1 Introduction
A fundamental problem in spectral graph theory is the classification and characterization of graphs with bounded eigenvalues. When we talk about eigenvalues of a graph we always refer to its adjacency matrix. In this paper, we study the families of graphs with eigenvalues bounded from below. Let
$\mathcal {G}(\lambda )$
be the family of graphs with smallest eigenvalue at least
$-\lambda $
. For the sake of comparison, we mention the family
$\mathcal {G}'(\lambda )$
of graphs with spectral radius (or, equivalently, largest eigenvalue) at most
$\lambda $
.
Remark. Since we rarely work with subgraphs that are not induced, all subgraphs are induced throughout this paper. We refer to subgraphs that are not necessarily induced as general subgraphs.
The Cauchy interlacing theorem implies that both
$\mathcal {G}'(\lambda )$
and
$\mathcal {G}(\lambda )$
are closed under taking subgraphs. It is a natural question to ask whether it is possible to define each of these families by a finite set of forbidden subgraphs.
Definition 1.1. Given a family
$\mathcal {G}$
of graphs that is closed under taking subgraphs, a family
$\mathcal {F}$
of graphs is a forbidden subgraph characterization of
$\mathcal {G}$
if the family
$\mathcal {G}$
consists exactly of graphs that do not contain any member of
$\mathcal {F}$
as a subgraph, and a graph F is a minimal forbidden subgraph for
$\mathcal {G}$
if F itself is not in
$\mathcal {G}$
but every proper subgraph of F is in
$\mathcal {G}$
.
Note that the most economical forbidden subgraph characterization of
$\mathcal {G}$
consists precisely of the minimal forbidden subgraphs for
$\mathcal {G}$
. Thus the existence of a finite forbidden subgraph characterization of
$\mathcal {G}$
is equivalent to the finiteness of the minimal forbidden subgraphs for
$\mathcal {G}$
. In 1992 Bussemaker and Neumaier made the following remark in [Reference Bussemaker and Neumaier6, p 599]:
It would be interesting to know the set of numbers m,
$-m$
such that
$\mathcal {G}_m^\#$
[the set of minimal forbidden subgraphs for
$\mathcal {G}'(m)$
] or
$\mathcal {G}_{-m}^\#$
[the set of minimal forbidden subgraphs for
$\mathcal {G}(m)$
] are finite; however, these seem to be very difficult problems.
The specific families
$\mathcal {G}'(2)$
and
$\mathcal {G}(2)$
are well understood. One of the earliest results dates back to 1970 when Smith [Reference Smith29] determined all the connected graphs in
$\mathcal {G}'(2)$
– they are general subgraphs of the extended Dynkin diagrams in Figure 1. The family
$\mathcal {G}(2)$
is much richer and more complex – it contains not only all the graphs in
$\mathcal {G}'(2)$
, but also all the line graphs.Footnote
1
The classification of
$\mathcal {G}(2)$
culminated in a beautiful theorem of Cameron, Goethals, Seidel, and Shult [Reference Cameron, Goethals, Seidel and Shult7] who related
$\mathcal {G}(2)$
to root systems that occur in the classification of semisimple Lie algebras (see Theorem 2.17). We refer the reader to the monograph [Reference Cvetković, Rowlinson and Simić12] for a comprehensive account of
$\mathcal {G}(2)$
.
These classification theorems can be used to establish quantitative answers to Bussemaker and Neumaier’s problems. For
$\mathcal {G}'(2)$
, Cvetković, Doob, and Gutman [Reference Cvetković, Doob and Gutman9, Theorem 2.8] determined that there are
$18$
minimal forbidden subgraphs. For
$\mathcal {G}(2)$
, Rao, Singhi, and Vijayan [Reference Rao, Singhi and Vijayan26, Theorem 4.1] observed that the number of vertices in any minimal forbidden subgraph is at most
$37$
, which was eventually perfected to
$10$
by Kumar, Rao, and Singhi [Reference Kumar, Rao and Singhi23]. A computer search by Bussemaker and Neumaier [Reference Bussemaker and Neumaier6, p 596] established that there are, in total,
$1,812$
minimal forbidden subgraphs for
$\mathcal {G}(2)$
.

Figure 1 Maximal connected graphs with spectral radius at most
$2$
. The number of vertices is one more than the given index. In particular,
$\widetilde {D}_4$
is actually a star with four leaves.
In a recent work, the authors of the current paper resolved the first problem of Bussemaker and Neumaier.Footnote 2
Theorem 1.2 (Theorem 1 of Jiang and Polyanskii [Reference Jiang and Polyanskii20]).
For every integer
$m \ge 2$
, let
$\beta _m$
be the largest root of
$x^{m+1} = 1 + x + \dots + x^{m-1}$
, and let
$\alpha _m := \beta _m^{1/2} + \beta _m^{-1/2}$
. The family
$\mathcal {G}'(\lambda )$
of graphs with spectral radius at most
$\lambda $
has a finite forbidden subgraph characterization if and only if
$\lambda < \lambda '$
and
$\lambda \not \in \left \{{\alpha _2, \alpha _3, \dots }\right \}$
, where

and
$\varphi $
is the golden ratio
$(1+\sqrt 5)/2$
.
The first main theorem of this paper resolves the remaining problem of Bussemaker and Neumaier –
$\mathcal {G}(\lambda )$
enjoys a straightforward threshold phenomenon.
Theorem 1.3. The family
$\mathcal {G}(\lambda )$
of graphs with smallest eigenvalue at least
$-\lambda $
has a finite forbidden subgraph characterization if and only if
$\lambda < \lambda ^*$
, where

