

Here, rij is the color of the edge from vertex i to vertex j. Vertex labels i and j take

values from 0 to q − 1, while color labels r run from 0 to q − 2. The corresponding

interaction tensor is

ΩA(0)A(1)...A(q−1) =∏

0≤i<j≤q−1

Ωa

(i)rija

(j)rij

. (3.10)

In (3.10), we took care of the directedness issue by requiring i < j, which is the same

as alphabetizing the lowercase indices in (3.7).

For q = 2, 4, and 6, the canonical solution (3.9) is the only solution. Starting at

q = 8, there are multiple solutions: That is, rij can be chosen differently from (3.9)

but still consistent with the requirement that we use only q − 1 colors and have one

edge of each color coming together at each vertex. For any such rij, we can still use

(3.10) to construct the interaction tensor. We exhibit the six different solutions for

q = 8 in figure 3.2. By “different,” we mean that there is no way to relabel the colors

and/or the vertices to map any of the six rij into one another. A striking point is

that the solutions have different symmetry groups, composed of up to three factors

of Z2.2

For q = 10, there are (we claim) 396 different interaction vertices, and none of

them have any symmetry. To justify this claim, and to proceed to larger q, we need to

give a more conceptually organized presentation. We do so in the next three sections,

starting with constraints on the symmetry groups in section 3.2, continuing with an

explicit construction in section 3.3, and concluding with a summary of the problem

2By Z2 we mean the integers modulo 2, or equivalently the multiplicative group 1,−1—notthe 2-adic integers.

66

Page 73: Real and p-Adic Physics

canonical asymmteric, 960

3 1

5 7

2

6

4 0

fully symmteric, 30

3 1

5 7

2

6

4 0

asymmteric, 1680

3 1

5 7

2

6

4 0

Z2, 2520

3 1

5 7

2

6

4 0

1

(a) G = Z 32

30 one-factorizations

Order of Aut(F) is 1344

asymmteric, 1680

3 1

5 7

2

6

4 0

Z2, 2520

3 1

5 7

2

6

4 0

Z22, 420

3 1

5 7

2

6

4 0

Z22, 420

3 1

5 7

2

6

4 0

1

(b) G = Z 22

630 one-factorizations

Order of Aut(F) is 64

asymmteric, 1680

3 1

5 7

2

6

4 0

Z2, 2520

3 1

5 7

2

6

4 0

Z22, 420

3 1

5 7

2

6

4 0

Z22, 420

3 1

5 7

2

6

4 0

1

(c) G = Z 22

420 one-factorizations

Order of Aut(F) is 96

asymmteric, 1680

3 1

5 7

2

6

4 0

Z2, 2520

3 1

5 7

2

6

4 0

Z22, 420

3 1

5 7

2

6

4 0

Z22, 420

3 1

5 7

2

6

4 0

1

(d) G = Z2

2520 one-factorizations

Order of Aut(F) is 16

canonical asymmteric, 960

3 1

5 7

2

6

4 0

fully symmteric, 30

3 1

5 7

2

6

4 0

asymmteric, 1680

3 1

5 7

2

6

4 0

Z2, 2520

3 1

5 7

2

6

4 0

1

(e) G = 1

1680 one-factorizations

Order of Aut(F) is 24

canonical asymmteric, 960

3 1

5 7

2

6

4 0

fully symmteric, 30

3 1

5 7

2

6

4 0

asymmteric, 1680

3 1

5 7

2

6

4 0

Z2, 2520

3 1

5 7

2

6

4 0

1

(f) G = 1

960 one-factorizations

Order of Aut(F) is 42

Figure 3.2: The six inequivalent melonic interactions and their symmetry fac-

tors for q = 8. Each interaction can be identified with an isomor-

phism class of one-factorizations, and we also list the number of one-

factorizations in each class as well as |Aut(F)|, the number of permu-

tations in S8 that preserve the one-factorizations in a given isomor-

phism class (see section 3.4 for an explanation of this notation). The

canonical coloring is the bottom right.

67

Page 74: Real and p-Adic Physics

3 1

5 7

2

6

4 0

3 1

5 7

2

6

4 0

S4

φ4

tr Φ4

Z4

Z2 Z2

ψa0a1a2ψa0b1b2ψb0a1b2ψb0b1a2

1

3 1

5 7

2

6

4 0

3 1

5 7

2

6

4 0

S4

φ4

tr Φ4

Z4

Z2 Z2

ψa0a1a2ψa0b1b2ψb0a1b2ψb0b1a2

1

3 1

5 7

2

6

4 0

3 1

5 7

2

6

4 0

S4

φ4

tr Φ4

Z4

Z2 Z2

ψa0a1a2ψa0b1b2ψb0a1b2ψb0b1a2

1

Figure 3.3: Left: The symmetry group of φ4 theory is the symmetric group S4.

Middle: The symmetry group of tr Φ4 theory is the cyclic group Z4.

Right: The symmetry group of the quartic tensor interaction is the

Klein group Z22.

of counting distinct interaction vertices in section 3.4.

3.2 Symmetry groups of interaction vertices

In ordinary scalar field theory where the scalar φ is real-valued, the symmetry group

of a q-fold interaction vertex φq is the permutation group Sq, with order q!, because

all propagators leading into the interaction vertex are equivalent and can be per-

muted arbitrarily without changing the structure of the interaction. In a matrix field

theory based on a Hermitian N × N matrix Φ, the symmetry group of a tr Φq in-

teraction vertex is the group Zq of cyclic permutations of the propagators. In the

q = 4 Klebanov-Tarnopolsky model, the symmetry group is Z2 × Z2 (not Z4), gen-

erated by the permutations (12)(34), and (13)(24) and sometimes referred to as the

Klein group. As should be clear from figure 3.3, a permutation in the Klein group

reorders propagators leading into the interaction vertex in such a way that we get

back to exactly the same diagram that we started with. This sort of permutation is

what we will call a coloring automorphism. If we look at the inner structure of the

interaction vertex, we see that a coloring automorphism permutes the vertices (each

one corresponding to an incoming propagator) in such a way as to preserve the colors

of each edge.

68

Page 75: Real and p-Adic Physics

To be precise: When we say that rij is a coloring of the complete graph Kq with

q vertices, what we mean is that ij labels an edge (so i 6= j), and each rij is chosen

from the set of “colors” 0, 1, . . . , q − 2, with the constraints that rij = rji and that

for fixed i, rij is a bijection from the q − 1 vertices that remain after i is omitted to

the set of colors. In other words, each edge leading into a given vertex is a different

color. A coloring automorphism is defined as a map i→ π(i) such that rπ(i)π(j) = rij

for all i and j. Let the group of coloring automorphisms be G. The main purpose of

the rest of this section is to limit the possibilities for G. Then in section 3.3 we will

show that all groups G not ruled out by the arguments of this section actually can

be realized.

Our first claim is that any coloring automorphism is an involution. Denote the

coloring automorphism by π. Assume that π is not the identity, since otherwise the

claim is trivial. For some vertex i, we have j ≡ π(i) 6= i. Then rij = rπ(i)π(j) = rjπ(j),

where in the first equality we remembered that π is a coloring automorphism. From

rji = rjπ(j) we can conclude that i = π(j) because, as noted previously, the coloring

r must be a bijection, for fixed j, from vertices i 6= j to colors.

Next we remember an elementary result of group theory: Any finite group G

consisting only of involutions is isomorphic to Zn2 for some n. First let’s show that

the group is abelian. Therefore let g and h be group elements. Because g and h are

involutions, we have (gh)−1 = h−1g−1 = hg. But because gh is also an involution,

(gh)−1 = gh. So gh = hg as required. Now the fundamental theorem of finite abelian

groups tells us that G must be a direct product of cyclic subgroups of prime-power

order. Because all elements of G are involutions, any cyclic subgroup must be a copy

of Z2, and the result is proven.

Another elementary point to note is that if a coloring automorphism preserves

any vertex, then it is necessarily the trivial automorphism that maps all vertices to

themselves. To see this, suppose π(i) = i for some vertex i, and consider any other

69

Page 76: Real and p-Adic Physics

G = Z2

1 0

G = Z 22

1 0

2 3

G = Z 32

5

6

4

7

1 0

2 3

1

G = Z2

1 0

G = Z 22

1 0

2 3

G = Z 32

5

6

4

7

1 0

2 3

1

G = Z2

1 0

G = Z 22

1 0

2 3

G = Z 32

5

6

4

7

1 0

2 3

1

Figure 3.4: The first three instances of maximally symmetrical interaction ver-

tices, with q = 2v and G = Zv2. We omitted one color from the q = 8

case, namely a seventh color with edges running from each corner to

the diametrically opposite corner, for example from 0 to 6.

vertex j. We have rij = rπ(i)π(j) = riπ(j), and because r is a bijection, for fixed i, from

vertices j 6= i to colors, we can conclude π(j) = j. It follows that any permutation

that gives a coloring automorphism consists of q/2 two-cycles.

We also note that color automorphisms that have a two-cycle in common must

be identical. For suppose there are color automorphisms π and π and a vertex i such

that π(i) = π(i). Then for any j we have that rij = rπ(i)π(j) = rπ(i)π(j) but also

rij = rπ(i)π(j). Since r is a bijection, it follows that π(j) = π(j) for all j.

It helps our imagination to think of the group G = Zn2 of coloring automorphisms

as reflections through n orthogonal planes which do not pass through any vertices. If

q = 2v and n = v, this line of thinking suggests that we can produce a coloring of Kq

whose automorphism group is Zv2: See figure 3.4 for the first few instances.

With these preliminaries in hand, we now come to the main result of this section:

If q = u2v where u is odd, then the largest that G can be is Zv2 if u = 1, or Zv−12

if u > 1. The arguments in the remainder of this section do not demonstrate the

existence of interaction vertices with any particular symmetry group G; rather, they

rule out larger symmetry groups.

The proof of our main result relies on the orbit-stabilizer theorem, which we

summarize here for the purposes of a self-contained presentation. If a group G acts

70

Page 77: Real and p-Adic Physics

on a set X, then the stabilizer stab i of an element i ∈ X is the subgroup of G of

elements which preserve i. Meanwhile, the orbit orb i is the subset of X consisting of

all images of i under the action of elements of G. The theorem says

| orb i|| stab i| = |G| . (3.11)

As a first application, let X be the set of q vertices, assume that G = Zn2 . The

stabilizer of a vertex i is the trivial group: This is because any π ∈ G that maps i

to itself must also map all other vertices to themselves. Invoking (3.11) we see that

each orbit contains 2n points. The union of all vertex orbits is all of X, and so there

must be

q ≡ u2v−n (3.12)

distinct vertex orbits. Already, (3.12) shows that G cannot be larger than Zv2, because

if n > v, q is not an integer.

To finish proving the main result, all we need to do is to exclude the possibility

that G = Zv2 when u > 1. This turns out to require somewhat more subtle reasoning

than we have used so far, but the essential idea is to consider the quotient of the

interaction vertex by G and show that, as a graph with q = u vertices, it leads to

an impossible coloring problem. Clearly, we could just set v = n throughout the

following paragraphs, but we refrain from doing so because keeping n ≤ v general

allows us to see some first hints on how to actually construct graphs with any allowed

symmetry.

Let’s start with a second application of the orbit-stabilizer theorem. Consider the

set of all edges of a fixed color. For any fixed color, there are q/2 such edges. Edges

of a fixed color are permuted among themselves by coloring automorphisms, and for

each edge we have

| orb e|| stab e| = |G| . (3.13)

71

Page 78: Real and p-Adic Physics

Following a standard trick, we divide both sides of (3.13) by | stab e| and then sum

over all distinct orbits to get

q

2=∑orbits

|G|| stab e|

. (3.14)

This can be written more usefully as

u = 2n−v+1∑orbits

1

| stab e|. (3.15)

To get proper mileage out of (3.15), we need some knowledge of the values of | stab e|—

which need not be unity! Consider however the following division of edges (still of

any fixed color) into two classes. There are edges which join two vertices which are

in the same vertex orbit; we will call these “internal” edges. All other edges we

will refer to as “external” edges. We observe that stab e is the trivial group for any

external edge, because if it weren’t, then there would be some coloring automorphism

that exchanges the edge’s two ends, and that would make the edge internal. So

| stab e| = 1 for external edges.

Next we want a count of external edges of a fixed color. It’s easier to start by

considering external edges of any color, i.e. the disjoint union of external edges of

each fixed color. There are(q2

)·22n such edges, because to specify one we must choose

a pair of vertex orbits, and then from each of those two vertex orbits we must choose

one vertex. Because the stabilizer of an external edge is trivial, its orbit must have

2n elements. So the count of external edge orbits of any color is

Ne =

(q

2

)· 2n = q

(u2v−n−1 − 1

2

). (3.16)

Suppose now 0 ≤ n < v. Then 2v−n−1 is a whole number, and because we have only

q− 1 colors to work with, there must be at least one color—call it red—with at least

72

Page 79: Real and p-Adic Physics

u2v−n−1 external edge orbits. Restricting to red edges only, we recall that | stab e| = 1

for each external edge, and so at least u2v−n−1 terms in the sum on the right hand side

of (3.15) must be unity. Comparing to (3.15), we see that—for red edges—the sum

over orbits works out perfectly with only external edges, implying that there can’t be

any red internal edges. This is informative and useful for constructing examples.

Now suppose n = v. If also u = 1, then from (3.16) we see that Ne = 0: There are

no external edges at all! This makes sense because there is only one vertex orbit, and

indeed the cubical vertex illustrated in figure 3.4 shows that it is entirely consistent

to have n = v and u = 1. Where things get dangerous is if n = v and u > 1. Then,

from (3.16), Ne = q(u − 1)/2. Since we have only q − 1 colors to work with, there

must be at least one color—again call it red—with at least (u + 1)/2 external edge

orbits. Considering only red edges the sum on the right hand side of (3.15) restricted

to external edges gives u + 1. This is disastrous, because adding in the contribution

of internal edges (if any) results in the absurd inequality u ≥ u+ 1. Another way to

put it is that if n = v and u is odd and greater than 1, then we can’t color even the

external edges consistently with only q − 1 colors—let alone the internal edges. See

figure 3.5.

3.3 Construction of interaction vertices

Since for any q we can construct an interaction vertex with the canonical ordering,

which has no coloring automorphisms, what remains to be shown is that for q equal

to twice times an even number, it is always possible to construct an interaction vertex

with any allowed non-trivial symmetry group. Figure 3.2 explicitly shows that this

is true for q = 8. We now prove inductively that it is also true for q > 8.

Let q′ = q/2 and assume that rij is a coloring of Kq′ with symmetry group G′

(which may be the trivial group). Split the q vertices into two sets of q′ vertices;

73

Page 80: Real and p-Adic Physics

3

4

1

0

2

5

q = 6

Z2

25 34

01

q = 3

1

Figure 3.5: A partial coloring for q = 6 which illustrates what goes wrong when

one demands too much symmetry. Here we propose Z2 symmetry,

which according to our general result is too much. On the left we

show a partial coloring with Z2 symmetry. On the right, we show

the same coloring modded out by the Z2 symmetry. The dots on the

right are vertex orbits, and the double lines of each color are edge

orbits. The problem is that we need two edge orbits between each

pair of vertex orbits, and all of them (in this simple case) must have

different colors. That means we need six colors just for the external

edges, and we only have five to work with—leaving us without a

second color to use in connecting the 01 and 34 orbits. Internal edges

would run within a vertex orbit, for example from 0 to 1.

74

Page 81: Real and p-Adic Physics

342

051

9

108

117

6

342

051

9

108

117

6

342

051

9

108

117

6

1

(a)

342

051

9

108

117

6

342

051

9

108

117

6

342

051

9

108

117

6

1

(b)

342

051

9

108

117

6

342

051

9

108

117

6

342

051

9

108

117

6

1

(c)

Figure 3.6: Constructing an interaction vertex with q = 12 and G = Z2 in three

steps. (a) Separate the vertices into two groups of 6, shown here as

inner and outer rings. We’ve connected each group only cyclically in

order to avoid clutter, but in the full construction, we start with two

copies of the complete graph K6, one colored according to rij and the

other according to rij = rij + q′ − 1. (b) Reconnect the 12 vertices

according to the prescription shown in the first two lines of (3.17).