and
$\rho $
is the unique real root of
$x^3 = x + 1$
.
Remark. The constants defined in Theorems 1.2 and 1.3 satisfy
$2 < \lambda ^* = \alpha _2 < \alpha _3 < \dots < \lambda '$
. The constant
$\rho $
defined in Theorem 1.3, known as the plastic number, is the smallest Pisot–Vijayaraghavan number.Footnote
3
As a byproduct, we determine all the limit points of the set of smallest eigenvalues of graphs. Let
$\Lambda _1$
consist of
$\lambda \in \mathbb {R}$
such that
$-\lambda $
is the smallest eigenvalue of some graph. Here we invert the smallest eigenvalues of graphs similarly to how we define
$\mathcal {G}(\lambda )$
. Before our work, Hoffman [Reference Hoffman18] characterized all the limit points of
$\Lambda _1$
in
$(-\infty , 2]$
. His result involves a technical qualification, which was conjectured to be always true, and was later established by Greaves et al. [Reference Greaves, Koolen, Munemasa, Sano and Taniguchi15, Theorem 1]. Doob [Reference Doob14, Theorem 9] observed that every real number in
$\left \{{\alpha _2, \alpha _3, \alpha _4, \dots }\right \} \cup [\lambda ',\infty )$
is a limit point of
$\Lambda _1$
, and conjectured that
$\alpha _2$
(which equals
$\lambda ^*$
),
$\alpha _3, \alpha _4 \dots $
are the only limit points of
$\Lambda _1$
in
$(2, \lambda ')$
. We refute this conjecture by finding all the limit points of
$\Lambda _1$
in
$(2, \lambda ')$
.
Corollary 1.4. For every
$\lambda> 2$
, the negative number
$-\lambda $
is a limit point of the set of smallest eigenvalues of graphs if and only if
$\lambda \ge \lambda ^*$
.
We next turn our attention to signed graphs, which are graphs whose edges are each labeled by
$+$
or
$-$
. Throughout the paper we decorate variables for signed graphs with the
$\pm $
superscript. When we talk about eigenvalues of a signed graph
$G^\pm $
on n vertices, we refer to its signed adjacency matrix – the
$n \times n$
matrix whose
$(i,j)$
-th entry is
$1$
if
$ij$
is a positive edge,
$-1$
if
$ij$
is a negative edge, and
$0$
otherwise.
It still makes sense to speak of forbidden subgraph characterization of a family of signed graphs with eigenvalues bounded from below. Our second main theorem establishes the same threshold phenomenon for the family of signed graphs with smallest eigenvalue at least
$-\lambda $
.
Theorem 1.5. The family
$\mathcal {G}^\pm (\lambda )$
of signed graphs with smallest eigenvalue at least
$-\lambda $
has a finite forbidden subgraph characterization if and only if
$\lambda < \lambda ^*$
.
Notice that if
$\mathcal {F}^\pm $
is a finite forbidden subgraph characterization of
$\mathcal {G}^\pm (\lambda )$
, then the set of all-positive signed graphs in
$\mathcal {F}^\pm $
is a finite forbidden subgraph characterization of
$\mathcal {G}(\lambda )$
. Therefore, for
$\lambda < \lambda ^*$
, Theorem 1.5 implies Theorem 1.3. However, in Section 2, we still provide the complete proof of Theorem 1.3 to illustrate the key ideas.
Finally, we turn our attention to the largest eigenvalue of a signed graph. We denote the eigenvalues of a signed graph
$G^\pm $
by
$\lambda _1(G^\pm ) \le \lambda _2(G^\pm ) \le \dots $
in ascending order, and by
$\lambda ^1(G^\pm ) \ge \lambda ^2(G^\pm ) \ge \dots $
in descending order. Observing that for every signed graph
$G^\pm $
,
$\lambda _1(G^\pm ) \ge -\lambda $
if and only if
$\lambda ^1(-G^\pm ) \le \lambda $
, where
$-G^\pm $
reverses all the edge signs of
$G^\pm $
, we obtain an immediate consequence of Theorem 1.5.
Corollary 1.6. The family
$\mathcal {G}^\mp (\lambda )$
of signed graphs with largest eigenvalue at most
$\lambda $
has a finite forbidden subgraph characterization if and only if
$\lambda < \lambda ^*$
.
Our motivation to understand the forbidden subgraph characterization of
$\mathcal {G}^\mp (\lambda )$
comes from the problem of determining the maximum size of a spherical two-distance set with two fixed angles (one acute and one obtuse) in high dimensions. For fixed
$-1 \le \beta < 0 \le \alpha < 1$
, let
$N_{\alpha ,\beta }(d)$
denote the maximum number of unit vectors in
$\mathbb {R}^d$
where all pairwise inner products lie in
$\left \{{\alpha ,\beta }\right \}$
.
In the special case
$\alpha = -\beta $
, which corresponds to equiangular lines, recent work [Reference Bukh5, Reference Balla, Dräxler, Keevash and Sudakov2, Reference Jiang and Polyanskii20] culminated in a solution [Reference Jiang, Tidor, Yao, Zhang and Zhao21] of Jiang, Tidor, Yao, Zhang and Zhao to the problem of determining
$N_{\alpha , -\alpha }(d)$
for sufficiently large d. We refer the reader to [Reference Jiang, Tidor, Yao, Zhang and Zhao21, Section 1] for earlier developments on equiangular lines with fixed angles in high dimensions.
For the general case
$-1 \le \beta < 0 \le \alpha < 1$
, in their subsequent work [Reference Jiang, Tidor, Yao, Zhang and Zhao22], Jiang et al. proposed a conjecture on the limit of
$N_{\alpha ,\beta }(d)/d$
as
$d\to \infty $
. To state their conjecture, we need the following spectral graph theoretic quantity.
Definition 1.7. Given
$\lambda> 0$
and
$p \in \mathbb {N}$
, define the quantity

where
$ \left \lvert {G^\pm } \right \rvert $
is the number of vertices of
$G^\pm $
,
$\operatorname {mult}(\lambda , G^\pm )$
is the multiplicity of
$\lambda $
as an eigenvalue of
$G^\pm $
, and
$\chi (G^\pm )$
is the chromatic number of the signed graph
$G^\pm $
.
We postpone the definition of
$\chi (G^\pm )$
, which takes values in
$\mathbb {N}^+ \cup \left \{{\infty }\right \}$
, to Section 4 (see Definition 4.1). We now state the conjecture on
$N_{\alpha ,\beta }(d)$
.
Conjecture 1.8 (Conjecture 1.11 of Jiang et al. [Reference Jiang, Tidor, Yao, Zhang and Zhao22]).
Fix
$-1 \le \beta < 0 \le \alpha < 1$
. Set
$\lambda = (1-\alpha )/(\alpha - \beta )$
and
$p = \lfloor -\alpha /\beta \rfloor + 1$
. Then

Conjecture 1.8 was confirmed in [Reference Jiang, Tidor, Yao, Zhang and Zhao22] for
$p \le 2$
and for
$\lambda \in \left \{{1,\sqrt 2, \sqrt 3}\right \}$
separately. Building on the framework developed there, for all
$\lambda < \lambda ^*$
, we establish Conjecture 1.8 as an application of Corollary 1.6, and we reduce the error term
$o(d)$
to a constant depending only on
$\alpha $
and
$\beta $
.
The rest of the paper is organized as follows. We prove Theorem 1.3 and Corollary 1.4 in Section 2, and we prove Theorem 1.5 in Section 3. Part of the proofs are computer-assisted with validated numerics. In Appendix A we explain our computer-aided proof and how anyone can recreate it independently. In Section 4 we give an application of Theorem 1.5 to spherical two-distance sets. In Section 5, we discuss open problems related to the classification of graphs in
$\mathcal {G}(\lambda ^*) \setminus \mathcal {G}(2)$
, and a possible way to establish more instances of Conjecture 1.8.
2 Forbidden subgraphs for
$\mathcal {G}(\lambda )$
We break the proof of Theorem 1.3 into three cases
$\lambda < 2$
,
$\lambda \in [2, \lambda ^*)$
and
$\lambda \ge \lambda ^*$
, where
$\lambda ^* \approx 2.01980$
is defined as in Theorem 1.3. It is worth pointing out the spectral graph theoretic interpretation of the peculiar constant
$\lambda ^*$
.
Proposition 2.1 [Reference Jiang and Polyanskii20, Lemma 5(e)].
For every
$n \in \mathbb {N}^+$
, define the graph
$E_{2,n}$
as in Figure 2. As
$n \to \infty $
,
$\lambda ^1(E_{2,n})$
increases to
$\lambda ^*$
, or equivalently
$\lambda _1(E_{2,n})$
decreases to
$-\lambda ^*$
.Footnote
4