(c) Add the last color according to the third line of (3.17). In the

parlance of section 3.2, these last edges are the internal edges, while

the ones colored in the previous step are the external edges.

notationally this can be done by identifying a vertex first by indicating which set it’s

in, say with a Greek index α = 0 or 1, and then which vertex within the set it is, say

with a Roman index i ∈ 0, 1, . . . , q′ − 1. Let rij ≡ rij + q′ − 1. We now claim that

the coloring rαi,βj of Kq defined by

r0i,0j = r1i,1j = rij ,

r0i,1j = r1i,0j = rij ,

r0i,1i = r1i,0i = q − 1 ,

(3.17)

has symmetry group G = G′ × Z2. See figure 3.6 for a diagrammatic illustration of

how this coloring is generated. That the coloring (3.17) inherits the G′ symmetry is

clear from the fact that, for any π ∈ G′, we have that

rαπ(i),βπ(j) = rαi,βj . (3.18)

75

Page 82: Real and p-Adic Physics

But by construction the coloring (3.17) is also invariant under the permutation τ that

acts on the Greek indices as

α→ α + 1 mod 2 , β → β + 1 mod 2 , (3.19)

while leaving the Roman indices unchanged. Furthermore, the coloring (3.17) has no

other symmetries. For coloring automorphisms that only swap around Roman indices

are in one-to-one correspondence with the coloring automorphisms of rij. And if a

coloring automorphism σ changes the Greek index of any index pair αi such that

σ(0i) = 1i′, then for any j

rσ(0i),σ(0j) = r1i′,σ(0j) . (3.20)

But since r0j,0k 6= r1j′,0k′ for all j, j′, k, k′, it follows that σ must change the Greek

index of all index pairs and so must be of the form σ = π τ for some π ∈ G′. This

completes the inductive proof.

In general, for a given symmetry, there are multiple interaction vertices different

from the one generated by the prescription (3.17). An exception, however, occurs

for the maximally symmetric vertex when q = 2v. In this case there is only one

vertex with G = Z v2 . For G consists of the identity element and q − 1 permutations

that commute amongst each other and each consists of q/2 two-cycles. And amongst

permutations in Sq consisting of q/2 two-cycles, one can at most form a set of q − 1

elements that commute amongst each other but don’t share a two-cycle; and any two

such sets are equivalent by conjugation. But if we consider the edges of a given color,

say red, in a maximally symmetric colored graph, then this sub-graph is invariant

under the permutation π that swaps vertices connected by a red edge. But π must

commute with all permutations in G, and so it follows that π ∈ G. Moreover, if we

explicitly write π = (a0a1)(a2a3) . . . (aq−1aq), then we recognize that each two-cycle in

π corresponds to a red edge; in other words, each non-trivial element of G is precisely

76

Page 83: Real and p-Adic Physics

associated with a one-factor. Since there are q− 1 non-trivial elements of G, the one-

factors are all specified once G is specified. And we have argued that G is essentially

unique.

3.4 One-factorizations and equivalent interaction

terms

In this section we determine the conditions under which two theories with actions

of the form (3.2) are equivalent. As discussed briefly in section 3.1, the choice of an

interaction tensor ΩA(0)A(1)···A(q−1) corresponds to a coloring problem on the complete

graph Kq on q vertices.

More precisely, these interaction tensors are in one-to-one correspondence with

ordered one-factorizations of Kq. A one-factor of Kq is a set F of edges such that each

vertex of Kq belongs to a unique edge in F . A one-factorization of Kq is a partition

of the edge-set of Kq into q − 1 one-factors Fi. We denote a one-factorization as

F = F0, . . . , Fq−2, and when speaking of one-factorizations we do not distinguish

between different orderings of the Fi. An ordered one-factorization is a (q − 1)-tuple

F = (F0, . . . , Fq−2), i.e. we have imposed a particular order on a one-factorization.

To see that ordered one-factorizations correspond to interaction vertices, recall that

the fields ψa0...aq−2 don’t have any built-in symmetry under interchange of indices ai.

Each such index is associated with a definite color: For example, we could say that

red is associated with the first index a0, then green with a1, blue with a2, and so

forth. And each color is associated with a one-factor: red for F0, green for F1, blue

for F2, etc.

One can check that any graph automorphism π : Kq → Kq carries (ordered) one-

factorizations into other (ordered) one-factorizations via the actions πF = (π(F0), . . . , π(Fq−2))

and πF ≡ πF , respectively. Two ordered one-factorizations F and G are said to be

77

Page 84: Real and p-Adic Physics

isomorphic if there exists a graph automorphism π such that F = πG. Similarly,

F ' G means there exists π so that F = πG. We can also consider the natural action

of any τ ∈ Sq−1 on ordered one-factorizations τF = (Fτ(0), . . . , Fτ(q−2)). This action

simply permutes the colors associated to each one-factor. In this language, F ' G if

and only if there exists a graph automorphism π and a permutation of colors τ such

that πF = τG.

Given an ordered one-factorization F , we can form the following interaction tensor:

ΩFA(0)...A(q−1) ≡q−2∏r=0

∏〈ij〉∈Fr

Ωa

(i)r a

(j)r, (3.21)

where 〈ij〉 is the edge in Kq running between the vertices i and j and ar is the color

index associated to the one-factor Fr. In writing 〈ij〉, we implicitly assume i < j.

Conversely, any degree q interaction term in which each constituent ψA has a single

color index contracted with each other ψA arises from an interaction tensor of the

form (3.21).

Note that for distinct ordered one-factorizations F and G we end up with distinct

interaction tensors ΩF 6= ΩG. We will however show that if their underlying one-

factorizations F and G are isomorphic then ΩF and ΩG give rise to equivalent theories.

Our argument proceeds in two parts: First we discuss equivalence under the action of

τ ∈ Sq−1 permuting colors, and then we consider the action of a permutation π ∈ Sq

of vertices.

Suppose we make the following field redefinition: φA = ψτ−1A, where τ−1 ∈ Sq−1

and τA ≡ aτ−1(0) . . . aτ−1(q−2). Then:

ΩFA(0)...A(q−1)φA(0) · · ·φA(q−1)

= ΩτFA(0)...A(q−1)ψ

A(0) · · ·ψA(q−1)

. (3.22)

So, modulo a linear field redefinition implemented by a permutation matrix, the

interaction term only depends on the underlying one-factorization F .

78

Page 85: Real and p-Adic Physics

Gq 2 4 6 8 10 12 14

1 0 0 1 2 396 526,910,769 ∼ 1.13× 1018

Z2 1 0 0 1 0 4851 0Z2

2 0 1 0 2 0 0 0Z3

2 0 0 0 1 0 0 0

Table 3.1: The number of isomorphism classes of one-factorizations of Kq with

given symmetry group G = Aut(F). The values for q = 12 and 14

were determined using [65] and [66], respectively; the exact count for

q = 14 is quoted in the main text.

Now view a graph automorphism as a permutation on q vertices, π ∈ Sq. With

the change of indices B(i) = Aπ(i) one can show:

(σψ)πΩFA(0)...A(q−1)ψA(0) · · ·ψA(q−1)

= ΩFA(0)...A(q−1)ψAπ−1(0) · · ·ψAπ

−1(q−1)

= ΩπFB(0)...B(q−1)ψ

B(0) · · ·ψB(q−1)

.

(3.23)

So any two isomorphic ordered one-factorizations give the same interaction term, up

to a sign.3 We can get rid of this sign by sending g → −g.4

We should point out here that isomorphic one-factorizations give equivalent inter-

actions to all orders in N . As we will show in sections 3.7 and 3.8, even non-isomorphic

one-factorizations lead to the same two-point and four-point functions in the melonic

limit, up to a rescaling of g by a power of the order of the coloring automorphism

group G.

For future reference, we note that by Aut(F) we mean all graph automorphisms π

such that πF = F . Aut(F) is the exactly the vertex automorphism group described

in section 3.2. In contrast, Aut(F) consists of all graph automorphisms π such that

πF = F , that is all π that can be undone by a permutation τ ∈ Sq−1 of the colors

3In the presence of multiple interaction terms in the lagrangian, this argument that a permutationof vertices does not lead to a new theory continues to hold true. But it will no longer be true thatan interaction term only depends on the underlying one-factorization F , see section 3.6.

4See [67] and [68] for a careful discussion, in the context of the q = 4 fermionic tensor model,of how the Hamiltonian even at the quantum level transforms in the degree 1 sign representationunder permutations of indices and of how this affects the spectrum of the theory.

79

Page 86: Real and p-Adic Physics

πF = τF . Clearly Aut(F) ≤ Aut(F). In general Aut(F) has a much richer group

structure than Aut(F). See for instance table 3.2 in section 3.5 where, for q = 12, we

list the ten possible values for |Aut(F)| when |Aut(F)| = 2.

In the case q = 12, the vast majority of isomorphism classes of one-factorizations

have trivial Aut(F). It is known in the literature [65] that there are a total of

526,915,620 non-isomorphic one-factorizations (the sum of the two entries in the q =

12 column in figure 3.1) and exactly 252,282,619,805,368,320 distinct one-factorizations

of K12. Implementing the orderly algorithm used in [64] and [65] on the set of one-

factors invariant under a given involution π ∈ S12, we find that there are 1,008,649,635,840

one-factorizations of K12 with Aut(F) = Z2 and that they fall into 4851 isomorphism

classes under permutations of vertices, see section 3.5.

For q = 14, Aut(F) can only be trivial and there are 1,132,835,421,602,062,347

nonisomorphic one-factorizations of K14 [66].

For q ≤ 14, a summary of the number of isomorphism classes of one-factorizations

and therefore the number of inequivalent interaction terms is given in table 3.1. It is

known that for sufficiently large q, the number of non-isomorphic one-factorizations

N(q) satisfies logN(q) ∼ 12q2 log q [69].

3.5 Finding the number of isomorphism classes of

one-factorizations for q = 12 using the orderly

algorithm

In this section we discuss in detail how to find the numbers of melonic interactions

with trivial and non-trivial automorphism groups for q = 12.

We have seen that melonic interaction vertices can be identified with isomorphism

classes of one-factorizations of complete graphs. For q = 12 the total number of

80

Page 87: Real and p-Adic Physics

isomorphism classes was found in [65]. To classify melonic theories it is desirable to

also know how many of these isomorphism classes correspond to melonic interactions

with a non-trivial automorphism group. A one-factorization corresponds to such an

interaction if there exists a vertex permutation π ∈ S verticesq that leaves all one-factors

invariant. We will now show, for q = 12, how to employ the orderly algorithm of [64]

and [65] to find the total number of isomorphism classes of one-factorizations as well

as the number of isomorphism classes corresponding to melonic interactions with

non-trivial automorphisms.

A one-factor Fi can be represented by a set of six pairs

Fi =a0, a1, a2, a3, a4, a5, a6, a7, a8, a9, a10, a11

(3.24)

where all the ai are distinct and belong to the set 0, 1, ..., 11.

By forming a set of I pair-wise disjoint one-factors, we get a partial one-factorization

HI ,

HI = F0, F1, ..., FI−1 (3.25)

where 1 ≤ I < 12. A (full) one-factorization is a partial one-factorization with I = 11.

In the notation of section 3.4, H11 = F .

A key point in implementing the orderly algorithm consists in imposing a lexico-

graphic order amongst partial one-factorizations. One first imposes an order amongst

the one-factors as follows:

Associate to each one-factor Fi a unique ordered 12-tuple Fi by writing it in the

form (3.24) with a2i < a2i+1 and a2i < a2i+2 for all i. Then take F to be given

by

(a0, a1, a2, a3, a4, a5, a6, a7, a8, a9, a10, a11) . (3.26)

Given two distinct one-factors Fi and F ′i , we impose an order such that Fi < F ′i

81

Page 88: Real and p-Adic Physics

if the left-most digit ai by which Fi and F ′i differ is smaller for Fi.

Having ordered the one-factors, we can now order the partial one-factorizations:

Associate to each partial one-factorization HI a unique ordered I-tuple HI by

writing it in the form (3.25) with Fi < Fi+1 for all i. Then take HI to be given

by

(F0, F1, ..., FI−1) . (3.27)

Given two distinct partial one-factorizations HI and H ′I , we impose an order

such that HI < H ′I if the left-most digit ai by which HI and H ′I differ is smaller

for HI .

Now, given a partial one-factorization, we can act on it with a permutation π ∈ S12

by permuting the digits of all the one-factors:

ai → π(ai) . (3.28)

With these preliminaries in place, we are ready to describe the orderly algorithm.

The algorithm works by constructing non-isomorphic partial one-factorizations HI

and then using these to construct larger non-isomorphic partial one-factorizations

HI+1. The way one ensures that all constructed partial one-factorizations HI are

non-isomorphic is by always selecting from any isomorphism class only the one with

the lowest lexicographic ordering. The steps of the algorithm can be described as

follows:

1. Consider all partial one-factorizations H1 containing exactly one-factor and dis-

card the ones for which there exists a permutation π ∈ S12 such that πH1 < H1.

Denote the set of remaining partial one-factorizations H1 by H1.

2. Consider all partial one-factorizations H2 obtained by adjoining to each member

H1 of H1 a one-factor disjoint from the one-factor in H1. Discard the ones for

82

Page 89: Real and p-Adic Physics

which there exists a permutation π ∈ S12 such that πH2 < H2. Denote the set

of remaining partial one-factorizations H2 by H2.

3. Continue on as in step 2 to obtain sets Hn of ordered partial one-factorizations

for n = 3, 4, . . . , 10.

4. Consider all one-factorizations H11 obtained by adjoining to each member H10

of H10 a one-factor disjoint from all the one-factors in H10. Discard the ones for

which there exists a permutation π ∈ S12 such that πH11 < H11. Denote the

set of remaining one-factorizations H11 by H11. This set will contain exactly

one one-factorizations for each isomorphism class of one-factorizations, namely

the one with the lowest lexicographic ordering.

The efficiency of the above algorithm can be improved in several ways. In consid-

ering in step 1 the partial one-factorizations H1 containing one one-factor, one can

restrict consideration to one-factors containing the pair 0, 1. For the one-factors

adjoined to the members of H1 in step 2, one can restrict consideration to one-factors

containing the pair 0, 2. Similarly one can consider restricted sets of one-factors

in implementing the rest of step 3 as well as step 4, where one needs only adjoin

one-factors containing 0, 11 to the members of H10.

The algorithm as described above will find all isomorphism classes of one-factorizations

as was done in [65]. To find only the isomorphism classes corresponding to melonic

interactions with non-trivial automorphism groups, the algorithm must be modified.

As we have seen in section 3.2, for q = 12 the only possible non-trivial automorphism

group is Z2. So for any one-factorization there can be at most a single non-trivial

permutation π ∈ S12 that leaves all one-factors invariant, and we have also seen in

section 3.2 that such a permutation must consist of six two-cycles. Let S12 denote

the subset of permutations in S12 that consist of six two-cycles. In total there are 11!!

such permutations.

83

Page 90: Real and p-Adic Physics

For any one-factorization H11 whose one-factors are invariant under a permu-

tation π ∈ S12, one can act on it with some permutation σ ∈ S12 to generate a

one-factorization σH11 whose one-factors are invariant wrt. any other permutation

π′ ∈ S12. The isomorphism classes of one-factorizations corresponding to a melonic

vertex with G = Z2 each contains a number of elements divisible by 11!! as each of

them contains the same number of one-factorizations for each member of S12. For

the purpose of finding the number of such isomorphism classes by constructing ex-

actly one example from each class, we can therefore restrict attention to one-factors

invariant under one specific permutation π ∈ S12, say

π = (0, 1)(2, 3)(4, 5)(6, 7)(8, 9)(10, 11) . (3.29)

The modifications by which we need to adjust steps 1 to 4 above, then, consist of:

Throughout the implementation of the algorithm, one should only consider one-

factors that are invariant under the permutation (3.29). In total there are 331

such one-factors.

Throughout steps 1 to 4, when checking for each partial one-factorization HI

whether there exists a permutation π such that πHI < HI , one should only

consider permutations π that respect the symmetry under (3.29), ie. permuta-

tions that only mix the 331 one-factors invariant under π amongst themselves,

i.e. permutations that commute with π. This subgroup of S12, the centralizer

CS12(π), contains 46,080 = 12!11!!

= 6! · 26 permutations.

These two restrictions make the orderly algorithm run much faster. One finds that

the total number of isomorphism classes corresponding to G = Z2 melonic interaction

vertices is 4851. Having generated one representative of each of these classes, one can

act on the representatives with permutations in S12 and delete duplicates to find that

there are 97,032,192 one-factorizations consisting of one-factors that are each invariant

84

Page 91: Real and p-Adic Physics

|Aut(F)| 2 4 6 8 10 12 16 24 48 240n 3697 944 13 104 1 38 36 10 6 2

Table 3.2: The number of non-isomorphic one-factorizations n with |Aut(F)| = 2

for each automorphism group order.

under (3.29). Multiplying this number with 11!! one finds the total number of one-

factorizations all of whose one-factors are invariant with respect to some non-trivial

permutation, namely 1,008,649,635,840.

An alternative way to obtain this number of one-factorizations is to use the

orbit-stabilizer theorem in the form (3.14). Acting on the 4851 non-isomorphic one-

factorizations with the centralizer CS12(π), one can find the order of the automorphism

group Aut(F) for each of these one-factorizations. One can then carry out the sum

4851∑i=1

|S12||Aut(F i)|

, (3.30)

and recover the number 1,008,649,635,840. The terms in the sum (3.30) assume but

ten different values, and the sum can be evaluated straightforwardly using the data in

table 3.2, which lists the number of non-isomorphic one-factorizations corresponding

to interaction vertices with Z2 symmetry for any order of Aut(F).

3.6 Isomorphism classes of ordered one-factorizations

for q = 12

For a complete graph of order q, we can consider the set of all ordered one-factorizations.

The orbits of this set under permutations τ ∈ Sq−1 of the one-factors are equal to the

one-factorizations of the graph. The orbits of the one-factorizations under permuta-

tions π ∈ Sq of the vertices are equal to the isomorphism classes that we identified

with melonic interaction vertices and tabulated in figure 3.1. But another set one can

consider is the orbits of ordered one-factorizations under permutations π ∈ Sq of the

85

Page 92: Real and p-Adic Physics

vertices. The relations between these sets and orbits are depicted diagrammatically

in figure 3.7.

Figure 1: The number of equivalence classes of ordered one-factorizations under permutationsof edges and vertices for q = 12.

automorphism group ofthe interaction vertex 1 Z2

ordered one-factorizations 11! · 252,281,611,155,732,480 11! · 1,008,649,635,840one-factorizations 252,281,611,155,732,480 1,008,649,635,840

isomorphism classes ofordered one-factorizations 21,023,467,596,311,040 168,108,272,640

isomorphism classes ofone-factorizations 526,910,769 4851

ordered one-factorizationsS one-factorsq−1

S one-factorsq−1

one-factorizations

S verticesq S vertices

q

ordered one-factorizations

isomorphism classes of isomorphism classes of

one-factorizations

1

Figure 3.7: The relationship between ordered one-factorizations, one-

factorizations, and their isomorphism classes under vertex per-

mutations. An arrow labeled by a group indicates that the set at the

head end can be identified with the orbits of the set at the tail end

under the action of the group.

We argued in section 3.2 that while an ordered one-factorization of a graph can

immediately be translated into a melonic interaction vertex, permutations of the ver-

tices of the graph correspond to commuting fields past each other, and permutations

of the one-factors correspond to field redefinitions, neither of which operations yield

physically distinct theories, so that in counting the number of theories via ordered

one-factorizations one should mod out by S verticesq and S one-factors

q−1 . But this reasoning

only applies to lagrangians with a single interaction term. For theories with multiple

interaction terms, one can still commute the fields around in each interaction term,

and so one should still mod out by S verticesq . But field-redefinitions can only be ap-

plied globally, and so one should no longer count the number of interaction terms by

modding out by S one-factorsq−1 . A crude upper bound on the number of distinct theories

with n interaction vertices of order q is the n-th power of the number of isomorphism

classes of ordered one-factorizations. The number of isomorphism classes of ordered

one-factorizations, call it C, is therefore not without interest in the context of melonic

86

Page 93: Real and p-Adic Physics

models. In the remainder of this section we will derive this number in the case of

q = 12 and for interaction vertices with and without trivial automorphism groups.

Letting X denote a set acted on by a finite group G, Xg denote the set of elements

in X that are left invariant by g ∈ G , and X/G denote the set of orbits of X under

G, Burnside’s lemma tells us that

|X/G| = 1

|G|∑g∈G

|Xg| . (3.31)

By taking X to be the set of ordered one-factorizations and G to be S12 acting on

vertices, we can use Burnside’s lemma to find C. If in the sum over g ∈ G in (3.31)

we take g to be the identity element, then |Xg| is equal to |X| itself, that is the total

number of ordered one-factorizations, which in turn is equal to 11!M , where M is the

total number of one-factorizations. From [65] we know that

M = 252,282,619,805,368,320 . (3.32)

If in the sum over g ∈ G in (3.31) we consider the sum over the permutations g ∈ S12,

with S12 defined in section 3.5, then this partial sum will count each ordered one-

factorization whose one-factors are all invariant under some permutation g ∈ S12

exactly once. This number is equal to 11!M2, where M2 is the total number of one-

factorizations whose one-factors are all invariant under some permutation g ∈ S12.

We found in section 3.5 that this number is equal to

M2 = 1,008,649,635,840 . (3.33)

If in the sum over g ∈ G in (3.31), g is not the identity element nor belongs to S12,

then |Xg| = 0.

87

Page 94: Real and p-Adic Physics

In conclusion, Burnside’s lemma tells us that

C =1

12!(11!M + 11!M2) = 21,023,635,704,583,680 . (3.34)

This, the number of isomorphism classes of ordered one-factorizations, can be split

into C1, the number of classes corresponding to interaction vertices with trivial auto-

morphism group, and C2, the number of classes corresponding to interaction vertices

with automorphism group Z2,

C = C1 + C2 . (3.35)

To find out how C splits into C1 and C2 we can use the orbit-stabilizer theorem in

the form of (3.14) and (3.30),

|X| =∑

orbits o

|G||stab(o)|

. (3.36)

The orbits of ordered one-factorizations under S12 acting on vertices fall into two sets.

There are C1 orbits o for which |stab(o)| = 1 and C2 orbits o for which |stab(o)| = 2.

Hence,

11!M = 12!

(C − C2

1+C2

2

). (3.37)

Solving for C2, one finds that

C2 = 168,108,272,640 . (3.38)

The numbers of ordered and un-ordered one-factorizations and the numbers of iso-

morphism classes of these under vertex permutations for q = 12 are summarized in

the table in table 3.3.

88

Page 95: Real and p-Adic Physics

automorphism group ofthe interaction vertex trivial Z2

ordered one-factorizations 11! · 252,281,611,155,732,480 11! · 1,008,649,635,840

one-factorizations 252,281,611,155,732,480 1,008,649,635,840

isomorphism classes ofordered one-factorizations 21,023,467,596,311,040 168,108,272,640

isomorphism classes ofone-factorizations 526,910,769 4851

Table 3.3: The number of equivalence classes of ordered one-factorizations under

permutations of one-factors and vertices for q = 12.

3.7 The two-point function and the Schwinger-Dyson

equation

The kinetic term in the action (3.2) gives rise to a free propagator that in momentum

space is given by

Ga0a1...aq−2b0b1...bq−2

0 (ω) =√

sgn(−1)sgn(ω)

|ω|sq−2∏i=0

Ωaibi ≡ G0(ω)

q−2∏i=0

Ωaibi , (3.39)

where the matrix Ωab is the inverse of Ωab, that is, ΩabΩbc = δca. With the normaliza-

tion of the kinetic term as given in (3.2), we must choose

√sgn(−1) =

1 if sgn(−1) = 1

i if sgn(−1) = −1 .

(3.40)

Once we include an interaction term, the propagator picks up loop corrections,

but in the limit where g → 0 and N → ∞ such that g2N(q−1)(q−2)

2 is kept fixed, only

melonic diagrams survive. The melonic contributions to the free propagator can all

be obtained by iteratively applying the melonic insertion to the free propagator as

shown on figure 3.8.

Adding together the free propagator and all the melonic corrections yields the

89

Page 96: Real and p-Adic Physics

dressed propagator

Ga0a1...aq−2b0b1...bq−2(ω) ≡ G(ω)

q−2∏i=0

Ωaibi . (3.41)

Just as in the cases of the SYK model [70] and the Klebanov-Tarnopolsky tensor

model with q = 4 [52], the dressed propagator satisfies the Schwinger-Dyson equa-

tion depicted schematically in figure 3.9. It first seems surprising that the right-most

propagator in figure 3.9 has no melonic insertion. This is correct because any Feyn-

man diagram contributing to the dressed propagator must have a right-most melonic

insertion, and a free propagator attaches to it from the right.

To work out the exact mathematical expression for the Schwinger-Dyson equation,

it is necessary to consider the automorphism group of the interaction as well as the

number of sign flips involved in index contraction.

The first melonic correction to the free propagator, depicted on the right-hand

side of figure 3.8, contains two interactions each of which is described by a colored

graph with vertices labeled from 0 to q − 1. For a given Feynman diagram corre-

sponding to the first melonic correction, define a permutation σ ∈ Sq by requiring

that if an external propagator enters the left interaction vertex at vertex i, then an

external propagator enters the right interaction vertex at vertex σ(i); and likewise

requiring that an internal propagator connected to vertex j in the left interaction

i σ(i)

j

k

σ(j)

σ(k)

i σ(i)

j

k

σ(j)

σ(k)

...

...︸ ︷︷ ︸q− 1 undressed propagators

= +

...

...

︸ ︷︷ ︸q− 1 dressed propagators

1

Figure 3.8: A melonic insertion.

90

Page 97: Real and p-Adic Physics

i σ(i)

j

k

σ(j)

σ(k)

...

...︸ ︷︷ ︸q− 1 undressed propagators

= +

...

...

︸ ︷︷ ︸q− 1 dressed propagators

1

Figure 3.9: The Schwinger-Dyson equation for the dressed propagator.

vertex connects to vertex σ(j) in the right interaction vertex. See figure 3.10.

We claim now that such a Feynman diagram is melonic if and only if σ belongs

to the automorphism group G of the interaction. The claim can be proved by the

following reasoning:

The two interactions are connected by q−1 internal propagators, each of which

carry q − 1 threads, giving a total of (q − 1)2 internal threads.

The q − 1 threads of the left external propagator must connect to the q − 1

threads of the right external propagator, so there are at most (q − 1)(q − 2)

internal threads that can partake in index loops.

Since at least two internal threads are needed for an index loop, there can be at

i σ(i)

j

k

σ(j)

σ(k)

i σ(i)

j

k

σ(j)

σ(k)

...

...︸ ︷︷ ︸q− 1 undressed propagators

= +

...

...

︸ ︷︷ ︸q− 1 dressed propagators

1

Figure 3.10: To get a melonic diagram, the external legs must be connected to

vertices in the two interactions that are related by an automorphism.

91

Page 98: Real and p-Adic Physics

i σ(i)

j

k

σ(j)

σ(k)

i σ(i)

j

k

σ(j)

σ(k)

...

...︸ ︷︷ ︸q− 1 undressed propagators

= +

...

...

︸ ︷︷ ︸q− 1 dressed propagators

1

Figure 3.11: All index loops can be brought to have a uniform orientation by

an even number of arrow flips. To get a uniform orientation of any

external thread an odd number of arrows flips are requisite.

most (q− 1)(q− 2)/2 index loops—the number required for a melonic diagram.

This maximum number of loops is achieved when and only when the external

threads only pass once through the internal propagators and the each index

loop contains exactly two internal threads.

Labeling the vertex that the left external propagator is incident to by i, the fact

that the external threads must pass only once through the internal propagators

is equivalent to the statement that rij = rσ(i)σ(j) for all j 6= i.

The fact that each index loop must contain exactly two internal threads is

equivalent to the statement that for any two indices j and k different from i,

rjk = rσ(j)σ(k).

We see then that the melonic diagrams are exactly those for which rlm = rσ(l)σ(m)

for all l,m, ie. those for which σ ∈ G. Since in writing down the Feynman diagram

corresponding to the first melonic correction we may take the left external propagator

to be incident to any vertex i, while the right external propagator can only be incident

to vertices i′ for which i′ = σ(i) for some σ ∈ G, the total number of Feynman

diagrams contributing is q|G|. And for each additional melonic insertion we pick up

an extra copy of this factor.

The other subtlety to consider before writing down the Schwinger-Dyson equation

is whether the contraction of index loops causes a sign difference between O(N) and

92

Page 99: Real and p-Adic Physics

Sp(N) symmetric models. To account diagrammatically for the possibility of Sp(N)

symmetry, we replace each thread with an arrow. The orientation of any arrow can

be flipped at the cost of a factor of σω, and index loops of arrows with uniform

orientation can be contracted to give a factor of N without picking up a sign.

For the first melonic correction to the free propagator we may take the threads

of all propagators to point from left to right as in figure 3.11. For a given Feynman

diagram described by a permutation σ ∈ G as explained above, we may for any l

and m take the the arrow between vertices l and m to have the same orientation as

the arrow between vertices σ(l) and σ(m) since any automorphism involves an even

number of arrow flips. In this case, all index loops consist of two arrows with one

orientation and two arrows with the opposite orientation, and so all the loops may

be contracted without picking up a sign. But to obtain a uniform orientation of the

external threads running from left to right and recover the index structure of the free

propagator, it is necessary for each of the q− 1 colors to flip one thread in one of the

two interactions. Hence, we must in total flip an odd number off arrows, and so we

conclude that we pick up a factor of σΩ with each melonic insertion. This fact was

shown in [27] to apply for q = 4, but we see now that it holds true for all q.

Taking into account the symmetry of the interaction and the index structure of

the corrections, we arrive finally at the Schwinger-Dyson equation

G(t1 − t2) = G0(t1 − t2)

+ q|G|(g

|G|

)2

σΩ N(q−1)(q−2)

2

∫dt dt′G(t1 − t)G(t− t′)q−1G0(t′ − t2) .

(3.42)

To slightly shorten expressions and remove the explicit dependence on G, we find it

useful to introduce the definition

g ≡√

q

|G|N

(q−1)(q−2)2 g . (3.43)

93

Page 100: Real and p-Adic Physics

3.7.1 The IR solution

The Schwinger-Dyson equation can be expressed in term of convolutions as

G(t) = G0(t) + σΩg2(G ∗Gq−1 ∗G0)(t) . (3.44)

In the infra-red limit, g is large, and to leading order in 1/g we can set G(t) = 0. To

solve (3.44) in the IR, we will adopt the methodology of [27], and so we introduce

multiplicative characters defined as

πs(t) ≡ |t|s, πs,sgn(t) ≡ |t|s sgn(t) . (3.45)

The Fourier transform of these multiplicative characters are given by

F [πs](ω) = Γ(πs+1) π−s−1(ω) , F [πs,sgn](ω) = Γ(πs+1,sgn) π−s−1,sgn(ω) , (3.46)

where Γ(πs) and Γ(πs,sgn) are instances of Gel’fand-Graev gamma functions, see sec-

tion 1.6. Using the above, we may express the bare propagator in momentum and

position space as

G0(ω) =√

sgn(−1) π−s,sgn(ω) , G0(t) =√

sgn(−1) Γ(π1−s,sgn) πs−1,sgn(t) .

(3.47)

As an ansatz for solving (3.44) with the left-hand side set to zero, we choose

G(t) = b π− 2q,sgn(t) , (3.48)

and so the IR Schwinger-Dyson equation in position space reads

−π1−s,sgn = σΩ g2 bq π− 2

q,sgn ∗ π 2(1−q)

q,sgn∗ π1−s,sgn . (3.49)

94

Page 101: Real and p-Adic Physics

Using (3.46) to Fourier transform the multiplicative characters, we get the momentum

space Schwinger-Dyson equation

−πs,sgn = σΩ g2 bq Γ(π q−2

q,sgn)Γ(π 2−q

q,sgn) π 2−q

q,sgn ∗ π (q−2)

q,sgn∗ πs,sgn . (3.50)

The multiplicative characters all cancel, and we read off directly that

bq =−1

σΩ g2 Γ(π q−2q,sgn)Γ(π 2−q

q,sgn)

. (3.51)

The s dependence is seen to cancel out entirely. Insofar as the ansatz (3.48) describes

the true IR behavior, the dressed propagators of the various non-local theories with

different values of s all flow to the same function in the IR. In the case of bosonic

theories, however, where the potential can be unbounded from below, the physical

significance of the IR solution (3.48) is uncertain. Nonetheless, we explicitly include

bosonic theories under the scope of theories we are subjecting to formal perturbation

theory since bosonic theories are required in order to write down an adelic relation

connecting Archimedean and p-adic theories, as we will see in section (3.8.1).

In section 3.7.3 we will see that for a large subset of the p-adic theories, the full

Schwinger-Dyson equation (3.42) can be solved exactly, and in these cases one can

verify by inspection of the full solution that the two-point function flows to the IR

solution described by equations (3.48) and (3.51).