Figure 2
$E_{2,n}$
.
2.1 Proof of Theorem 1.3 for
$\lambda < 2$
Hoffman demonstrated several sequences of graphs
$G_1, G_2, \dots $
such that
$G_m$
is a subgraph of
$G_{m+1}$
for every m and
$\lim _{m \to \infty } \lambda _1(G_m) \le -2$
. To state these results, we introduce the following notions.
Definition 2.2. Given a nonempty vertex subset A of a graph F,
$\ell \in \mathbb {N}$
, and
$m \in \mathbb {N}^+$
,
-
(a) the path extension
$(F, A, \ell )$ is obtained from F by adding a path
$v_0 \dots v_\ell $ of length
$\ell $ , and connecting
$v_0$ to every vertex in A;Footnote 5
-
(b) the path-clique extension
$(F, A, \ell , K_m)$ is further obtained from
$(F, A, \ell )$ by adding a clique of order m, and connecting every vertex in the clique to
$v_\ell $ ;
-
(c) the clique extension
$(F, A, K_m)$ is obtained from F by adding a clique of order m, and connecting every vertex in the clique to every vertex in A.
The following figure consists of schematic drawings of the path extension
$(F, A, \ell )$
, the path-clique extension
$(F, A, \ell , K_m)$
, and the clique extension
$(F, A, K_m)$
.

We compile some of Hoffman’s computation [Reference Hoffman18] and two classical results in the following lemma.
Lemma 2.3. Denote by
$C_n$
the cycle of length n, and
$V_2(C_n)$
a set of two adjacent vertices of
$C_n$
. The path-clique extensions and clique-extensions of
$C_n$
satisfy:
-
(c1)
$\lim _{m\to \infty } \lambda _1(C_n, V_2(C_n), \ell , K_m) \le -2$ for fixed
$n \ge 3$ and
$\ell \in \mathbb {N}$ ;
-
(c2)
$\lim _{m\to \infty } \lambda _1(C_n, V_2(C_n), K_m) = -2$ for fixed
$n \ge 3$ .
Denote by
$K_n$
the complete graphs with n vertices. The path-clique extensions of
$K_n$
satisfy:
-
(k)
$\lim _{m\to \infty } \lambda _1(K_m, V(K_m), \ell , K_m) = -2$ for fixed
$\ell \in \mathbb {N}$ .
Denote by
$\overline {K}_2$
the null graph with
$2$
vertices. The path-clique extensions and the clique extensions of
$\overline {K}_2$
satisfy:
-
(n1)
$\lim _{m\to \infty } \lambda _1(\overline {K}_2, V(\overline {K}_2), \ell , K_m) = -2$ for fixed
$\ell \in \mathbb {N}$ ;
-
(n2)
$\lim _{m\to \infty } \lambda _1(\overline {K}_2, V(\overline {K}_2), K_m) = -2$ .
Denote by
$P_n$
the path of length n (with
$n+1$
vertices), and
$S_n$
the star with n leaves. Their smallest eigenvalues satisfy:
-
(p)
$\lim _{n\to \infty } \lambda _1(P_n) = -2$ ;
-
(s)
$\lambda _1(S_n) = -\sqrt {n}$ .
Remark. In Lemma 2.3, (c1) to (n2) are taken directly from [Reference Hoffman18, Lemmas 2.4 to 2.9], and (p) and (s) follow from the classical results
$\lambda ^1(P_n) = 2\cos (\pi /(n+2))$
and
$\lambda ^1(S_n) = \sqrt {n}$
.
Lemma 2.3 allows us to build a finite set of forbidden subgraphs for
$\mathcal {G}(\lambda )$
. To state the result, we introduce the following definition.
Definition 2.4 (Extension family).
Given a graph F and
$\ell , m \in \mathbb {N}^+$
, the extension family
$\mathcal {X}(F, \ell , m)$
of F consists of the path-extension
$(F, A, \ell )$
, the path-clique extension
$(F, A, \ell _0, K_m)$
, and the clique extension
$(F, A, K_m)$
, where A ranges over the nonempty vertex subsets of F, and
$\ell _0$
ranges over
$\left \{{0, \dots , \ell -1}\right \}$
.
Lemma 2.5. For every
$\lambda < 2$
, there exist
$\ell , m \in \mathbb {N}^+$
such that the extension families
$\mathcal {X}(C, \ell , m)$
and
$\mathcal {X}(D, \ell , m)$
are both disjoint from
$\mathcal {G}(\lambda )$
, where C is the claw graph and D is the diamond graph (see Figure 3).

Figure 3 The claw graph C and the diamond graph D.
Proof. From Lemma 2.3(p), we obtain
$\ell \in \mathbb {N}^+$
such that
$P_\ell \not \in \mathcal {G}(\lambda )$
. Clearly, for every nonempty A, the path extensions
$(C, A, \ell )$
and
$(D, A, \ell )$
, each of which contains
$P_\ell $
as a subgraph, are not in
$\mathcal {G}(\lambda )$
. With hindsight, using Lemma 2.3(c1, c2, n1, n2), we choose
$m \in \mathbb {N}^+$
such that none of the following graphs

where
$\ell _0 \in \left \{{0, \dots , \ell + 1}\right \}$
, is in
$\mathcal {G}(\lambda )$
.
It suffices to prove that every clique extension and every path-clique extension in the extension families
$\mathcal {X}(C, \ell , m)$
and
$\mathcal {X}(D, \ell , m)$
contains one of the graphs in (1) as a subgraph. Label the vertices of C and D as in Figure 3, and pick an arbitrary
$\ell _0 \in \left \{{0, \dots , \ell - 1}\right \}$
. The clique extension
$(C, A, K_m)$
and the path-clique extension
$(C, A, \ell _0, K_m)$
respectively contain

and the clique extension
$(D, A, K_m)$
and the path-clique extension
$(D, A, \ell _0, K_m)$
respectively contain

Therefore the extension families
$\mathcal {X}(C, \ell , m)$
and
$\mathcal {X}(D, \ell , m)$
are both disjoint from
$\mathcal {G}(\lambda )$
.
The next result shows that forbidding a star and an extension family of F effectively forbids F itself in every sufficiently large connected graph.
Lemma 2.6. For every graph F and
$k, \ell , m \in \mathbb {N}^+$
, there exists
$N \in \mathbb {N}$
such that for every connected graph G with more than N vertices, if no member in
$\left \{{S_k}\right \} \cup \mathcal {X}(F, \ell , m)$
is a subgraph of G, then neither is F.
Proof. With hindsight, we choose
$N = vd^{\ell }$
, where d is the Ramsey number
$R(k, 2^{v + 1}m + v)$
, and v is the number of vertices of F. Suppose that G is a connected graph with more than N vertices that contains no member in
$\left \{{S_k}\right \} \cup \mathcal {X}(F, \ell , m)$
as a subgraph. Assume for the sake of contradiction that the subgraph of G induced on a vertex subset, say V, is isomorphic to F.
Since G is a connected graph with more than
$vd^{\ell }$
vertices, we claim that either there exists a vertex u at distance at least
$\ell $
from V, or the maximum degree of G is at least d. Indeed, assume for the sake of contradiction that every vertex of G is within distance
$\ell -1$
from V, and the maximum degree of G is less than d. For
$i \in \left \{{0, 1, \dots , \ell -1}\right \}$
, let
$V_i$
be the set of vertices at distance i from V. Clearly
$V_0 = V$
, and so
$ \left \lvert {V_0} \right \rvert = v$
. Using the assumption on the maximum degree of G, one can inductively show that
$ \left \lvert {V_i} \right \rvert \le vd^i$
. Therefore the number of vertices in G expressed as
$\sum _{i = 0}^{\ell - 1} \left \lvert {V_i} \right \rvert $
is at most
$\sum _{i=0}^{\ell - 1}vd^i = v(d^\ell - 1)/(d-1) \le vd^\ell $
, which yields a contradiction.
We break the rest of the proof into two cases.
Case 1: There exists a vertex u at distance at least
$\ell $
from V. Thus the subgraph
$G[V]$
and a shortest path from V to u contains the path extension
$(F, A, \ell )$
as a subgraph for some nonempty
$A \subseteq V$
, which is a contradiction.
Case 2: The maximum degree of G is at least d. Let u be a vertex of G with the maximum degree, and let
$N(u)$
be the set of neighbors of u in G. Since
$S_k$
is not a subgraph of G, the subgraph
$G[N(u)]$
cannot have an independent set of size k. Because
$d = R(k, 2^{v + 1}m + v)$
, the subgraph
$G[N(u)]$
contains a clique of order
$2^{v + 1}m + v$
, and so there exists a clique
$G[U]$
of size
$2^{v + 1}m$
that is vertex-disjoint from F.
We further partition the vertices in U as follows. For every subset A of V, let
$U(A)$
be the set of vertices in U that are adjacent to all vertices in A and not adjacent to any vertex in
$V \setminus A$
. By the pigeonhole principle, there exists a subset A of V such that
$ \left \lvert {U(A)} \right \rvert \ge 2m$
. If A is nonempty, then
$G[V \cup U(A)]$
contains the clique extension
$(F, A, K_m)$
as a subgraph, which is a contradiction.
Hereafter we may assume that
$ \left \lvert {U(\varnothing )} \right \rvert \ge 2m$
, hence the distance between V and
$U(\varnothing )$
is at least
$2$
. Let
$v_0 \dots v_{\ell _0}$
be a shortest path between V and
$U(\varnothing )$
, where
$v_0$
and
$v_{\ell _0}$
are respectively at distance
$1$
from V and
$U(\varnothing )$
. Let
$A_1$
be the nonempty set of vertices in V that are adjacent to
$v_0$
, and denote by
$v_{\ell _0 + 1}$
an arbitrary vertex in
$U(\varnothing )$
that is adjacent to
$v_{\ell _0}$
. We may assume that
$\ell _0 < \ell - 1$
because otherwise
$G[V \cup \left \{{v_0, \dots , v_{\ell _0}, v_{\ell _0 + 1}}\right \}]$
would contain the path extension
$(F, A_1, \ell )$
as a subgraph, which is a contradiction. Let
$A_2$
be the nonempty set of vertices in
$U(\varnothing )$
that are adjacent to
$v_{\ell _0}$
. If
$ \left \lvert {A_2} \right \rvert \ge m$
, then
$G[V \cup \left \{{v_0, \dots , v_{\ell _0}}\right \} \cup A_2]$
contains the path-clique extension
$(F, A_1, \ell _0, K_m)$
as a subgraph. Otherwise
$ \left \lvert {U(\varnothing ) \setminus A_2} \right \rvert> m$
, and so
$G[V \cup \left \{{v_0, \dots , v_{\ell _0}, v_{\ell _0 + 1}}\right \} \cup (U(\varnothing ) \setminus A_2)]$
contains the path-clique extension
$(F, A_1, \ell _0 + 1, K_m)$
.
Combining Lemmas 2.5 and 2.6, we obtain a finite set of forbidden subgraphs for
$\mathcal {G}(\lambda )$
that forbids the claw graph and the diamond graph in every sufficiently large connected graph. The following result, which is an immediate consequence of [Reference van Rooij and Wilf27, Theorem 4], gives a sufficient condition for line graphs.
Theorem 2.7 (Theorem 4 of van Rooij and Wilf [Reference van Rooij and Wilf27]).
Every graph that contains neither the claw graph nor the diamond graph as a subgraph is a line graph.
Furthermore, we can bootstrap to obtain a finite set of forbidden subgraphs for
$\mathcal {G}(\lambda )$
that forces every sufficiently large connected graph to be the line graph of a tree whose complexity is uniformly bounded.
Lemma 2.8. For every
$\lambda < 2$
, there exist
$N \in \mathbb {N}$
and a finite family
$\mathcal {F}_1$
that is disjoint from
$\mathcal {G}(\lambda )$
such that for every connected graph G with more than N vertices, if G contains no member in
$\mathcal {F}_1$
as a subgraph, then there exists a rooted tree
$H \in \mathcal {T}_N$
such that G is the line graph of H, where
$\mathcal {T}_N$
is the family of rooted trees such that every connected component obtained from removing the root has at most N vertices.
Proof. We obtain
$\ell , m \in \mathbb {N}^+$
from Lemmas 2.3 and 2.5 such that the following family