3.7.2 The zoo of theories

As we are working in Euclidean space, the reality of the action (3.2) dictates that

G(t) should be real too. The ansatz (3.48) can therefore only provide an accurate

description of the IR behavior if b is real, that is, if the right-hand side of (3.51) is

positive. In other words, the sign of Γ(π q−2q,sgn)Γ(π 2−q

q,sgn) must be opposite to that

of σΩ. For any choice of number field R or Qp and any choice of sign function in

95

Page 102: Real and p-Adic Physics

K condition τ Γ(π q−2q,sgn)Γ(π 2−q

q,sgn) σΩ σψ explanation

R 1 − 2πqq−2

tan(πq

)1 1 O(N) bosonic

R −1 − 2πqq−2

cot(πq

)1 −1 O(N) fermionic

C 1 − π2q2

(q−2)2 1 1 O(N) bosonic

Qp p odd 1 − (p2/q−1)(p2−p2/q)

p(p−p2/q)2 1 1 O(N) bosonic

Qp p odd ε (p2/q+1)(p2+p2/q)

p(p−p2/q)2 −1 −1 Sp(N) fermionic

Qp p ≡ 1 mod 4 p 1/p −1 −1 Sp(N) fermionic

Qp p ≡ 1 mod 4 εp 1/p −1 −1 Sp(N) fermionic

Qp p ≡ 3 mod 4 p −1/p 1 −1 O(N) fermionic

Qp p ≡ 3 mod 4 εp −1/p 1 −1 O(N) fermionic

Q2 1 −5·41/q−4−161/q

2(41/q−2)2 1 1 O(N) bosonic

Q2 −1 −1/4 1 −1 O(N) fermionic

Q2 2 1/8 −1 −1 Sp(N) fermionic

Q2 −2 −1/8 1 −1 O(N) fermionic

Q2 3 −1/4 1 −1 O(N) fermionic

Q2 −3 (1+41/q)(4+41/q)

2(2+41/q)2 −1 −1 Sp(N) fermionic

Q2 6 −1/8 1 −1 O(N) fermionic

Q2 −6 1/8 −1 −1 Sp(N) fermionic

Table 3.4: Table of melonic theories, Archimedean or ultrametric. Shaded rows

indicate theories for which an exact solution to the Schwinger-Dyson

equation is given in subsection 3.7.3.

the action, this requirement together with the sign constraint (3.5) uniquely specifies

whether the theory must be bosonic or fermionic and O(N) and Sp(N) symmetric.

Generalizing the table of theories in [27] to all values of q, table 3.4 lists the melonic

theories with renormalization group flow from a free theory in the UV to a strongly

interacting fixed point in the IR. The choice of τ in table 3.4 amounts to a choice of

the sign function, as explained in section 1.6. Often, the value of τ involves ε, which

stands for any integer that is not a square modulo p.

96

Page 103: Real and p-Adic Physics

3.7.3 Full solution to the Schwinger-Dyson equation for direction-

dependent theories

Recall that any p-adic number x can be represented by a series expansion

x = pv(x)

∞∑m=0

cm pm (3.52)

where cm ∈ 0, 1, ..., p− 1 and c0 6= 0.

It was shown in [27] for q = 4 that the Schwinger-Dyson equation (3.44) can be

solved exactly for p-adic theories when the sign function is direction-dependent, that

is, when the sign function sgn(x) depends not only on the norm |x| but also on the first

p-adic digit c0 in the expansion (3.52). These theories are indicated by shaded rows

in table 3.4, and the solutions of [27] generalize straightforwardly to higher values of

q.

For each of the theories, the solution can be written as

G(t) = G(|t|) sgn(t) , (3.53)

where G(|t|) is the unique real root of a q-th order polynomial.

For the direction-dependent theories with odd p, ie. those with a sign function

defined by τ = p or τ = εp, G(|t|) is given by the real solution to the equation

G(|t|) =√

sgn(−1)Γ(π1−s,sgn)|t|s−1

(1− g2

p|t|2 G(|t|)q

). (3.54)

For p = 2 and τ = −1 or τ = 3, G(|t|) is given by the real solution to

G(|t|) =√

sgn(−1)Γ(π1−s,sgn)|t|s−1

(1− g2

22|t|2 G(|t|)q

). (3.55)

97

Page 104: Real and p-Adic Physics

And p = 2 and τ = ±2 or τ = ±6, G(|t|) is given by the real solution to

G(|t|) =√

sgn(−1)Γ(π1−s,sgn)|t|s−1

(1− g2

23|t|2 G(|t|)q

). (3.56)

In all cases, we choose√

sgn(−1) as in (3.40).

3.8 The four-point function

As shown in [52], the four-point function of the Klebanov-Tarnopolsky tensor model

with rank three fermions has the same structure as the SYK model [70], and this

result generalizes to values of q > 4 [55]. Working to sub-leading order in the melonic

limit, the four-point correlator can be decomposed as

ΩA1A2ΩA3A4

⟨T (ψA1(t1)ψA2(t2)ψA3(t3)ψA4(t4))

⟩=

N2(q−1)G(t12)G(t34) +N q−1

(Γ(t1, t2, t3, t4) + σψΓ(t1, t2, t4, t3)

)+O(N q−2) ,

(3.57)

where Γ stands for a sum of so-called ladder Feynman diagrams that can be expanded

as Γ =∑

n Γn according to the number n of sets of q−2 rungs in the ladder diagrams.

Schematically, for q = 8:

Γ(t1, t2, t3, t4) =

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A3

A4

...

...

A1

A2

i j

σ(i) σ(j)

k

l

σ(k)

σ(l)

1

+

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A3

A4

...

...

A1

A2

i j

σ(i) σ(j)

k

l

σ(k)

σ(l)

1

+

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A3

A4

...

...

A1

A2

i j

σ(i) σ(j)

k

l

σ(k)

σ(l)

1

+ ...

(3.58)

Here black lines represent propagators while gray lines stand for contraction through

the matrix ΩAB. But for a more exact understanding of the four-point function, we

98

Page 105: Real and p-Adic Physics

must also consider the threads running within each interaction vertex and endow each

line with an orientation. For the index contraction in (3.57), the arrows of the first

ladder diagram are oriented as

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A3

A4

...

...

A1

A2

i j

σ(i) σ(j)

k

l

σ(k)

σ(l)

1

. (3.59)

By flipping two arrows, each corresponding to q−1 interchanges of indices in a matrix

Ωab or Ωab, we get an oriented diagram,

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A3

A4

...

...

A1

A2

i j

σ(i) σ(j)

k

l

σ(k)

σ(l)

1

. (3.60)

Since an even number of arrow flips were performed, there is no σΩ dependence so

that

Γ0(t1, t2, t3, t4) = σψG(t1 − t3)G(t2 − t4) . (3.61)

The absence of a factor of σΩ in (3.61) owes directly to the choice of contraction with

matrices ΩA1A2ΩA3A4 rather than say ΩA1A2ΩA4A3 in (3.57), and so is ultimately the

result of a convention. The important question to ask is whether the insertion of

each new set of rungs in a ladder diagram leads to a factor of σΩ. Consider therefore,

as in figure 3.12, the diagram obtained by appending an extra set of rungs to some

oriented ladder diagram.

The propagator that forms part of the top rail of the ladder immediately to the

left of the newly appended set of rungs may be incident to any vertex i of the top

interaction vertex. The propagator that forms part of the top rail immediately to the

right of the appended sets of rungs may be incident to any other vertex j 6= i of the

99

Page 106: Real and p-Adic Physics

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A1

A2

A3

A4

A3

A4

...

...

A1

A2

i j

σ(i) σ(j)

k

l

σ(k)

σ(l)

1

Figure 3.12: Insertion of an extra set of rungs in an oriented ladder. Not all

internal threads in the interactions nor all rungs have been drawn

so as to not clutter up the figure.

topmost interaction vertex. But for a given choice of i and j, in order to get a ladder

diagram that is not suppressed in the melonic limit, the two propagators of the bottom

rail immediately to the left and right of the bottom interaction vertex must be given,

respectively, by σ(i) and σ(j) for some σ ∈ G. And similarly the appended rungs

must connect vertex l of the top interaction vertex to vertex σ(l) of the bottom one for

all l 6= i, j. These facts follow immediately from the above discussion of the two-point

function once we note that 1) the threads of the propagators incident to vertices i

and σ(i) all partake of the same index loops; and 2) from the point of view of index

contraction we may look upon the propagators incident to vertices i and σ(i), the

arrows running to and between A3 and A4, and the two appended interaction vertices

and the rungs running between them as all forming part of a melonic insertion into

one single propagator. Since there are q choices of vertex i, (q − 1) choices of vertex

j, and |G| choices of permutation σ, the total number of melonic Feynman diagrams

that contribute to the ladder diagram in question is q(q − 1)|G|.

As in our consideration of the two-point function, we can assume that the orien-

100

Page 107: Real and p-Adic Physics

tation of an arrow between any two vertices i′ and j′ of the top interaction vertex is

the same as the orientation of the arrow between σ(i′) and σ(j′) since automorphisms

induce an even number of arrow flips. As to the q−2 arrows that make up the rungs,

these all have the same orientation prior to flipping any arrows, as one can find by

computing the four-point function via functional differentiation. As in figure 3.12 we

will take these arrows to point downwards rather than upwards, but this is an arbi-

trary choice that does not affect the parity of arrow flips needed to make all index

loops have a uniform orientation of arrows.

To determine whether appending an extra set of rungs gives rise to an overall

factor of σΩ, we need to consider all the index loops involved. These fall into four

types:

1. The index loops running from i through one of the rungs of the ladder and

through σ(i) without passing through A3 and A4, as illustrated in blue on

figure 3.12. Because there are q − 2 such loops and they all require the same

number of arrow flips to obtain uniform orientation, no net factor of σΩ is

introduced on account of these index loops.

2. The index loops running from j through A3 and A4 and back to j through one

of the rungs in the ladder, illustrated in red on figure 3.12. Again these q − 2

loops are even in number and require the same number of arrow flips for uniform

orientation, so again no net factor of σΩ is introduced.

3. The index loops that run between the two interaction vertices and consist of

four threads each, two in two rungs and two within the two interaction vertices.

There are (q − 2)(q − 3)/2 such loops, an example of which is illustrated in

yellow on figure 3.12. Each of these index loops consists of two threads with

one orientation and two index loops with the opposite orientation, so once again

no net factor of σΩ is introduced in bringing about uniform orientation.

101

Page 108: Real and p-Adic Physics

4. Lastly, there is the index loop that runs between i and j and σ(i) and σ(j),

illustrated in green on figure 3.12. Since we are assuming that the index loops

of the original ladder diagram, prior to insertion of an extra set of rungs, each

had a uniform orientation of arrows, two arrows must be flipped in order to give

this index loop a uniform orientation (the arrow between σ(i) and σ(j) and the

arrow between σ(j) and A4 on figure 3.12) and again we do not pick up any net

factor of σΩ.

In summary, there are no relative sign differences in ladder diagrams between SO(N)

and Sp(N) symmetric tensor models.

Having worked out the subtleties relating to Sp(N) symmetry and the automor-

phism group of the interaction vertex, we are ready to write down the recursive

relation describing ladder diagrams:

Γn+1(t1, t2, t3, t4) =

∫dt dt′K(t1, t2, t, t

′)Γn(t, t′, t3, t4) (3.62)

where the integration kernel is given as

K(t1, t2, t, t′) = σψq(q − 1)|G|

(g

|G|

)2

N(q−1)(q−2)

2 G(t1 − t)G(t2 − t′)G(t− t′)q−2

= σψ(q − 1)g2G(t1 − t)G(t2 − t′)G(t− t′)q−2 .

(3.63)

In the IR, we can plug in the expression for the dressed propagator derived in

section 3.7.1 to obtain

K(t1, t2, t3, t4) =− σψσΩ(q − 1)

Γ(π q−2q,sgn)Γ(π 2−q

q,sgn)

π− 2q,sgn(t13) π− 2

q,sgn(t24) π 2(2−q)

q,sgn

(t34) .

(3.64)

Following [71] and [52] and defining

v(t1, t2) ≡ πh− 2q,sgn(t12), (3.65)

102

Page 109: Real and p-Adic Physics

the integral eigenvalue-equation to solve in order to find the scaling dimensions of

two-particle operators is given by

v(t1, t2) =1

g(h, q)

∫dt3 dt4K(t1, t2, t3, t4) v(t3, t4) . (3.66)

Changing variables to

u ≡ t13 , v ≡ t42 , (3.67)

makes it manifest that the integral in the eigenvalue equation is a convolution since

∫dt3 dt4 π− 2

q,sgn(t13)π− 2

q,sgn(t24) π

h+2(1−q)q−h,sgn

(t34) =

sgn(−1)

∫du dv π− 2

q,sgn(u) π− 2

q,sgn(v) π

h+2(1−q)q

,sgn(t12 − u− v) .

(3.68)

Invoking the sign constraint (3.5), the eigenvalue equation can therefore be written

as

πh− 2q,sgn = − (q − 1)

g(h, q)Γ(π q−2q,sgn)Γ(π 2−q

q,sgn)

π− 2q,sgn ∗ π− 2

q,sgn ∗ πh+

2(1−q)q

,sgn. (3.69)

Fourier-transforming this equation using (3.46), the multiplicative characters cancel,

as do two of the gamma functions, and we find that

g(h, q) = −(q − 1)Γ(π q−2

q,sgn)Γ(πh+ 2−q

q,sgn)

Γ(π 2−qq,sgn)Γ(πh+ q−2

q,sgn)

. (3.70)

This formula is valid for real as well as p-adic numbers, and for an action (3.2) with

a kinetic term with any sign function.

On the real numbers, selecting the usual sign function in the action (3.2), the

above equation reproduces the fermionic result of the SYK and melonic tensor models,

equations (3.73) in [71] and (6.8) in [55]. Selecting the trivial sign on the reals, the

above equation reproduces the bosonic result, equation (4.14) in [52] with d = 1.

103

Page 110: Real and p-Adic Physics

On the p-adic numbers, there are multiple inequivalent sign functions, each of

which can be labeled by a p-adic number τ as explained in section 1.6. For the sign

functions characterized by τ = p or τ = εp, g(h, q) reduces to 1− q. For τ = ε, g(h, q)

is a non-constant function in h, but the equation g(h, q) = 1 never has a solution for

the p-adic theories.

3.8.1 Adelic product formula for the integral eigenvalues

In [30] it is demonstrated how, by invoking the functional equations of suitably cho-

sen Dirichlet L-functions, one can derive an adelic product formula for the signed

Gel’fand-Graev gamma functions. By selecting a fixed rational number τ , one picks

out a sign function for each of the number fields R and Qp. For each of these fields

with associated sign function sgnτ (x) there is a signed character πs,sgn(x) and an asso-

ciated gamma function Γ(πs,sgn). Taking the product over all these gamma functions

for any fixed complex number s, one gets the usual sign function of τ :

∏K=R,Qp∀p

Γ(πs,sgn) =

1 for τ > 0

−1 for τ < 0 .

(3.71)

Note that for any choice of τ there will be many fields K for which sgnτ (x) is

the trivial sign character, ie. unity, so that πs,sgn reduces to πs. For example, this

happens for K = R when τ is positive.

As an immediate consequence of (3.71), we find that for fixed rational τ , complex

h, and even number q, the integral eigenvalues g(h, q) satisfy the adelic product

formula

∏K=R,Qp∀p

g(h, q)

1− q= 1 . (3.72)

For most choices of τ , this product mixes bosonic and fermionic, and O(N) and Sp(N)

104

Page 111: Real and p-Adic Physics

symmetric theories.

3.9 Outlook

Given the relative uniqueness and simplicity of melonic theories with quartic and sex-

tic vertices, it is surprising that melonic theories proliferate and diversify for larger

orders q of the interaction vertex. Already at q = 8, there are six different types of

vertices, with up to Z32 symmetry. The situation can be compared with matrix mod-

els, where if we restrict to quartic vertices, the most commonly studied interactions

are tr Φ4 and (tr Φ2)2. The first of these has Z4 symmetry, and the second has Z32

symmetry. If we restrict to only one of these two vertices, then we get remarkably

different behavior in the large N limit: tr Φ4 leads to a planar limit, while (tr Φ2)2

leads to bubble diagrams. Mixing the two gives some interesting modifications of

scaling exponents of the pure tr Φ4 theory [72, 73, 74, 75]. In contrast, for melonic

theories the treatment of the two-point and four-point functions proceeds almost iden-

tically for the myriad theories we can construct. The one salient difference among

the theories considered here is that the effective coupling constant that enters into

the self-consistent treatment of two- and four-point functions includes the inverse half

power of the order of the symmetry group of the interaction vertex. An interesting

question for future work is to see whether combining different interaction vertices

could significantly alter the analysis, for example through some cancellations or mod-

ified infrared scaling. Another interesting question for future work is to examine the

extent to which the operator counting of [67] generalizes to higher melonic theories.

The construction of q-fold interaction vertices in this chapter is, by necessity,

somewhat detailed, amounting to a coloring of the complete graph of q vertices with

q − 1 colors so that all edges meeting at a vertex have different colors. The combi-

natorial problem of counting all such colorings is formidable, and it is related to the

105

Page 112: Real and p-Adic Physics

problem of one-factorizations of the complete graph, for which results are available

only for modest values of q. However, as demonstrated in section 3.3 one can con-

firm the existence of colorings with all possible symmetries groups, namely Zn2 with n

bounded above by v if q = 2v and by v−1 if q = u2v with odd u greater than 1. While

this construction of interaction vertices is rooted in the natural generalizations of the

Klebanov-Tarnopolsky model [52], we cannot claim to have exhausted all possible

generalizations with melonic limits. Here is a useful way to frame the question: If we

restrict the field content to a single real field ψa0a1...aq−2 with q − 1 N -valued indices

all of which must be contracted amongst each other, is the most general interaction

vertex with a melonic limit a linear combination of the interaction vertices that can

be described in terms of colorings of a complete graph of q vertices?

106

Page 113: Real and p-Adic Physics

4

Holographic Duals of Nontrivial Characters in

p-adic AdS/CFT

This chapter is based on [4] coauthored with Steven S. Gubser and Christian Jepsen.

The paper was dedicated to the memory of Peter G. O. Freund, whose discoveries

are foundational to much of this thesis. We thank M. Heydeman for substantial

input in early phases of this work, in particular for identifying the U(1) gauge field

configurations as related to Paley graphs.

Of Peter Freund’s many ideas in theoretical physics, it was clear that p-adic

string theory [7, 6, 9], also studied by Volovich in [8], was one of his favorites. The

strangeness of the p-adic numbers, the unexpectedness of Freund and Olson’s idea

to replace the reals with the p-adics on the boundary of the open string worldsheet,

and the simplicity of the resulting scattering amplitudes, all contribute to the charm

of the subject. The deep question of why the reciprocal of the Veneziano amplitude

factorizes into a product of its p-adic relatives remains mysterious. It causes one to

wonder whether, even now, we have fully plumbed the depths of perturbative string

dynamics.

In p-adic AdS/CFT [10, 11], one looks at some of the surprising features of p-

adic string theory in a new light. Recall that (see section 1.8) that in an important

107

Page 114: Real and p-Adic Physics

precursor to p-adic AdS/CFT [42], Zabrodin defined a free massless scalar action over

the Bruhat-Tits tree Tp, whose boundary is the projective line P1(Qp). Integrating out

the bulk scalar was shown to result in the correlators that Freund and Olson needed to

obtain the analog of the Veneziano amplitude for p-adic strings. An updated version

of this integrating out process is the computation of holographic Green’s functions in

p-adic AdS/CFT, with the Bruhat-Tits tree playing the role of the bulk geometry.1

Free massless scalars in the bulk are just the beginning: In p-adic AdS/CFT one

wants to consider mass terms and their relation to boundary conformal dimension,

and also non-linear bulk dynamics dual to n-point boundary Green’s functions with

more interesting structure than can be obtained from Wick contractions.

In [42] as well as later works, attention focused on scalar fields in the bulk ge-

ometry, dual to scalar operators on the boundary. Likewise on the field theory side,

the study of the operator product expansion [41] focused on scalar operators. Even

the gravitational dynamics of [25] is a scalar theory, because the bulk variable is edge

length on Tp. The dual boundary operator is found to be a scalar whose scaling di-

mension equals the dimension of the boundary as a vector space over Qp. Boundary

theory correlators involving sign characters were considered in [9, 29, 28, 30, 31] in

connection with p-adic string amplitudes and supersymmetry. Work on fermionic

p-adic field theories continued in [32] with a study of a relative of the Gross-Neveu

model, and the recent work [27] investigates both fermionic and bosonic melonic the-

ories over Qp. General comments on higher spin can be found in [11]. But no bulk

dual of non-scalar operators was suggested in any of these works. Another precursor

of p-adic AdS/CFT is the stochastic cellular model eternal inflation studied in [77];

but there too the treatment was restricted to scalar operators on the boundary (best

understood as an analog of future infinity in de Sitter space). The aim of this chapter

is to present first steps in a larger project of finding bulk duals to known non-scalar

1See however [76] for a recent study of holography involving p-adic numbers in the context of acontinuous bulk geometry.

108

Page 115: Real and p-Adic Physics

boundary theories on P1(Qp).

As suggested in [42, 30, 27], we consider boundary Green’s functions of the form

G(x) =C sgnx

|x|2∆p

, (4.1)

where C is a constant, | · | denotes the p-adic norm, and sgnx is a multiplicative sign

character on Qp, as defined in section 1.6. We want to inquire, when and how can we

extract a Green’s function like (4.1) from a bulk construction?

To further motivate the study of Green’s function of the type (4.1), consider the

corresponding Fourier space expression

G(k) = C(sgn k)|k|2∆−1p . (4.2)

For comparison, fermionic correlators in ordinary AdSd+1/CFTd take the form G(k) =

Cγµkµ|k|2∆−d−1 (where now | · | is the norm on Rd instead of the p-adic norm); see for

example [78]. Our assertion is that the factor sgn k is in rough analogy to the the factor

γµkµ/|k| that appears in the Archimedean case. The first point of similarity is that

(sgn k)2 = 1, just as (γµkµ/|k|)2 = 1. Furthermore, in analogy to the transformations

of γµkµ/|k| under rotations, sgn k is a representation of the rotation group on Qp,

which comes from multiplication by p-adic numbers with norm 1. This is an abelian

group, so the only finite dimensional irreducible representations are one-dimensional.

There are certainly more complicated representations than just sign characters, so

this chapter should be considered only a first foray into the potentially large subject

of p-adic AdS/CFT with spin.

In section 1.6 we claimed that for odd p, there were four distinct sign characters

on Q×p . It will be useful to us to have explicit formulas for each of them.2 First

2Sign characters over Q2 are also well known, but their relation to holographic constructions ismore intricate.

109

Page 116: Real and p-Adic Physics

express any x ∈ Q×p as

x = pvx(x0 + x1p+ x2p2 + . . . ) , (4.3)

where vx ∈ Z, x0 ∈ F×p , and all other xi ∈ Fp. Here Fp is the finite field of p

elements, namely Z/pZ, which we identify with the set 0, 1, 2, . . . , p− 1. And F×p is

the non-zero elements of Fp, which form a multiplicative group. There are two sign

characters on F×p : the trivial one which maps all elements to 1, and the quadratic

residue character n→ (n|p) where (n|p), also denoted(np

), is the Legendre symbol.

It is defined so that (n|p) = 1 if n is a square in F×p and −1 otherwise.3 On Q×p , there

are four choices of sign character:

1. We can map all x ∈ Q×p to 1. This is the trivial character.

2. We can map x→ (x0|p).

3. We can map x→ (−1)vx . This means we assign p itself a sign of −1.

4. We can map x→ (−1)vx(x0|p).

This list exhausts all the sign characters on Q×p . Until we get to section 4.6, we are

going to focus exclusively on the second case: That is, we will hereafter define

sgnx =

(x0

p

). (4.4)

We will narrow our field of inquiry in two other ways. First, we will restrict attention

to nearest neighbor interactions in the bulk, expressible in terms of a classical action

either on Tp or on its line graph L(Tp). This is analogous to restricting to the lowest

non-trivial order in derivatives in Archimedean anti-de Sitter space. Second, the

boundary for us will always be Qp rather than an extension of Qp.

3The definition of the Legendre symbol is traditionally extended to all of Fp by defining (0|p) = 0,and to all integers by first reducing them modulo p. Then (n|p) = 0 or 1 precisely when n is aquadratic residue modulo p.

110

Page 117: Real and p-Adic Physics

The organization of the rest of the chapter is as follows. In section 4.1, we de-

scribe the nearest neighbor actions on Tp and L(Tp) that we will need, both for bosons

and fermions. In order to obtain a factor of sgnx in the final holographic two-point

functions, we need to introduce a non-dynamical U(1) gauge field. Indeed, the factor

of sgn x in (4.1) can be thought of as a Wilson line obtained by integrating the U(1)

gauge field along the shortest path on L(Tp) between the boundary points 0 and x.

The particular gauge field configurations that we need are described in section 4.2.

The main technical steps in extracting the holographic two-point functions are out-

lined in section 4.3, which deals with bulk-to-boundary propagators, and section 4.4,

which recounts the holographic prescription. We detour briefly in section 4.5 into an

account of dynamical gauge fields in the bulk, and then in section 4.6 we summarize

how to modify the gauge fields so as to get any sign character one wants in the final

two-point function (4.1) (for p 6= 2).

4.1 Nearest neighbor actions

In this section we will formulate nearest neighbor actions, first for bosons in sec-

tion 4.1.1 and then for fermions in section 4.1.2. A key ingredient will be a non-

dynamical U(1) gauge field.

4.1.1 Bosonic actions

Starting with a complex-valued function φa on vertices a of a directed graph, we can

define the gradient of φ as the following complex-valued function on edges of the

graph:4

dφe = φt(e) − φs(e) . (4.5)

4Nothing so far privileges complex numbers: φa and dφe could be valued in any linear space V ,and then ωe as used in (4.6)-(4.7) would need to be valued in linear functions on V . For the mostpart we do not need such a general viewpoint.

111

Page 118: Real and p-Adic Physics

a

b

φa

φb

e eiθe Dφe = eiθeφb − φa

Figure 4.1: The gauge covariant derivative Dφe on a small section of a directed

graph.

Here e is an oriented edge with starting point s(e) and terminus t(e).

If we start from a function ωe on directed edges, then we define

dTωa =∑t(e)=a

ωe −∑s(e)=a

ωe , (4.6)

so that ∑e

ωedφe =∑a

(dTωa)φa , (4.7)

possibly up to issues of boundary terms and/or convergence. The equality (4.7)

is an analog of integration by parts. It can be useful to think of d = dea as a

rectangular matrix with one edge-valued index e and one vertex-valued index a. Then,

for example, dφe =∑

a deaφa, and dTωa =∑

e deaωe.

A crucial ingredient in our constructions is a non-dynamical U(1) gauge field.

Because the graph is discrete, instead of a gauge-covariant derivative Dµ = ∂µ + iAµ,

we are going to consider modifying (4.5) to

Dφe = eiθeφt(e) − φs(e) , (4.8)

where θe is essentially∫Aµdx

µ across the edge e. See figure 4.1. Upon gauge trans-

formations

φa → eiλaφa θe → θe − dλe , (4.9)

112

Page 119: Real and p-Adic Physics

we see that

Dφe → eiλs(e)Dφe . (4.10)

Evidently, in the absence of loops, we can use (4.9) with λe = −θe to remove the

phase from (4.8), so that D = d. On L(Tp), there are loops, so non-trivial gauge field

configurations exist.

We now consider the action

Sφ =∑e

|Dφe|2 +∑a

m2|φa|2 . (4.11)

Here and below, | · | acting on a target space field is the norm on C. Varying (4.11)

with respect to φ∗a gives

D†Dφa +m2φa = 0 , (4.12)

where D† is the adjoint of D. (Explicitly, we can write D = Dea as a rectangular

matrix, and then D∗, DT , and D† all have obvious definitions.) A helpful result for

calculations to come is

D†Dφa = oaφa −∑t(e)=a

e−iθeφs(e) −∑s(e)=a

eiθeφt(e) , (4.13)

where oa is the number of edges incident upon a. (On Tp, oa = p+ 1 for all vertices,

while on L(Tp), oa = 2p for all vertices.)

The definition (4.8) might seem asymmetrical, and one might prefer instead eiθe/2φt(e)−

e−iθe/2φs(e) on the right hand side. But for purposes of forming the action (4.11), the

overall phase of Dφe doesn’t matter because only |Dφe|2 enters. In other words, there

is U(1) gauge freedom on edges which we fix in (4.8) by locking the phase of Dφe to

φs(e).

113

Page 120: Real and p-Adic Physics

4.1.2 Fermionic actions

On a graph, the natural notion of a Dirac operator has to do with the exterior deriva-

tive. See for example [79], where the Dirac operator on a graph D is essentially the

signed adjacency matrix on the clique graph of G. We will consider a simplifica-

tion of this general development, in which the only operator we need is the gradi-

ent, rendered gauge covariant as in the previous section. Explicitly, we introduce a

Grassmann-complex-valued function ψa on vertices of a directed graph, and another

such function χe on edges. We define

Dψe = eiθeψt(e) − ψs(e) (4.14)

and introduce the action

Sψ =∑e

[iχ∗eDψe + iχeD∗ψ∗e +mχ∗eχe]−

∑a

Mψ∗aψa . (4.15)

The kinetic terms in (4.15) are constructed in the spirit of b∂c lagrangians, where b

is replaced by an edge field χe and c is replaced by a vertex field ψa. The action is

real once we assume that conjugation exchanges the order of factors. We need ψa and

χe to be complex in order to make the mass terms possible. It appears that the two

mass coefficients are independently meaningful, but in fact there is a global scaling

symmetry ψa → λψa and χe → (λ∗)−1χe, where λ ∈ C is a constant, which preserves

the kinetic terms while rescaling m→ |λ|2m and M → |λ|−2M .

The action (4.15) is invariant under the gauge transformation

ψa → eiλaψa χe → eiλs(e)χe θe → θe − dλe , (4.16)

114

Page 121: Real and p-Adic Physics

and the equations of motion are

iDψe +mχe = 0 iD†χa +Mψa = 0 (4.17)

(and the complex conjugates of these equations).5 From the two equations (4.17) it

follows that

d†dψa +mMψa = 0 . (4.18)

The equivalence of (4.18) and (4.12) is comparable to the way the massive Dirac

equation implies the massive Klein-Gordon equation.

4.2 The background geometries

The non-dynamical U(1) gauge fields on L(Tp) that we are going to study encode the

Legendre symbol (α|p). Consider first the case p ≡ 1 mod 4. Label the vertices of the

complete graph Kp with elements of Fp. Pick a directed structure on Kp, and define

a map

e→ α(e) = t(e)− s(e) (4.19)

from directed edges to F×p . Set eiθe = (α(e)|p) on each edge. Because (α|p) is an

even function of α ∈ F×p , the choice of eiθe doesn’t depend on the directed structure

we picked. Because eiθe is always real, the operator D†D also doesn’t depend on the

directed structure.

Now consider the case p ≡ 3 mod 4. Again label the vertices of Kp with elements

of Fp. Introduce a directed structure on Kp such that an edge runs from a vertex a

to another vertex b iff b − a is a square in F×p . This prescription uniquely specifies

the direction of every edge in Kp because for any α ∈ F×p , either (α|p) = 1 or else

5As in the scalar case, the overall phase of Dψe doesn’t matter because in (4.15) we form theproduct iχ∗eDψe, and we can adjust the phase of χe to keep the overall prefactor equal to i.

115

Page 122: Real and p-Adic Physics

(−α|p) = 1, due to the fact that (α|p) is an odd function of p. Set eiθe = i on all

edges.

We will refer to the directed structures and gauge fields on Kp as Paley construc-

tions, since for p ≡ 1 mod 4 the edges with eiθe = 1 form a Paley graph (without

reference to the directed structure), while for p ≡ 3 mod 4, the directed structure that

we picked forms a Paley digraph.

There are p+1 edges incident upon each vertex A of Tp, of which one edge is below

located below A (that is, one edge lies on the path from A to the boundary point at

infinity), while p edges are located above A. The vertices in L(Tp) corresponding to

the above-lying edges we think of as forming a copy of Kp, and each of these vertices

is also connected to the vertex corresponding to the edge below A. We parametrize

A using a pair (xA, zA) as in section 1.8. Recall that xA ∈ Qp and zA is an integer

power of p, with (xA, zA) identified with (x′A, z′A) iff zA = z′A and |xA − x′A|p ≤ |zA|p.

We will parametrize elements of L(Tp) by using the same coordinates (xA, zA) to label

the vertex of L(Tp) immediately below A; usually we will write instead (xa, za) since

we use lowercase letters to label vertices of L(Tp). To fix a directed structure and

non-dynamical gauge field on L(Tp), we adopt the same Paley construction on each

Kp, and the rest of the edges are directed downward, from (x, z) to (x, z/p), with

eiθe = 1 for p ≡ 1 mod 4 and eiθe = i for p ≡ 3 mod 4. We will refer to edges inside a

copy of Kp as horizontal, and the others as vertical.

If we use a → b to denote a directed edge, then for p ≡ 1 mod 4, our choice of

gauge field is

vertical edges: θ(x,z)→(x, zp)= 0

horizontal edges: θ(x,z)→(x+α zp,z) =

π

2

[1−

p

)]for α ∈ F×p .

(4.20)

In the second line of (4.20), we bear in mind that the directed edge (x, z)→(x+ α z

p, z)

116

Page 123: Real and p-Adic Physics

Figure 4.2: Left: L(T5) in blue and red with T5 shown in dashed green. The blue

edges have θe = 0, while the red edges have θe = π. The pentagram

figures are the Paley constructions, and edges within them are called

horizontal. Their orientation doesn’t matter. The vertical edges are

the ones connecting each pentagram with a vertex of the pentagram

below it, and our convention is for all of them to be directed down-

ward.

Right: L(T3) in black, with T3 shown in dashed green. All the edges

have θe = π/2.

exists only for half the elements of F×p ; but which half doesn’t matter. For p ≡ 3 mod 4,

our choice of gauge field is

vertical edges: θ(x,z)→(x, zp)=π

2

horizontal edges: θ(x,z)→(x+α zp,z) =

π

2for α ∈ F×p with

(αp

)= 1 .

(4.21)

See figure 4.2 for a depiction of small subgraphs of L(Tp) showing also the choice of

gauge fields (4.20) and (4.21) for p = 5 and 3, respectively.