is disjoint from
$\mathcal {G}(\lambda )$
.
We obtain
$N_0$
from Lemma 2.6 such that for every connected graph G with more than
$N_0$
vertices, if no member in
$\left \{{S_4}\right \} \cup \mathcal {X}(C, \ell , m) \cup \mathcal {X}(D, \ell , m)$
is a subgraph of G, then neither is C nor D. With hindsight, we choose
$N = \max (N_0, (m+1)^{\ell +2})$
. Suppose that G is a connected graph with more than N vertices, and suppose that no member in
$\mathcal {F}_1$
is a subgraph of G. By our choice of
$N_0$
, we know that neither C nor D is a subgraph of G.
Theorem 2.7 implies that G is the line graph of another connected graph, say H. Since
$P_\ell $
is not a subgraph of G, the path
$P_{\ell + 1}$
cannot be a general subgraph of H. In particular, the diameterFootnote
6
of H is at most
$\ell $
. Let d be the maximum degree of H. If
$d < m + 2$
, then H has at most
$(m+1)^{\ell +1}$
vertices, and at most
$(m+1)^{\ell +2}$
edges, and so G has at most
$(m+1)^{\ell +2}$
vertices, which is a contradiction.
We may assume that
$d \ge m + 2$
. Let u be a vertex of H with degree d. We claim that H is a tree. Suppose on the contrary that H contains a cycle as a general subgraph. Since H does not contain
$P_{\ell + 1}$
as a general subgraph, the length of any cycle in H cannot exceed
$\ell + 1$
. Suppose in addition that u is on a cycle in H. Let
$C_n$
be a shortest cycle containing u, where
$n \in \left \{{3, \dots , \ell + 1}\right \}$
. By the choice of
$C_n$
, the vertex u has at least
$d - 2 \ge m$
neighbors outside
$C_n$
. Thus H contains
$C_n$
with
$S_m$
attached at u as a general subgraph, and so G contains
$(C_n, V_2(C_n), K_m)$
as a subgraph, which is a contradiction. Therefore u is not on any cycle in H. Let
$C_n$
be any cycle in H, where
$n \in \left \{{3, \dots , \ell + 1}\right \}$
. Let
$P_{\ell _0}$
be a shortest path between u and
$C_n$
, where
$\ell _0 \in \left \{{1, \dots , \ell }\right \}$
. Since u is not on any cycle, u has at least
$d - 1> m$
neighbors outside
$C_n$
and
$P_{\ell _0}$
. Thus H contains
$C_n \cup P_{\ell _0}$
with
$S_m$
attached at u as a general subgraph, and so G contains
$(C_n, V_2(C_n), \ell _0 - 1, K_m)$
as a subgraph, which is a contradiction.
We now view H as a tree rooted at u. We claim that u is the only vertex with degree larger than m. Suppose on the contrary that there is another vertex
$u'$
with degree at least
$m + 1$
in H. Let
$P_{\ell _0}$
be a shortest path between u and
$u'$
, where
$\ell _0 \in \left \{{1, \dots , \ell }\right \}$
. Thus H contains
$P_{\ell _0}$
with two vertex-disjoint stars
$S_m$
respectively attached to u and
$u'$
as a (general) subgraph, and so G contains
$(K_m, V(K_m), \ell _0 - 1, K_m)$
as a subgraph, which is a contradiction. Therefore all the vertices but u in H has degree at most m. After the root u is removed from H, each connected component has diameter at most
$\ell $
and degree at most m, and so each connected component has at most
$m^{\ell +1} \le N$
vertices.
The last ingredient for the proof of Theorem 1.3 for
$\lambda < 2$
is the following lemma attributed to Dickson, who used it to prove a result about perfect numbers in number theory.
Lemma 2.9 (Lemma A of Dickson [Reference Dickson13]).
For every
$n \in \mathbb {N}^+$
, the partially ordered set
$(\mathbb {N}^n, \le )$
, in which
$(a_1, \dots a_n) \le (b_1, \dots , b_n)$
if and only if
$a_i \le b_i$
for every i, does not contain infinite antichains.
Dickson gave two proofs of Lemma 2.9, one of which uses induction, while the other uses Hilbert’s basis theorem. For the reader’s convenience, we provide a short combinatorial proof based on the infinite Ramsey’s theorem.
Proof. Suppose that
$s_1, s_2, \dots $
is an infinite sequence of distinct tuples in
$\mathbb {N}^n$
. For every
$i < j$
, color the edge
$s_i s_j$
by
$0$
if
$s_i < s_j$
, otherwise by any
$k \in \left \{{1, \dots , n}\right \}$
such that
$s_{i,k}> s_{j,k}$
. The infinite Ramsey’s theorem provides an infinite subset
$I \subseteq \mathbb {N}^+$
such that the edges
$s_i s_j$
, where
$i, j \in I$
and
$i\neq j$
, all receive the same color c. Because
$(\mathbb {N}, \le )$
does not contain infinite descending chains, it must be the case that
$c = 0$
. This implies that
$\left \{{s_i}\colon {i \in I}\right \}$
is an infinite ascending chain, and in particular
$\left \{{s_1, s_2, \dots }\right \}$
is not an antichain.
We are ready to establish the first main theorem for
$\lambda < 2$
.
Proof of Theorem 1.3 for
$\lambda < 2$
.
Let
$N \in \mathbb {N}$
and
$\mathcal {F}_1$
be given by Lemma 2.8, and set

where
$\mathcal {T}_N$
is the family of rooted trees such that every connected component obtained from removing the root has at most N vertices. Setting
$\mathcal {F}_2$
to be the family of graphs that are minimal in
$\widetilde {\mathcal {F}}_2$
under taking subgraphs, one can check that
$\mathcal {F}_0 \cup \mathcal {F}_1 \cup \mathcal {F}_2$
is a forbidden subgraph characterization of
$\mathcal {G}(\lambda )$
.
It suffices to prove that
$\mathcal {F}_2$
is finite. Let
$T_1, \dots , T_n$
be an enumeration of rooted trees on at most N vertices. We encode
$G \in \mathcal {F}_2$
by
$t_G \in \mathbb {N}^n$
as follows. Let H be the rooted tree in
$\mathcal {T}_N$
such that
$G = L(H)$
. After removing the root u from H, we view each connected component as a tree rooted at the vertex that is a child of u in H. Set
$t_G := (t_1, \dots , t_n)$
, where
$t_i$
is the number of occurrences of
$T_i$
as a connected component in the graph obtained by removing u from H. Because no member of
$\mathcal {F}_2$
is a subgraph of any other, one can deduce that
$\left \{{t_G}\colon {G \in \mathcal {F}_2}\right \}$
is an antichain in
$(\mathbb {N}^n, \le )$
, and so
$\mathcal {F}_2$
is finite by Lemma 2.9.
2.2 Proof of Theorem 1.3 for
$\lambda \in [2, \lambda ^*)$
We prove a stronger result from which the first main theorem for
$\lambda \in [2, \lambda ^*)$
and its corollary for
$\lambda \in (2,\lambda ^*)$
follow.
Theorem 2.10. For every
$\lambda \in [2, \lambda ^*)$
, the number of connected graphs in
$\mathcal {G}(\lambda ) \setminus \mathcal {G}(2)$
is finite.
Proof of Theorem 1.3 for
$\lambda \in [2, \lambda ^*)$
assuming Theorem 2.10.
It is known that
$\mathcal {G}(2)$
has a finite forbidden subgraph characterization – in fact, there are
$1,812$
minimal forbidden subgraphs for
$\mathcal {G}(2)$
; see [Reference Bussemaker and Neumaier6]. In view of Theorem 2.10, it suffices to show that if G is a minimal forbidden subgraph for
$\mathcal {G}(\lambda )$
, then G is a minimal forbidden subgraph for
$\mathcal {G}(2)$
, or G, after removing some vertex, is a connected graph in
$\mathcal {G}(\lambda ) \setminus \mathcal {G}(2)$
.
Indeed, let G be a minimal forbidden subgraph for
$\mathcal {G}(\lambda )$
. Because
$\lambda _1(G) < -\lambda \le -2$
, the graph G contains a minimal forbidden subgraph for
$\mathcal {G}(2)$
as a subgraph, say F. Clearly both G and F are connected. We may assume that F is a proper subgraph of G because otherwise we are done already. Consider the graph H obtained after removing v from G, where v is a vertex of G that is furthest from F. Using the connectivity of F, one can then show that H is connected. Finally,
$H \in \mathcal {G}(\lambda )$
because of the minimality of G, and
$H \not \in \mathcal {G}(2)$
because F is a subgraph of H.
Proof of Corollary 1.4 for
$\lambda \in (2,\lambda ^*)$
.
It follows immediately from Theorem 2.10 that
$-\lambda $
is not a limit point of the set of smallest eigenvalues of graphs for
$\lambda \in (2,\lambda ^*)$
The proof of Theorem 2.10 centers around the notion of generalized line graphs. Although we do not need their definition, we state it nevertheless for concreteness.
Definition 2.11 (Cocktail party graphs and generalized line graphs).
The cocktail party graph
$\overline {a K_2}$
is obtained from the complete graph on
$2a$
vertices by deleting a perfect matching. Given a graph G with vertices
$v_1, \dots , v_n$
, and
$a_1, \dots a_n \in \mathbb {N}$
, the generalized line graph
$L(G; a_1, \dots , a_n)$
is obtained from the line graph
$L(G)$
of G by adjoining n vertex-disjoint cocktail party graphs
$\overline {a_1K_2}, \dots , \overline {a_nK_2}$
where every vertex of the ith cocktail party graph
$\overline {a_iK_2}$
is adjacent to every vertex of
$L(G)$
that contains
$v_i$
. See Figure 4 for an example.