The assignments of eiθe in (4.20) and (4.21) are preserved under the maps (x, z)→

(rx+b, z/|r|p) for all b ∈ Qp and r ∈ (Q×p )2. This is fortunate because the correspond-

ing boundary maps, x→ rx+b, applied to two distinct points x1 and x2 in Qp, are the

ones that preserve the desired two-point function, G(x1, x2) = sgn(x1−x2)/|x1−x2|2∆p ,

up to some power of the scale factor |r|p.

117

Page 124: Real and p-Adic Physics

4.3 Bulk-to-boundary propagators

In order to compute holographic two-point functions, a key ingredient is the bulk-to-

boundary propagator in momentum space. In section 4.3.1 we review the calculation

of this propagator in the case of complex scalars on Tp. Then in section 4.3.2 we work

it out for scalars and fermions on L(Tp).

4.3.1 Scalars on Tp

Consider complex scalars on Tp with action

S =∑E

|dφE|2 +∑A

m2|φA|2 , (4.22)

where vertices are labeled A = (xA, zA) and edges are labeled E. For brevity, let’s

write zA = pv. Then, as recounted in [10], a useful solution of the equations of motion

following from (4.22) is

φA(k) = fvγp(kzA)χ(−kxA) where fv = |zA|1−∆p +Q|k|2∆−1

p |zA|∆p (4.23)

where k ∈ Q×p and

Q = −p1−2∆ . (4.24)

Here χ(ξ) is the additive character on Qp, given explicitly by χ(ξ) = e2πiξ where

ξ is the fractional part of ξ ∈ Qp. Here we have used the p-adic Gaussian γp as

defined in section 1.4. The dimension ∆ is related to the mass by

m2 = − 1

ζp(∆− 1)ζp(−∆). (4.25)

118

Page 125: Real and p-Adic Physics

The solution φA(k) in (4.23) can be thought of as a bulk to boundary propagator

because it is the disturbance of φA in the bulk that corresponds to deforming the

boundary field theory by a term∫dxχ(kx)O(x), where O is the operator dual to

φ.6 From the form of φA(k) given in (4.23) we can pick out the k-dependence of the

Fourier space holographic two-point function for O: G(k) ∝ |k|2∆−1.

4.3.2 Scalars and fermions on the line graph

We would now like to find solutions analogous to (4.23) on L(Tp) with the directed

structures and gauge fields as outlined in section 4.2. The invariance of the back-

ground geometry under translations xa → xa + b indicates that we should be able to

require that fields should depend on xa through a factor χ(−kxa). We immediately

encounter the need to multiply in a factor of γp(kza), because by itself, χ(−kxa) is

not single valued on L(Tp), whereas γp(kza)χ(−kxa) is (and the same logic dictated

that (4.23) must include a factor of γp(kzA)). In short, we are lead to essentially the

same ansatz as (4.23):

φa(k) = uvfvγp(kza)χ(−kxa) where fv = |za|1−∆p +Q|k|2∆−1

p |za|∆p (4.26)

and we have written za = pv. For later convenience, we have introduced the prefactor

uv where

u =

1 for p ≡ 1 mod 4

i for p ≡ 3 mod 4 .

(4.27)

The coefficient Q in (4.26) may depend on k, but we don’t expect it to depend on |k|p

since the explicit factor of |k|2∆−1p already is the dependence we expect for the k-th

Fourier mode of a holographic two-point function with dimension ∆. The aims of

6The field φa, the operator O and the deformation of the conformal field theory action, shouldin the end be real. This can be accomplished by always considering superpositions of Fourier modeswith equal and opposite k.

119

Page 126: Real and p-Adic Physics

the following calculation are to verify that the ansatz (4.26) does solve (4.12) and to

determine ∆ and Q. The strategy is to plug (4.26) into (4.12) and extract a difference

equation for fv. We can handle the fermionic case by replacing φa by ψa in (4.26)

and plugging into (4.18). In the fermionic case, we assume m 6= 0 so that χe can be

determined by the first equation in (4.17).

The factor γp(kza) means that the equation of motion (4.12) is trivially satisfied

for vertices a such that vk + v < −1, where vk ∈ Z is the valuation of k: that is

k = pvk(k0 + k1p+ k2p2 + . . . ) with k0 ∈ F×p . (4.28)

Let’s first show that the equation of motion is also trivially satisfied when v = −vk−1.

In this case, the only non-zero terms in the equation of motion (4.12) are the ones

corresponding to zb = pza, i.e. in the Paley construction above the point a. Explicitly,

the equation of motion reads

−f−vk∑α∈Fp

χ(−k[xa + zaα]) = 0 , (4.29)

where we are using the fact that the p vertices in L(Tp) above (xa, za) are (xa+zaα, pza)

where α runs over Fp. Recalling that v = −vk − 1, we see that the sum in (4.29) is

proportional to ∑α∈Fp

χ(−k0α/p) = 0 . (4.30)

So (4.29) is indeed satisfied trivially and gives us no information about fv.

Let’s move on to the case v > −vk. The factor χ(−kxb) now has the same value

for all vertices b neighboring a, as well as for b = a. Also, γp(kzb) = 1 at all these

vertices. Thus we may discard the factor γp(kza)χ(−kxa) from the ansatz (4.26) and

120

Page 127: Real and p-Adic Physics

work directly with φa = fv. By plugging in to (4.12) we find

(2p+m2)fv − pfv+1 − fv−1 = 0 . (4.31)

The form of (4.31) is the same for p ≡ 1 mod 4 and p ≡ 3 mod 4 because of the

overall prefactor uv in (4.26). One can view this factor as a change of gauge in

the p ≡ 3 mod 4 case which removes the factors of i from the covariant derivatives

along vertical edges while leaving them unchanged within the Paley constructions.

The only property of the Paley constructions we need in order to get (4.31) is that

contributions to D†Dφa from the p − 1 vertices connected to a by a horizontal edge

cancel out. The second order difference equation (4.31) is solved by fv = p−∆v = |za|∆p

and fv = p(∆−1)v = |za|1−∆p where

2p+m2 = p1−∆ + p∆ . (4.32)

We assume that ∆ is real, and the standard prescription is to choose it as the larger

of the two roots of (4.32), so that ∆ > 1/2. Note that 2(√p − p) < m2 < 0 when

∆ ∈ (1/2, logp(p+√p(p− 1))), and it is positive otherwise.7

In the case of fermions, the discussion up to this point proceeds unchanged, except

that m2 is replaced by mM .

To summarize progress so far: We have shown that the ansatz (4.26) trivially

satisfies the equations of motion for v < −vk, while for v > −vk we have shown that

it is consistent with the equations of motion provided we impose the mass-dimension

relation (4.32). But we have no information yet about Q. This information comes

from a boundary condition at v = −vk, and it turns out that it encodes the sign

character that we need in order to obtain two-point functions of the desired form

7Note that ∆ = 12 + is gives m2 real but violating the lower bound m2 > 2(

√p − p). A similar

result was already noted in [11] for scalars on Tp. It is tempting to think that these complex valuesof ∆ correspond to unstable actions, but they may nevertheless have some interesting role to play.

121

Page 128: Real and p-Adic Physics

(4.4). The equation of motion (4.12) for v = −vk reads

(2p+m2)f−vkχ(−kxa)− f−vk+1

∑α∈Fp

χ(−k[xa + zaα])

− uf−vk∑α∈F×p

p

(−k[xa + za

α

p

])= 0 .

(4.33)

There is no f−vk−1 term in (4.33) because the factor of γp(kzb) vanishes when zb =

p−vk−1, so (4.33) is only first order in differences rather than second order. Hence

it can indeed be thought of as a boundary condition for the second order equation

(4.31). In the last term of (4.33), we are using the fact that the p− 1 vertices in the

same Paley construction as xa are(xa + za

αp, za

). This last term is proportional to

the Gauss sum: ∑α∈Fp

p

(−k0α

p

)=

√p

u

(k0

p

). (4.34)

(In (4.34), the α = 0 term in the sum vanishes, so including it is optional. The form

(4.34) makes it clear that we are taking a Fourier transform of the Legendre symbol

over Fp.) Simplifying, and using (4.32), we obtain

[p1−∆ + p∆ −√p

(k0

p

)]f−vk − pf−vk+1 = 0 . (4.35)

Plugging the ansatz for fv in (4.26) into (4.35), one arrives at

[p1−∆ + p∆ −√p

(k0

p

)](1 +Q)− p(p∆−1 + p−∆Q) = 0 , (4.36)

which reduces to

Q = p12−∆

(k0

p

)= p

12−∆ sgn k . (4.37)

122

Page 129: Real and p-Adic Physics

4.4 Two-point functions

With the bulk-to-boundary propagators in hand, we now turn to the computation of

the holographic two-point functions, first in section 4.4.1 for real scalars on Tp and

then in section 4.4.2 for complex scalars on L(Tp), and finally in section 4.4.3 for

fermions on L(Tp).

4.4.1 Scalars on Tp

As a warmup, consider a complex scalar φA on Tp, as in (4.22)-(4.25). Implement a

cutoff by fixing the values of φA for all vertices (xA, zA) with |zA|p = |ε|p, where ε = pvε

and vε is an integer. Let Σε denote the set of vertices with |zA|p > |ε|p, together with

the edges with at least one vertex having |zA|p > |ε|p. Let ∂Σε be the edges with only

one vertex in Σε. We orient edges downward (away from the Qp boundary), so that

when E ∈ ∂Σε, t(E) ∈ Σε and s(E) 6∈ Σε. The vertices in Σε are allowed to fluctuate,

while vertices with |zA|p < |ε|p are ignored. The cutoff action is

Sε =∑E∈Σε

|dφE|2 +∑A∈Σε

m2|φA|2 . (4.38)

We now need an improvement of the partial integration formula (4.7) to include

boundary terms:8

∑E∈Σε

ωEdφE =∑A∈Σε

(dTωA)φA −∑E∈∂Σε

ωEφs(E) . (4.39)

8If ∂Σε included edges for which s(E) ∈ Σε while t(E) 6∈ Σε, then in place of (4.39) we wouldneed

∑E∈Σε

ωEdφE =∑A∈Σε

(dTωA)φA +∑

E∈∂Σεt(E)6∈Σε

ωEφt(E) −∑

E∈∂Σεs(E)6∈Σε

ωEφs(E).

123

Page 130: Real and p-Adic Physics

Using (4.39) we see that

Sε =1

2

∑A∈Σε

[φ∗A(d†dφA +m2φA) + (dTd∗φ∗A +m2φ∗A)φA

]− 1

2

∑E∈∂Σε

[φ∗s(E)dφE + (dφ∗E)φs(E)

].

(4.40)

The first line of (4.40) vanishes on-shell, leaving only the boundary terms. Recalling

that −Son-shellε is the generating function of connected Green’s functions and following

the logic of [78], we see that a cutoff version of the Green’s function can be computed

as

Gε(k) =dφE(k)

φs(E)(k)for any E ∈ ∂Σε , (4.41)

where, crucially, we have plugged in the solution φA = φA(k) from (4.23). We have

to choose |kε|p < 1 in order to avoid having a vanishing denominator in (4.41). As

long as we work at fixed k, this is not a problem, since our eventual aim is to take

ε→ 0 p-adically. Straightforward calculation of the right hand side of (4.41) gives

Gε(k) =fvε−1 − fvε

fvε= − 1

ζp(∆− 1)+ |k|2∆−1

p |ε|2∆−1p

Qp∆

ζp(2∆− 1)+ . . . , (4.42)

where to obtain the first equality we used (4.23), and to obtain the second we ex-

panded in p-adically small ε. The omitted terms go to 0 more quickly than the ones

shown provided ∆ > 1/2, which is true of the larger of the two roots of the relation

(4.25). The first term in (4.42) is k-independent, so in position space it gives rise to a

pure contact term. Dropping this term, we define the Fourier space Green’s function

as

G(k) = limε→0

Gε(k)

|ε|2∆−1p

=Qp∆

ζp(2∆− 1)|k|2∆−1

p . (4.43)

124

Page 131: Real and p-Adic Physics

Plugging in Q = −p1−2∆ from (4.24) and recalling the Fourier transform9

∫Qpdk χ(kx)|k|s =

ζp(1 + s)/ζp(−s)|x|1+s

p

, (4.44)

we obtain

G(x) =p∆ζp(2∆)

ζp(2∆− 1)2

1

|x|2∆p

. (4.45)

A somewhat more involved derivation of (4.45) in [10] makes it clear that the overall

normalization of G(x) is a subtle issue. Changing the location of the cutoff by one

lattice spacing results in changing G(x) by an O(1) multiplicative factor. We should

in short view (4.43) as a reasonable but non-unique prescription for normalizing the

two-point function.

4.4.2 Scalars on the line graph

For complex scalars φa on L(Tp), the extraction of a Green’s function from the bulk

to boundary propagator (4.26) proceeds almost exactly as in the warmup exercise

outlined in the previous section. Formally, in (4.38)-(4.43), one replaces d → D,

A → a, and E → e. Let’s inquire a little more closely why this works. The set

Σε comprises vertices with |za|p > |ε|p and edges with at least one vertex having

|za|p > |ε|p. The boundary ∂Σε consists of vertical edges only, and these edges all

have s(e) 6∈ Σε. Thus the partial integration formula (4.39) can indeed be carried

over to scalars on L(Tp) just by replacing d → D, A → a, and E → e. Likewise,

the subsequent manipulation of the action in (4.40) and the formula (4.41) for the

Green’s function carry over with the same alterations. The calculation (4.42) carries

over unaltered because of our careful inclusion of a factor of uv in the scalar ansatz

(4.26); a more conceptual way to say it is that this factor brings us to a gauge where

D = d on vertical edges. The result (4.43) carries over unaltered, and if we plug in

9An exposition of of Fourier integrals including (4.44) can be found in [27].

125

Page 132: Real and p-Adic Physics

Q = p12−∆ sgn k, as given in (4.37), we obtain

G(k) =

√p sgn k

ζp(2∆− 1)|k|2∆−1

p . (4.46)

Using the Fourier integral

∫Qpdk χ(kx)|k|s sgn k = ups+

12

sgnx

|x|s+1p

, (4.47)

we arrive at

G(x) =up2∆

ζp(2∆− 1)

sgnx

|x|2∆p

. (4.48)

In terms of the boundary field theory, G(x) = 〈Oφ(x)Oφ(0)†〉 whereOφ is the operator

dual to φ. If we assume that translation by x is implemented in the boundary theory

by a unitary operator U(x), then Oψ(x) = U(x)†Oψ(0)U(x), and

G(x)∗ = 〈Oφ(0)Oφ(x)†〉 = 〈Oφ(−x)Oφ(0)†〉 = G(−x) . (4.49)

The relation G(x)∗ = G(−x) is indeed satisfied by (4.48): For p ≡ 1 mod 4, G(x) is

real and even under x → −x, while for p ≡ 3 mod 4, it is imaginary and odd under

x→ −x.

4.4.3 Fermions on the line graph

To derive a holographic two-point function for the operator Oψ dual to ψ, we start

from the cutoff action

Sε =∑e∈Σε

[iχ∗eDψe + iχeD∗ψ∗e +mχ∗eχe]−

∑a∈Σε

Mψ∗aψa , (4.50)

126

Page 133: Real and p-Adic Physics

where Σε is defined as in section 4.1.1. After integration by parts,

Sε =∑e∈Σε

[1

2χ∗e(iDψe +mχe) +

1

2χe(iD

∗ψ∗e −mχ∗e)]

+∑a∈Σε

[1

2ψa(−iDTχ∗a +Mψ∗a) +

1

2ψ∗a(−iD†χa −Mψa)

]−∑e∈∂Σε

[i

2χ∗eψs(e) +

i

2χeψ

∗s(e)

].

(4.51)

This is an off-shell result. On shell, the first two lines vanish. By equating minus the

on-shell action with the generating functional of connected Green’s functions, we can

see by following the logic of [78] that a sensible definition of the cutoff Fourier space

Green’s function Gε(k) is

χe(k) = −iGε(k)ψs(e)(k) , (4.52)

provided Gε(k) turns out to be real. Assuming m 6= 0, we can rewrite (4.52) as

1

mDψe(k) = Gε(k)ψs(e)(k) (4.53)

Plugging in the fermionic bulk-to-boundary propagator (identical to (4.26) with φa →

ψa and Q given by (4.37)), we obtain in place of (4.42) and (4.43) the results

mGε(k) =fvε−1 − fvε

fvε= − 1

ζp(∆− 1)+ |k|2∆−1

p |ε|2∆−1p

Qp∆

ζp(2∆− 1)+ . . .