Figure 4 A graph G and a schematic drawing of its generalized line graph
$L(G; 2, 1, 0, 3)$
.
Just like line graphs, all the generalized line graphs are in
$\mathcal {G}(2)$
, and they have a finite forbidden subgraph characterization.
Theorem 2.12 (Theorem 2.1 of Hoffman [Reference Hoffman17]).
The smallest eigenvalue of a generalized line graph is at least
$-2$
.
Theorem 2.13 (Cvetković, Doob, and Simić [Reference Cvetković, Doob and Simić10, Reference Cvetković, Doob and Simić11], and Rao, Singhi, and Vijayan [Reference Rao, Singhi and Vijayan26]).
The minimal forbidden subgraphs for the family
$\mathcal {D}_\infty $
of generalized line graphs are listed in Figure 5.
The key observation for the proof of Theorem 2.10 is that for every minimal forbidden subgraph F for
$\mathcal {D}_\infty $
, there exists an extension family of F disjoint from
$\mathcal {G}(\lambda )$
. We first deal with path extensions and clique extensions.
Lemma 2.14. For every minimal forbidden subgraph F in Figure 5 for the family
$\mathcal {D}_\infty $
of generalized line graphs, and every nonempty vertex subset A of F, the path extensions and the clique extensions of F satisfy

Moreover, the equality holds in the first inequality if and only if
$F = G_4$
and
$A \in \left \{{\left \{{3}\right \}, \left \{{4}\right \}}\right \}$
.
We prove Lemma 2.14 under computer assistance in Appendix A. The next result takes care of path-clique extensions.
Lemma 2.15. Suppose that A is a nonempty vertex subset of a graph F and
$\lambda \ge 2$
. If the path extensions of F satisfy

then there exists
$m \in \mathbb {N}^+$
such that the path-clique extensions of F satisfy

Proof. Pick
$\ell _0 \in \mathbb {N}$
such that
$\lambda _1(F, A, \ell ) < -\lambda $
for every
$\ell \ge \ell _0$
. Clearly
$\lambda _1(F, A, \ell , K_m) < -\lambda $
for every
$\ell \ge \ell _0$
and every
$m \in \mathbb {N}^+$
. It suffices to show that for every
$\ell \in \left \{{0, \dots , \ell _0 - 1}\right \}$
there exists
$m \in \mathbb {N}^+$
such that
$\lambda _1(F, A, \ell , K_m) < -\lambda $
.

Figure 5 Minimal forbidden subgraphs for
$\mathcal {D}_\infty $
.
Let
$v_0 \dots v_{\ell _0}$
be the path added to F to obtain
$(F, A, \ell _0)$
, where the vertex
$v_0$
is connected to every vertex in A. Set
$\lambda _0 := -\lambda _1(F, A, \ell _0)$
, and let
$\boldsymbol {x} \colon V(F) \cup \left \{{v_0, \dots , v_{\ell _0}}\right \} \to \mathbb {R}$
be an eigenvector of
$(F, A, \ell _0)$
associated with
$-\lambda _0$
.
Now fix
$\ell \in \mathbb {N}$
with
$\ell < \ell _0$
, and let
$m \in \mathbb {N}^+$
be determined later. Set
$V_\ell := V(F) \cup \left \{{v_0, \dots , v_\ell }\right \}$
, and identify the vertex set of
$(F, A, \ell , K_m)$
with
$V_\ell \cup V(K_m)$
. We abuse notation and write
$x_i$
in place of
$x_{v_i}$
for
$i \in \left \{{\ell , \dots , \ell _0}\right \}$
. Define
$\tilde {\boldsymbol {x}}\colon V_\ell \cup V(K_m) \to \mathbb {R}$
by

We claim that
$\sum _{v \in V_\ell }x_v^2> 0$
. Indeed, assume for the sake of contradiction that
$x_v = 0$
for
$v \in V_\ell $
. Using
$-\lambda _0x_i = \sum _{u\sim v_i}x_u$
for
$i \in \left \{{\ell , \dots , \ell _0-1}\right \}$
, where the sum is taken over all vertices u that are adjacent to the vertex
$v_i$
in
$(F, A, \ell _0)$
, one can prove inductively that
$x_i = 0$
for
$i \in \left \{{\ell + 1, \dots , \ell _0}\right \}$
, which contradicts the assumption that
$\boldsymbol {x}$
is a nonzero vector.
Because
$\sum _{v \in V_\ell } x_v^2> 0$
, clearly
$\tilde {\boldsymbol {x}}$
is a nonzero vector. We compute

Moreover we can compute
$\tilde {\boldsymbol {x}}^\intercal A_{(F, A, \ell , K_m)} \tilde {\boldsymbol {x}}$
as follows

Since
$\boldsymbol {x}$
is an eigenvector of
$(F, A, \ell _0)$
associated with
$-\lambda _0$
, we obtain that

Thus
$\tilde {\boldsymbol {x}}^\intercal A_{(F, A, \ell , K_m)} \tilde {\boldsymbol {x}}$
can be simplified to

The Rayleigh principle says that
$\lambda _1(F, A, \ell , K_m)$
is at most

which, as
$m \to \infty $
, approaches

Here we used the above claim that the denominator in the limit is positive.
Recall that
$\lambda _0 = -\lambda _1(F, A, \ell _0)> \lambda \ge 2$
. It suffices to show that
$(x_\ell + x_{\ell +1})x_{\ell +1} \le 0$
. In fact, we prove inductively that
$(x_i + x_{i+1})x_{i+1} \le 0$
for
$i \in \left \{{\ell , \dots , \ell _0 - 1}\right \}$
. The base case where
$i = \ell _0 - 1$
follows immediately from
$-\lambda _0x_{\ell _0} = x_{\ell _0-1}$
and
$\lambda _0> 2$
. For the inductive step, using
$-\lambda _0x_{i+1} = x_i + x_{i+2}$
and
$\lambda _0> 2$
, we obtain

which is nonpositive by the inductive hypothesis.
We are ready to prove Theorem 2.10.
Proof of Theorem 2.10.
Suppose that
$\lambda \in [2, \lambda ^*)$
. Let
$\mathcal {F}$
denote the set of minimal forbidden subgraphs for the family
$\mathcal {D}_\infty $
. Combining Lemmas 2.14 and 2.15, we choose
$\ell , m \in \mathbb {N}^+$
such that for every
$F \in \mathcal {F}$
, the extension family
$\mathcal {X}(F, \ell , m)$
is disjoint from
$\mathcal {G}(\lambda )$
. From Lemma 2.3(s), we know that
$S_5 \not \in \mathcal {G}(\lambda )$
. In particular, no graph in
$\mathcal {G}(\lambda )$
contains any member in the following family