G(k) = limε→0

Gε(k)

|ε|2∆−1p

=Qp∆

mζp(2∆− 1)|k|2∆−1

p =

√p sgn k

mζp(2∆− 1)|k|2∆−1

p ,

(4.54)

where as usual we dropped the k-independent term from Gε(k) before taking the ε→ 0

limit. The presence of the factor of m in the denominator of the final expression in

127

Page 134: Real and p-Adic Physics

(4.54) makes sense because with a coupling

Sint =

∫Qpdx[ψdef(x)∗Oψ(x) +Oψ(x)†ψdef(x)

](4.55)

between a renormalized version ψdef of the boundary limit of ψ and a boundary

operator Oψ, the scaling symmetry must act as ψdef → λψdef and Oψ → (λ∗)−1Oψ.

So the two-point function G(k) of Oψ with O†ψ scales as G(k) → |λ|−2G(k), which

matches the scaling of 1/m.

In short, we arrive at

G(x) =up2∆

mζp(2∆− 1)

sgnx

|x|2∆p

. (4.56)

It seems at first surprising that the symmetry of G(x) under x→ −x is the same for

bosons and fermions. The reason is that this symmetry was accompanied by complex

conjugation, which reverses the order of operator multiplication without introducing

signs related to the statistics of the operators.

4.5 Gauge field dynamics

So far we have considered only non-dynamical U(1) gauge fields. Let’s now consider

how we might add a kinetic term. On a general directed graph G, define a face f

to be any subgraph of G with three vertices only, all of which are required to be

connected, let’s say by edges e1, e2, and e3. To each face we assign (arbitrarily) a

direction, meaning a direction around which we think of the boundary of the face

circulating. If the direction of edge ei matches the direction assigned to f , then we

set s(f, ei) = 1; otherwise we set s(f, ei) = −1.10 Then, starting from a function θe

10This is of course the next step after (4.5) in constructing the incidence matrix of the cliquegraph; see [79].

128

Page 135: Real and p-Adic Physics

defined on directed edges, we set

dθf =3∑i=1

s(f, ei)θei . (4.57)

Because each θe is essentially a line integral∫Aµdx

µ over the corresponding edge, the

derivative dθf is essentially the integral∫

12Fµνdx

µ ∧ dxν over the face. So an obvious

analog of the Maxwell action is

Sθ =∑f

1

2(dθf )

2 . (4.58)

Given a function Ff on faces, we can immediately read off from (4.57) the adjoint

operator (equivalently, the transpose):

d†Fe =∑∂f3e

s(f, e)Ff , (4.59)

where the sum is over all the faces whose boundary includes e. Evidently—in the

absence of couplings to other fields—the equation of motion following from (4.58) is

d†dθe = 0 . (4.60)

Although (4.58) is indeed the obvious free field action for a U(1) gauge field on a

directed graph G, it is not entirely satisfactory, since we feel that the phases eiθe

rather than the angles θe should be the fundamental variables. To say it another way,

if dθf ∈ 2πZ for some face f , we would like to say that it is equivalent to having

dθf = 0, but that is not reflected by the action (4.58).

The treatment is altogether more natural if we first define a Wilson line. The task

is not much harder for gauge group U(n). Starting therefore with a matter field φ

129

Page 136: Real and p-Adic Physics

mapping vertices to Cn, we can define

Dφe = Ueφt(e) − φs(e) (4.61)

where Ue ∈ U(n). Now consider a directed path γ in our directed graph, from one

vertex, s(γ), to another, t(γ). It is simpler to rule out back-tracking and self-crossing,

but this is not really necessary. For edges e ∈ γ, let s(γ, e) = 1 if the direction of e

matches the overall direction of γ, and −1 otherwise. Now define

Uγ =∏e∈γ

U s(γ,e)e , (4.62)

where the order of factors follows the direction of γ: That is, the first edge in γ

corresponds to the leftmost factor, and the last edge corresponds to the rightmost

factor. The operator Uγ is a Wilson line for the path γ.

Let ∂af be the directed path around a face f , starting and ending on a chosen

vertex a. Then U∂af is well defined, and using it we can form the action

SU = −1

2

∑f

tr(U∂af + U−1∂af

) , (4.63)

in analogy to the Yang-Mills action. The choice of vertex a on the boundary of each

face can be made arbitrarily due to the cyclicity of the trace. To find the equation of

motion, choose an edge e and consider a face f with e ∈ ∂f . One can check that Ue

is the first factor in Us(f,e)∂s(e)f

. Therefore, upon a variation

Ue → (1 + iαH)Ue , (4.64)

where α is a small parameter and H is Hermitian, we find also Us(f,e)∂s(e)f

→ (1 +

130

Page 137: Real and p-Adic Physics

iαH)Us(f,e)∂s(e)f

. We define the variation

δSU =∂SU∂α

∣∣∣∣α=0

= − i2

∑∂f3e

tr(HUs(f,e)∂s(e)f

) + c.c. , (4.65)

where the sum is over all faces whose boundary includes e. In short, the lattice

Yang-Mills equation reads

Im∑∂f3e

tr(HUs(f,e)∂s(e)f

) = 0 for all hermitian H . (4.66)

Returning to the U(1) case, we can simplify the equations of motion (4.66) to

Im∑∂f3e

Us(f,e)∂f = 0 , (4.67)

where we omit to specify the starting and ending point of ∂f because in an abelian

theory it doesn’t matter. We wish to consider a stronger condition, which we will

refer to as the complexified equation of motion:

∑∂f3e

Us(f,e)∂f = 0 . (4.68)

Imposing the condition (4.68) clearly implies the equations of motion, but the reverse

is in general not true. This is reminiscent of the situation in continuum Yang-Mills

theory where self-duality of the field strength implies the equations of motion, but

not vice versa. The point of interest for us is that the gauge field configurations we

have used throughout do satisfy (4.68). To verify this, we need the identity

∑y∈Fp

(x− yp

)(y

p

)= (−1)

p+12 , (4.69)

where x ∈ F×p is arbitrary and p is an odd prime. This convolution identity is known

as a Jacobi sum, and it can be proven starting from the Gauss sum (4.34). In the

131

Page 138: Real and p-Adic Physics

following two paragraphs, we explain how (4.68) reduces to (4.69), as well as giving

some further indications of what (4.68) means physically.

For p ≡ 1 mod 4, the claim (4.68) comes down to the claim that U∂f = +1 for

half of the p − 1 faces adjoining a given edge e, and −1 for the other half. Before

we check this claim, let’s note that since all the Ue are real, the lattice Yang-Mills

equation (4.67) is satisfied trivially. It is not even clear that we should be demanding

(4.67) if we restrict the gauge fields to Ue = ±1, since (4.67) was derived on the

assumption that we could make an infinitesimal change to the gauge fields. However,

when all Ue = ±1, one can derive (4.68) as the condition that the action should

remain unaltered when just one of the Ue flips its sign. In any case, checking (4.68) is

trivial when e is a vertical edge, because it comes down to the observation that half

of the horizontal edges from any given vertex have sign +1 while the other have sign

−1. When e is a horizontal edge, say from 0 to x ∈ F×p , then we may write down

(4.68) explicitly as

(x

p

)+∑y∈F×py 6=x

(x

p

)(y − xp

)(−yp

)= 0 .

(4.70)

The first term in (4.70) comes from a face with two vertical edges plus the chosen

horizontal edge e, while the other p−2 terms come from faces with all edges horizontal.

The sum (4.70) obviously reduces to (4.69).

For p ≡ 3 mod 4, because all the Ue = i, one finds that U∂f is pure imaginary for all

faces. So in this case, (4.68) is equivalent to (4.67), and we can think of it as following

from requiring that the action is stationary under infinitesimal variations. We do not

see a useful way to understand how (4.68) arises from flipping the orientation of one

edge while preserving the property that Ue = i, since any configuration in which all

the U∂f are pure imaginary automatically has vanishing action. Once again, checking

(4.68) is trivial when e is a vertical edge because Us(f,e)∂f = i when the horizontal edge

132

Page 139: Real and p-Adic Physics

in f points toward e and −i when it points away. When e is a horizontal edge, say

from 0 to x ∈ F×p , then (4.68) becomes

i

(x

p

)− i

∑y∈F×py 6=x

(x

p

)(y − xp

)(−yp

)= 0 ,

(4.71)

and this again reduces to (4.69).

4.6 Other sign characters

Let’s now inquire how we might modify our constructions on L(Tp) so as to get the

other three sign characters listed in section 4. As recounted in section 1.6, we can

identify each sign character with some non-square τ ∈ Q×p :

sgnτ x =

1 if x = a2 − τb2 for some a, b ∈ Qp

−1 otherwise .

(4.72)

Indeed, sgnτ = sgnτ ′ if τ/τ ′ is a square. There are therefore essentially only four

choices for τ : 1, ε, p, and pε, where ε is any element of F×p with (ε|p) = −1. The

characters sgnτ can be explicitly evaluated, as follows:

p ≡ 1 mod 4 p ≡ 3 mod 4

sgn1 x = 1 sgn1 x = 1

sgnε x = (−1)vx sgnε x = (−1)vx

sgnp x = (x0|p) sgnp x = (−1)vx(x0|p)

sgnεp x = (−1)vx(x0|p) sgnεp x = (x0|p)

(4.73)

where x = pvx(x0 + x1p+ x2p2 + . . . ) ∈ Q×p as in (4.3), where x0 ∈ F×p .

133

Page 140: Real and p-Adic Physics

A useful intuition in the way a sign character arises is that a particle picks up a

phase Ue as it propagates across a link e. Because all vertical edges have the same

downward orientation and the same Ue, any phase that a particle picks up while

moving down from the boundary is precisely undone on its way back up. However,

on the shortest possible path on L(Tp) from one boundary point to Qp to another,

there is one horizontal edge. In a correlator 〈Oφ(x)Oφ(0)†〉, this horizontal edge runs

from (0, pvx+1) to (x0pvx , pvx+1). So it makes sense that 〈Oφ(x)Oφ(0)†〉 should include

a factor of

Ux ≡

U(0,pvx+1)→(x0pvx ,pvx+1) if (0, pvx+1)→ (x0p

vx , pvx+1) exists as an edge

U−1(x0pvx ,pvx+1)→(0,pvx+1) otherwise .

(4.74)

The factor (4.74) can be understood as a Wilson line for the path from 0 to x—where

all the factors for vertical edges canceled out. The choice of non-dynamical U(1)

gauge fields in section 4.2 can be understood as a way to get Ux = u sgnx—where the

factor of u is forced on us by the hermiticity condition U−x = U∗x . It is not entirely

clear from this discussion that the final Green’s function G(x) is proportional to Ux:

For instance, other paths exist on L(Tp) from 0 to x which do not pass through the

horizontal edge in question. However, it is a good guess that we can get the other

sign characters in (4.73) by setting

Ux = uτ sgnτ x , (4.75)

where uτ is an x-independent constant—either 1 or i—chosen so as to preserve the

condition U−x = U∗x . If we require (4.75) for all x ∈ Qp, and further require that

phases should be invariant under translations, then (4.75) amounts to a specification

of Ue for all horizontal edges. See figure 4.3 for example. As we have seen in the

discussion following (4.26), we can gauge away the phase on vertical edges.

134

Page 141: Real and p-Adic Physics

Figure 4.3: Left: L(T5) in blue and red. The blue edges have θe = 0, while

the red edges have θe = π. This geometry yields the sign character

(x0|5)(−1)vx .

Center: L(T3) in black. All the edges have θe = π/2. This geometry

yields the sign character (x0|3)(−1)vx .

Right: L(T3) in blue and red. The blue edges have θe = 0, while

the red edges have θe = π. This geometry yields the sign character

(−1)vx .

135

Page 142: Real and p-Adic Physics

The calculations of previous sections are for the most part straightforward to gen-

eralize to arbitrary sign characters, using the prescription (4.75) for choosing phases

on the horizontal edges of L(Tp). For an operator dual to a complex scalar on L(Tp)

with action (4.11), we find

G(x) = Csgnτ x

|x|2∆p

, (4.76)

up to divergent contact terms proportional to δ(x). We need uτ = 1 except when

p ≡ 3 mod 4 and τ = p or εp, and then uτ = i is required. The values of C and m2

can be determined from the following table:

τ C mass-dimension relation

1 p∆ζp(2∆)

ζp(2∆−1)2 m2 = − 1ζp(∆−1)ζp(−∆)

ε − 1+p2∆−2p−m2

1+p2−2∆−2p−m2

p∆ζp(4∆)

ζp(2∆)ζp(4∆−2)(2p+m2)2 = (1 + p2∆)(1 + p2−2∆)

p up2∆

ζp(2∆−1)2p+m2 = p1−∆ + p∆

εp up2∆

ζp(2∆−1)2p+m2 = p1−∆ + p∆

(4.77)

where u = 1 for p ≡ 1 mod 4 and u = i for p ≡ 3 mod 4. In all cases, we choose the

larger of the two possible values of ∆, assumed to be real.

The case τ = 1 is essentially the same as a complex scalar on Tp. The cases τ = p

and τ = εp are similar to one another, and we already presented in detail the cases

τ = p for p ≡ 1 mod 4 and τ = εp for p ≡ 3 mod 4.

Let us therefore focus on the one case with some new features: τ = ε, i.e. sgnε x =

(−1)vx . For simplicity, we consider only the complex scalar. The treatment proceeds

much as in sections 4.3.2 and 4.4.2, with a scalar ansatz

φa = fvγp(kza)χ(−kxa) . (4.78)

136

Page 143: Real and p-Adic Physics

We recall the notation za = pv, and we use vk as before to denote the valuation of k.

For v > −vk, by plugging (4.78) into the scalar equation of motion (4.12), we obtain

the difference equation

(2p+m2)fv − pfv+1 − fv−1 + (−1)v(p− 1)fv = 0 . (4.79)

The last term comes from the horizontal edges. The equation (4.79) is different from

all previous difference equations we’ve encountered in that it does not have constant

coefficients, but instead coefficients that are periodic modulo 2 in v.11 Up to an overall

multiplicative scaling, the general solution to (4.79) is

fv = (1 + q(−1)v)(|za|1−∆

p +Q|k|2∆−1p |za|∆p

), (4.80)

where Q is a coefficient which at this stage is undetermined. Plugging (4.80) into

(4.79), one finds

(2p+m2)2 = (1 + p2∆)(1 + p2−2∆)

q =p∆ + p1−∆ − 2p−m2

p− 1,

(4.81)

and we assume as usual that ∆ > 1/2 is real.

For v < −vk, the equation of motion (4.12) is satisfied trivially, so the boundary

condition that determines Q comes from v = −vk, where (4.12) reads

(2p+m2)f−vk − pf−vk+1 + (−1)vkf−vk∑α∈F×p

χ

(−kza

α

p

)= 0 . (4.82)

Using the obvious identity

∑α∈F×p

χ

(−kza

α

p

)= −1 , (4.83)

11For the sign characters (−1)vx(x0|p), the difference equation obtained for v > −vk has constantcoefficients because the terms from horizontal edges cancel out.

137

Page 144: Real and p-Adic Physics

We arrive at [2p+m2 − (−1)vk

]f−vk − pf−vk+1 = 0 . (4.84)

Plugging in (4.80) and using (4.81), we obtain

Q = (−1)vk+1p1−2∆ 1 + p2∆ − 2p−m2

1 + p2−2∆ − 2p−m2. (4.85)

To compute the holographic Green’s function, if we start with (4.42), we obtain

Gε(k) =1− q(−1)vε

1 + q(−1)vε

[p1−∆ + |k|2∆−1|ε|2∆−1 Qp∆

ζp(2∆− 1)+ . . .

]− 1 , (4.86)

where the omitted terms scale to 0 more quickly than the ones shown as ε→ 0 in the

p-adic norm. We now define

G(k) = limε→0

(1− q(−1)vε

1 + q(−1)vεGε(k)

|ε|2∆−1p

)=

Qp∆

ζp(2∆− 1)|k|2∆−1

p , (4.87)

where we dropped a k-independent term from Gε(k) before taking the limit. As

compared to (4.43), the definition (4.87) may seem a bit contrived. However, the extra

prefactor 1−q(−1)vε

1+q(−1)vεin (4.87) has no k-dependence, and its geometric mean between

even and odd vε is 1. So we maintain that (4.87) is the most sensible way to normalize

the Green’s function. Passing through a Fourier transform, we wind up with

G(x) = − 1 + p2∆ − 2p−m2

1 + p2−2∆ − 2p−m2

p∆ζp(4∆)

ζp(2∆)ζp(4∆− 2)

sgnε x

|x|2∆p

. (4.88)

4.7 Outlook

The main results of this chapter are summarized in (4.75)-(4.77): With a suitably

chosen configuration of a non-dynamical U(1) gauge field on the line graph L(Tp) of

the Bruhat-Tits tree, one may recover from a bulk complex scalar action on L(Tp) a

138

Page 145: Real and p-Adic Physics

holographic boundary two-point function proportional to any sign character one wants

over Qp (for any odd prime p). Equally, one can work with a bulk complex fermion.

The key is not the statistics of the gauge field (or the boundary operator), but rather

the bulk gauge field, which gives rise to the desired sign character essentially as a

Wilson line between two boundary points.