as a subgraph. Using Lemma 2.6, we obtain
$N \in \mathbb {N}$
such that for every connected graph G with more than N vertices, if no member in (2) is a subgraph of G, then neither is any
$F \in \mathcal {F}$
, and so G is a generalized line graph and is in
$\mathcal {G}(2)$
by Theorems 2.12 and 2.13. This implies that every connected graph in
$\mathcal {G}(\lambda ) \setminus \mathcal {G}(2)$
has at most N vertices.
We end this subsection with a remark on an alternative proof of Theorem 2.10. Instead of working with the family
$\mathcal {D}_\infty $
, one should be able to establish Lemma 2.14 for every minimal forbidden subgraph for
$\mathcal {G}(2)$
, and then prove Theorem 2.10 similarly. This alternative approach is more direct as it does not rely on generalized line graphs. However the drawback is obvious – there are
$1,812$
minimal forbidden subgraphs for
$\mathcal {G}(2)$
. To the best of our knowledge, there is no easily accessible database for these
$1,812$
graphs; for example, a complete list of these graphs was printed to microfiche in [Reference Bussemaker and Neumaier6], and a list of only
$92$
of these graphs up to
$8$
vertices was available in [Reference Cvetković, Rowlinson and Simić12, Fig. 2.4 and Table A1.2]. One certainly can implement the reasonably fast algorithm in [Reference Bussemaker and Neumaier6] to enumerate the minimal forbidden subgraphs for
$\mathcal {G}(2)$
. However as we strive to keep the computer-assisted part of the proof to a minimum, we work with the
$31$
minimal forbidden subgraphs for
$\mathcal {D}_\infty $
.
In fact, the family
$\mathcal {G}(2)$
and its subfamily
$\mathcal {D}_\infty $
are interchangeable when both families are restricted to sufficiently large connected graphs because of a characterization of
$\mathcal {G}(2)$
due to Cameron, Goethals, Seidel, and Shult. To state their characterization, we introduce the following definitions.
Definition 2.16 (Representation of graphs and root systems
$D_n$
and
$E_8$
).
Given a graph G and
$V \subseteq \mathbb {R}^n$
, we say that G is represented by V if the Gram matrix of V is equal to
$A_G + 2I$
, where
$A_G$
is the adjacency matrix of G. The root systems
$D_n$
and
$E_8$
are defined by the standard basis
$e_1, \dots e_n$
of
$\mathbb {R}^n$
as follows

Theorem 2.17 (Theorems 4.2, 4.3 and 4.10 of Cameron et al. [Reference Cameron, Goethals, Seidel and Shult7]).
For every connected graph G, the smallest eigenvalue of G is at least
$-2$
if and only if G is represented by a subset of
$D_n$
or
$E_8$
. Moreover
-
(a) a graph (not necessarily connected) is represented by a subset of
$D_n$ if and only if it is a generalized line graph, and
-
(b) a graph represented by a subset of
$E_8$ has at most
$36$ vertices, and its maximum degree is at most
$28$ .
In particular, Theorem 2.17 implies that every connected graph in
$\mathcal {G}(2)$
with more than
$36$
vertices is a generalized line graph.
2.3 Proof of Theorem 1.3 for
$\lambda \ge \lambda ^*$
Suppose that
$\left \{{F_1, \dots , F_n}\right \}$
is a finite forbidden subgraph characterization of
$\mathcal {G}(\lambda )$
. Because every graph that is not in
$\mathcal {G}(\lambda )$
contains
$F_i$
as a subgraph for some
$i \in \left \{{1, \dots , n}\right \}$
, no graph has its smallest eigenvalue in the open interval
$(\max \left \{{\lambda _1(F_i)}\colon {i \in \left \{{1, \dots , n}\right \}}\right \},-\lambda )$
. Recall that
$\Lambda _1$
consists of
$\lambda \in \mathbb {R}$
such that
$-\lambda $
is the smallest eigenvalue of some graph. The contrapositive of the above observation says the following.
Proposition 2.18. Let
$\lim _+ \Lambda _1 := \left \{{\lambda \in \mathbb {R}}\colon {(\lambda , \lambda + \varepsilon ) \cap \Lambda _1 \neq \varnothing \text { for every }\varepsilon> 0}\right \}$
be the set of right-sided limit points of
$\Lambda _1$
. The family
$\mathcal {G}(\lambda )$
does not have a finite forbidden subgraph characterization for any
$\lambda \in \lim _+ \Lambda _1$
.
In fact, we prove that
$\Lambda _1$
is dense in
$(\lambda ^*, \infty )$
, from which the first main theorem and its corollary for
$\lambda \ge \lambda ^*$
follow.
Theorem 2.19. For every
$\lambda> \lambda ^*$
, there exist graphs
$G_1, G_2, \dots $
such that
$\lim _{n\to \infty } \lambda _1(G_n) = -\lambda $
.
Proof of Theorem 1.3 for
$\lambda \ge \lambda ^*$
.
Theorem 2.19 implies that
$\lim _+\Lambda _1 \supseteq [\lambda ^*, \infty )$
, which implies through Proposition 2.18 that
$\mathcal {G}(\lambda )$
has no finite forbidden subgraph characterization for any
$\lambda \ge \lambda ^*$
.
Proof of Corollary 1.4 for
$\lambda \ge \lambda ^*$
.
It follows immediately from Theorem 2.19 that
$-\lambda $
is a limit point of the set of smallest eigenvalues of graphs for
$\lambda \ge \lambda ^*$
.
A large chunk of Theorem 2.19 is essentially established by Shearer [Reference Shearer28], who proved that the set of spectral radii of all graphs is dense in
$(\lambda ', \infty )$
, where
$\lambda ' = \sqrt {2+\sqrt {5}}$
. As was pointed out in [Reference Doob14], Shearer actually proved that the set of spectral radii of all caterpillar treesFootnote
7
is dense in
$(\lambda ', \infty )$
already. Since a caterpillar tree is bipartite, we rephrase Shearer’s result in terms of smallest eigenvalues.
Theorem 2.20 (Shearer [Reference Shearer28]; cf. Theorem 3 of Doob [Reference Doob14]).
For every
$\lambda \ge \lambda '$
, there exist caterpillar trees
$G_1, G_2, \dots $
such that
$\lim _{n\to \infty } \lambda _1(G_n) = -\lambda $
.
To fill the gap between
$\lambda ^* \approx 2.01980$
and
$\lambda ' \approx 2.05817$
, we use the following graphs.
Definition 2.21 (Rowing graphs).
Given a sequence
$(a_1, \dots , a_n)$
of natural numbers, a rowing graph
$R(a_1, \dots , a_n)$
is obtained from the path
$v_{-2}v_{-1}v_{0}v_{1}\dots v_{n}$
(called the central path) by attaching a vertex (called the coxswain) to
$v_0$
, and attaching a clique of order
$a_i$
to both
$v_{i-1}$
and
$v_i$
for every
$i \in \left \{{1, \dots , n}\right \}$
. See Figure 6 for an example of a rowing graph.