While technical in nature, the results of this chapter open up many further ques-

tions. To begin with, one could try to tackle the case p = 2. There are seven

non-trivial sign characters sgnτ x over Q2 (besides the trivial one), and they depend

not only on the 2-adic norm of x, but also on its second and third non-trivial 2-adic

digits. Preliminary indications are that one can engineer elaborations of T2, including

non-dynamical gauge fields, that allow one to recover these sign characters; however,

nearest neighbor interactions are not enough. This is not too surprising given that the

second and third 2-adic digits relate to paths on the tree with at least two or three

links. Perhaps this is a hint for how to go on to more complicated multiplicative

characters over Qp, which can depend on finitely many p-adic digits and will wind up

involving finite range interactions on Tp.

It is a bit unsatisfying that we have placed so few limitations on the types of

fields that are allowed on Tp. In particular, if on L(Tp) we give each horizontal edge a

phase θe = π and leave all vertical edges with phases θe = 0, then one winds up with a

correlator of the form (4.76) with the trivial sign character, and the only effect of the

phases is to switch the sign of C. One naturally asks if there is a positivity constraint

that can fix this sign. One could observe that the choice of phases just described fails

to satisfied the complexified equation of motion (4.68); however, the same critique

can be made of the choice we made to capture the character sgnε x = (−1)vx . A more

fundamental point of view is called for explaining why particular choices of the U(1)

gauge field are natural constructions on L(Tp), perhaps analogous to the way that

the spin connection is natural on a smooth manifold.

139

Page 146: Real and p-Adic Physics

Perhaps a related point is that we seem to have broken a lot of the symmetry of

Tp by introducing the distinction between horizontal and vertical edges on L(Tp). At

one level, this is not so disturbing, because the invariance we should require is dual

to the maps x → rx + b for all b ∈ Qp and r ∈ (Q×p )2. From a boundary theory

point of view, the operators Oφ and their two-point Green’s function transform with

non-trivial Jacobians under other elements of the p-adic conformal group PGL(2,Qp).

It would be very satisfying to give a full account in the bulk of how the corresponding

isometries of Tp (and L(Tp)) act on the U(1) gauge fields so as produce holographic

Green’s functions which are suitably covariant under the p-adic conformal group.

Having introduced the possibility of gauge field dynamics on L(Tp), another nat-

ural direction to explore is what the corresponding boundary operators are. If some

notion of conserved currents on the boundary is understood, perhaps the gravitational

dynamics of [25] could be refined or extended. We also hope that an enriched under-

standing of the geometry dual to p-adic conformal field theories will eventually impact

back on p-adic string theory, perhaps providing a better first-principles understanding

of Freund and Olson’s adaptation of the Veneziano amplitude and suggesting some

interesting generalizations.

140

Page 147: Real and p-Adic Physics

References

[1] S. S. Gubser, C. Jepsen, Z. Ji, and B. Trundy, “Continuum limits of sparse

coupling patterns,” Phys. Rev. D 98 (2018), no. 4 045009, 1805.07637.

[2] G. Bentsen, T. Hashizume, A. S. Buyskikh, E. J. Davis, A. J. Daley, S. S.

Gubser, and M. Schleier-Smith, “Treelike Interactions and Fast Scrambling

with Cold Atoms,” Physical Review Letters 123 (Sep, 2019).

[3] S. S. Gubser, C. Jepsen, Z. Ji, and B. Trundy, “Higher melonic theories,”

JHEP 09 (2018) 049, 1806.04800.

[4] S. S. Gubser, C. Jepsen, and B. Trundy, “Spin in p-adic AdS/CFT,” J. Phys. A

52 (2019), no. 14 144004, 1811.02538.

[5] G. Veneziano, “Construction of a crossing-symmetric, Regge behaved

amplitude for linearly rising trajectories,” Nuovo Cimento 57 (1968) 190–197.

[6] P. G. O. Freund and E. Witten, “Adelic string amplitudes,” Phys. Lett. B199

(1987) 191.

[7] P. G. O. Freund and M. Olson, “Nonarchimedean strings,” Phys. Lett. B199

(1987) 186.

[8] I. V. Volovich, “p-adic string,” Classical and Quantum Gravity 4 (1987), no. 4

L83.

141

Page 148: Real and p-Adic Physics

[9] L. Brekke, P. G. O. Freund, M. Olson, and E. Witten, “Non-archimedean string

dynamics,” Nucl. Phys. B302 (1988) 365.

[10] S. S. Gubser, J. Knaute, S. Parikh, A. Samberg, and P. Witaszczyk, “p-adic

AdS/CFT,” Commun. Math. Phys. 352 (2016), no. 3 1019–1059, 1605.01061.

[11] M. Heydeman, M. Marcolli, I. Saberi, and B. Stoica, “Tensor networks, p-adic

fields, and algebraic curves: arithmetic and the AdS3/CFT2 correspondence,”

1605.07639.

[12] J. Maldacena International Journal of Theoretical Physics 38 (1999), no. 4

1113–1133.

[13] E. Witten, “Anti-de Sitter space and holography,” Adv. Theor. Math. Phys. 2

(1998) 253–291, hep-th/9802150.

[14] S. S. Gubser, I. R. Klebanov, and A. M. Polyakov, “Gauge theory correlators

from noncritical string theory,” Phys. Lett. B 428 (1998) 105–114,

hep-th/9802109.

[15] A. L. Fitzpatrick, J. Kaplan, J. Penedones, S. Raju, and B. C. van Rees, “A

natural language for AdS/CFT correlators,” Journal of High Energy Physics

2011 (Nov, 2011).

[16] M. F. Paulos, “Towards Feynman rules for Mellin amplitudes in AdS/CFT,”

Journal of High Energy Physics 2011 (Oct, 2011).

[17] G. Mack, “D-independent representation of Conformal Field Theories in D

dimensions via transformation to auxiliary Dual Resonance Models. Scalar

amplitudes,” 2009.

[18] C. B. Jepsen and S. Parikh, “p-adic Mellin amplitudes,” 1808.08333.

142

Page 149: Real and p-Adic Physics

[19] C. B. Jepsen and S. Parikh, “Recursion relations in p-adic Mellin space,”

Journal of Physics A: Mathematical and Theoretical 52 (Jun, 2019) 285401.

[20] B. Swingle, “Entanglement renormalization and holography,” Physical Review

D 86 (Sep, 2012).

[21] F. Pastawski, B. Yoshida, D. Harlow, and J. Preskill, “Holographic quantum

error-correcting codes: toy models for the bulk/boundary correspondence,”

Journal of High Energy Physics 2015 (Jun, 2015).

[22] C.-L. Hung, A. Gonzalez-Tudela, J. I. Cirac, and H. J. Kimble, “Quantum spin

dynamics with pairwise-tunable, long-range interactions,” Proceedings of the

National Academy of Science 113 (Aug., 2016) E4946–E4955, 1603.05860.

[23] S. Ryu and T. Takayanagi, “Holographic Derivation of Entanglement Entropy

from the anti–de Sitter Space/Conformal Field Theory Correspondence,”

Physical Review Letters 96 (May, 2006).

[24] M. Heydeman, M. Marcolli, S. Parikh, and I. Saberi, “Nonarchimedean

Holographic Entropy from Networks of Perfect Tensors,” 2018.

[25] S. S. Gubser, M. Heydeman, C. Jepsen, M. Marcolli, S. Parikh, I. Saberi,

B. Stoica, and B. Trundy, “Edge length dynamics on graphs with applications

to p-adic AdS/CFT,” JHEP 06 (2017) 157, 1612.09580.

[26] S. S. Gubser, C. Jepsen, S. Parikh, and B. Trundy, “O(N) and O(N) and

O(N),” 1703.04202.

[27] S. S. Gubser, M. Heydeman, C. Jepsen, S. Parikh, I. Saberi, B. Stoica, and

B. Trundy, “Signs of the time: Melonic theories over diverse number systems,”

1707.01087.

143

Page 150: Real and p-Adic Physics

[28] I. Ya. Arefeva, B. G. Dragovic, and I. V. Volovich, “p-adic superstrings,” Phys.

Lett. B214 (1988) 339.

[29] I. Ya. Arefeva, B. G. Dragovic, and I. V. Volovich, “Open and Closed p-adic

Strings and Quadratic Extensions of Number Fields,” Phys. Lett. B212 (1988)

283–291.

[30] P. Ruelle, E. Thiran, D. Verstegen, and J. Weyers, “Adelic String and

Superstring Amplitudes,” Mod. Phys. Lett. A4 (1989) 1745.

[31] A. V. Marshakov and A. V. Zabrodin, “New p-adic string amplitudes,” Mod.

Phys. Lett. A5 (1990) 265.

[32] E. Yu. Lerner and M. D. Missarov, “Renormalization group in a fermionic

hierarchical model,” Theor. Math. Phys. 101 (1994) 1353–1360.

[33] F. Q. Gouvea, p-adic Numbers. Springer, 1997.

[34] L. Brekke and P. G. Freund, “p-adic numbers in physics,” Physics Reports 233

(1993), no. 1 1–66.

[35] V. Vladimirov, I. Volovich, and E. Zelenov, p-adic Analysis and Mathematical

Physics. Series on Soviet and East European mathematics. World Scientific,

1994.

[36] P. J. Sally, “An Introduction to p-adic Fields, Harmonic Analysis and the

Representation Theory of SL(2),” Letters in Mathematical Physics 46 (Oct,

1998) 1–47.

[37] J. T. Tate, Fourier analysis in number fields and Hecke’s zeta functions. PhD

thesis, Princeton University, 1950.

[38] I. M. Gel’fand, M. I. Graev, and I. Piatetski-Shapiro, Representation theory

and automorphic functions. Saunders, 1968.

144

Page 151: Real and p-Adic Physics

[39] Y. I. Manin, “p-Adic automorphic functions,” Sovr. Probl. Mat. 3 (1974) 5–92.

[40] E. Y. Lerner and M. D. Missarov, “p-Adic conformal invariance and the

Bruhat-Tits tree,” Lett. Math. Phys. 22 (1991) 123–129.

[41] E. Melzer, “Nonarchimedean Conformal Field Theories,” International Journal

of Modern Physics A 04 (1989), no. 18 4877–4908.

[42] A. V. Zabrodin, “Nonarchimedean Strings and Bruhat-Tits Trees,” Commun.

Math. Phys. 123 (1989) 463–483.

[43] S. S. Gubser and S. Parikh, “Geodesic bulk diagrams on the Bruhat–Tits tree,”

Physical Review D 96 (Sep, 2017).

[44] F. J. Dyson, “Existence of a phase transition in a one-dimensional Ising

ferromagnet,” Commun. Math. Phys. 12 (1969) 91–107.

[45] P. M. Bleher and J. G. Sinai, “Investigation of the critical point in models of

the type of Dyson’s hierarchical models,” Comm. Math. Phys. 33 (1973), no. 1

23–42.

[46] E. Yu. Lerner and M. D. Missarov, “Scalar Models of p-adic Quantum Field

Theory and Hierarchical Models,” Theor. Math. Phys. 78 (1989) 177–184.

[47] G. Bentsen, E. Davis, and M. Schleier-Smith, private discussions.

[48] V. Bonzom, R. Gurau, A. Riello, and V. Rivasseau, “Critical behavior of

colored tensor models in the large N limit,” Nucl. Phys. B853 (2011) 174–195,

1105.3122.

[49] R. Gurau and J. P. Ryan, “Colored Tensor Models - a review,” SIGMA 8

(2012) 020, 1109.4812.

145

Page 152: Real and p-Adic Physics

[50] S. Carrozza and A. Tanasa, “O(N) Random Tensor Models,” Lett. Math. Phys.

106 (2016), no. 11 1531–1559, 1512.06718.

[51] E. Witten, “An SYK-Like Model Without Disorder,” 1610.09758.

[52] I. R. Klebanov and G. Tarnopolsky, “Uncolored random tensors, melon

diagrams, and the Sachdev-Ye-Kitaev models,” Phys. Rev. D95 (2017), no. 4

046004, 1611.08915.

[53] S. Sachdev and J. Ye, “Gapless spin fluid ground state in a random, quantum

Heisenberg magnet,” Phys. Rev. Lett. 70 (1993) 3339, cond-mat/9212030.

[54] A. Kitaev. “A simple model of quantum holography.” Talks at KITP, April 7,

2015 and May 27, 2015.

[55] P. Narayan and J. Yoon, “SYK-like Tensor Models on the Lattice,” JHEP 08

(2017) 083, 1705.01554.

[56] S. Choudhury, A. Dey, I. Halder, L. Janagal, S. Minwalla, and R. Poojary,

“Notes on Melonic O(N)q−1 Tensor Models,” 1707.09352.

[57] F. Ferrari, V. Rivasseau, and G. Valette, “A New Large N Expansion for

General Matrix-Tensor Models,” 1709.07366.

[58] G. Tarnopolsky, “On large q expansion in the Sachdev-Ye-Kitaev model,”

1801.06871.

[59] C. Jepsen, “Adelic Tensor Models,” Princeton University Pre-thesis (2017).

[60] L. E. Dickson and F. H. Safford, “Solution to problem 8 (group theory),”

Amer. Math. Monthly 13 (1906) 150–151.

[61] E. N. Gelling, “On one-factorizations of a complete graph and the relationship

to round-robin schedules,” 1973. Ph.D. Thesis, Univ. Victoria, Canada.

146

Page 153: Real and p-Adic Physics

[62] E. N. Gelling and R. E. Odeh, “On 1-factorizations of a complete graph and

the relationship to round-robin schedules,” Congressus Numerantium 9 (1974)

213–221.

[63] J. H. Dinitz and D. R. Stinson, “A hill-climbing algorithm for the construction

of one-factorizations and Room squares,” SIAM Journal on Algebraic Discrete

Methods 8 (1987), no. 3 430–438.

[64] E. Seah and D. R. Stinson, “On the enumeration of one-factorizations of

complete graphs containing prescribed automorphism groups,” Mathematics of

computation 50 (1988), no. 182 607–618.

[65] J. H. Dinitz, D. K. Garnick, and B. D. McKay, “There are 526,915,620

nonisomorphic one-factorizations of K12,” Journal of Combinatorial Designs 2

(1994), no. 4 273–285.

[66] P. Kaski and P. R. Ostergard, “There are 1,132,835,421,602,062,347

nonisomorphic one-factorizations of K14,” Journal of Combinatorial Designs

17 (2009), no. 2 147–159.

[67] K. Bulycheva, I. R. Klebanov, A. Milekhin, and G. Tarnopolsky, “Spectra of

Operators in Large N Tensor Models,” Phys. Rev. D97 (2018), no. 2 026016,

1707.09347.

[68] I. R. Klebanov, A. Milekhin, F. Popov, and G. Tarnopolsky, “On the Spectra of

Eigenstates in Fermionic Tensor Quantum Mechanics,” 1802.10263.

[69] P. J. Cameron, Parallelisms of Complete Designs. London Mathematical

Society Lecture Note Series. Cambridge University Press, 1976.

[70] J. Polchinski and V. Rosenhaus, “The Spectrum in the Sachdev-Ye-Kitaev

Model,” JHEP 04 (2016) 001, 1601.06768.

147

Page 154: Real and p-Adic Physics

[71] J. Maldacena and D. Stanford, “Remarks on the Sachdev-Ye-Kitaev model,”

Phys. Rev. D94 (2016), no. 10 106002, 1604.07818.

[72] S. R. Das, A. Dhar, A. M. Sengupta, and S. R. Wadia, “New Critical Behavior

in d = 0 Large N Matrix Models,” Mod. Phys. Lett. A5 (1990) 1041–1056.

[73] F. Sugino and O. Tsuchiya, “Critical behavior in c = 1 matrix model with

branching interactions,” Mod. Phys. Lett. A9 (1994) 3149–3162,

hep-th/9403089.

[74] S. S. Gubser and I. R. Klebanov, “A Modified c = 1 matrix model with new

critical behavior,” Phys. Lett. B340 (1994) 35–42, hep-th/9407014.

[75] I. R. Klebanov and A. Hashimoto, “Nonperturbative solution of matrix models

modified by trace squared terms,” Nucl. Phys. B434 (1995) 264–282,

hep-th/9409064.

[76] F. Qu and Y.-h. Gao, “Scalar fields on pAdS,” Phys. Lett. B786 (2018)

165–170, 1806.07035.

[77] D. Harlow, S. H. Shenker, D. Stanford, and L. Susskind, “Tree-like structure of

eternal inflation: A solvable model,” Phys. Rev. D85 (2012) 063516,

1110.0496.

[78] N. Iqbal and H. Liu, “Real-time response in AdS/CFT with application to

spinors,” Fortsch. Phys. 57 (2009) 367–384, 0903.2596.

[79] O. Knill, “The Dirac operator of a graph,” 1306.2166.

148