Figure 6 A schematic drawing of the rowing graph
$R(2,0,2,1,0,8,3,1)$
.
We consider rowing graphs for the following two heuristics. On the one hand, a rowing graph is almost a line graph, whose smallest eigenvalue is at least
$-2$
– when the coxwain is removed from a rowing graph, it becomes the line graph of a caterpillar tree. On the other hand, the rowing graph
$R(a_1, \dots , a_n)$
contains
$E_{2,n}$
(see Figure 2), whose smallest eigenvalue decreases to
$-\lambda ^*$
as
$n \to \infty $
.
We adopt the shorthand

Lemma 2.22. Rowing graphs satisfy the following properties.
-
(a) The smallest eigenvalue of a rowing graph is at least
$-1-\sqrt {2}$ .
-
(b) There exists
$a_1 \in \mathbb {N}$ such that
$\lambda _1(R(a_1)) < -\lambda ' \approx -2.05817$ .
-
(c) For every
$\varepsilon> 0$ there exists
$\ell \in \mathbb {N}^+$ such that
$$ \begin{align*} &\lambda_1(R(a_1, \dots, a_{n_1}, 0^{(\ell)}, b_{1}, \dots, b_{n_2}))> \lambda_1(R(a_1, \dots, a_{n_1}, 0^{(\ell)})) - \varepsilon \\ &\qquad\qquad\qquad\qquad\text{ for every }n_1, n_2 \in \mathbb{N}, (a_1, \dots, a_{n_1}) \in \mathbb{N}^{n_1} \text{ and } (b_1, \dots, b_{n_2}) \in \mathbb{N}^{n_2}. \end{align*} $$
-
(d) For every
$n, \ell \in \mathbb {N}^+$ and
$(a_1, \dots , a_n) \in \mathbb {N}^n$ with
$a_n> 0$ , if
$\lambda _1(R(a_1, \dots , a_n, 0^{(\ell )})) \le -2$ , then for every
$\varepsilon> 0$ there exists
$a_{n+1} \in \mathbb {N}^+$ such that
$$\begin{align*}\lambda_1(R(a_1, \dots, a_{n-1}, a_n-1, a_{n+1}, 0^{(\ell - 1)})) < \lambda_1(R(a_1, \dots, a_n, 0^{(\ell)})) + \varepsilon. \end{align*}$$
-
(e) For every
$\varepsilon> 0$ , there exists
$m \in \mathbb {N}^+$ such that for every
$n \ge m$ and every
$(a_1, \dots , a_n) \in \mathbb {N}^n$ there exists
$k \in \left \{{1, \dots , m}\right \}$ such that
$$\begin{align*}\lambda_1(R(a_1, \dots, a_{k-1}, 0, a_k, \dots, a_n)) < \lambda_1(R(a_1, \dots, a_n)) + \varepsilon. \end{align*}$$
In the proof of Lemma 2.22, we work with vectors
$\boldsymbol {x}$
on the vertex set of a rowing graph, whose coxswain and central path are denoted by
$v_c$
and
$v_{-2}v_{-1} \dots $
, and we abuse notation and write
$x_c$
and
$x_i$
in place of
$x_{v_c}$
and
$x_{v_i}$
respectively.
Proof of (a).
Set
$\boldsymbol {a} = (a_1, \dots , a_n)$
and
$\lambda = \lambda _1(R(\boldsymbol {a}))$
. Let
$v_{-2} \dots v_n$
denote the central path of
$R(\boldsymbol {a})$
, and let
$v_c$
denote the coxswain attached to
$v_0$
. Let
$\boldsymbol {x}\colon V(R(\boldsymbol {a})) \to \mathbb {R}$
be a unit eigenvector of
$R(\boldsymbol {a})$
associated with
$\lambda $
. Clearly we have
$\lambda x_c = x_0$
, which implies that

Let L be the rowing graph
$R(\boldsymbol {a})$
with
$v_c$
removed, and let
$\boldsymbol {x}_L$
be
$\boldsymbol {x}$
restricted to
$V(L)$
. Notice that L is a line graph of a caterpillar tree, and so
$\lambda _1(L) \ge -2$
(see Footnote 1 on Page 2). Finally, we bound the smallest eigenvalue of
$R(\boldsymbol {a})$
as follows:

which implies that
$\lambda \ge -1-\sqrt {2}$
. In the above inequality, we assumed that
$\lambda \le -1$
which follows from the fact that
$R(\boldsymbol {a})$
has an edge.
Proof of (b).
Let
$\boldsymbol {x}$
be the vector that assigns
$1$
to
$v_{-2}$
,
$-2$
to
$v_{-1}$
,
$4$
to
$v_0$
,
$0$
to
$v_1$
,
$-2$
to the coxswain, and
$-4/a_1$
to every vertex in the clique of size
$a_1$
. The Rayleigh principle says that
$\lambda _1(R(a_1))$
is at most

which approaches
$-52/25 = -2.08$
as
$a_1 \to \infty $
.
Proof of (c).
Take
$\ell = \lceil 2/\varepsilon \rceil + 5$
. Let
$v_{-2} \dots v_{n_1 + \ell + n_2}$
denote the central path of the rowing graph
$R(\boldsymbol {a}, 0^{(\ell )}, \boldsymbol {b})$
, where
$\boldsymbol {a} = (a_1, \dots , a_{n_1})$
and
$\boldsymbol {b} = (b_1, \dots , b_{n_2})$
, and let
$\boldsymbol {x}\colon V(R(\boldsymbol {a},0^{(\ell )},\boldsymbol {b})) \to \mathbb {R}$
be a unit eigenvector of
$R(\boldsymbol {a},0^{(\ell )},\boldsymbol {b})$
associated with the smallest eigenvalue. Choose
$k \in \left \{{0, \dots , \ell - 1}\right \}$
such that
$x_{n_1+k}x_{n_1+k+1}$
reaches the minimum in absolute value. In particular, using the inequality
$ \left \lvert {x_{n_1+i}x_{n_1+i+1}} \right \rvert \le (x_{n_1+i}^2 + x_{n_1+i+1}^2)/2$
, we obtain

Notice that removing the edge
$v_{n_1+k} v_{n_1+k+1}$
disconnects
$R(\boldsymbol {a},0^{(\ell )},\boldsymbol {b})$
into two subgraphs, one of which is
$R(\boldsymbol {a}, 0^{(k)})$
, while the other is a line graph, denoted L, of a caterpillar tree. Clearly

As
$\ell \ge 6$
, the graph
$\widetilde {E}_8$
in Figure 1 is a proper subgraph of
$R(\boldsymbol {a}, 0^{(\ell )})$
, and so
$\lambda _1(R(\boldsymbol {a}, 0^{(\ell )})) < -2$
. Together with
$\lambda _1(L) \ge -2$
(see Footnote 1 on Page 2), we obtain

Let
$\boldsymbol {x}_R$
and
$\boldsymbol {x}_L$
be the unit eigenvector
$\boldsymbol {x}$
restricted to
$V(R(\boldsymbol {a},0^{(k)}))$
and
$V(L)$
. Finally, we bound the smallest eigenvalue of
$R(\boldsymbol {a},0^{(\ell )},\boldsymbol {b})$
as follows:

Proof of (d).
Set
$\boldsymbol {a} = (a_1, \dots , a_n, 0^{(\ell )})$
and
$\lambda = \lambda _1(R(\boldsymbol {a}))$
. Let
$\boldsymbol {x}\colon V(R(\boldsymbol {a})) \to \mathbb {R}$
be an eigenvector of
$R(\boldsymbol {a})$
associated with
$\lambda $
. Let
$v_{-2}\dots v_{n + \ell }$
denote the central path of the rowing graph
$R(\boldsymbol {a})$
. In
$R(\boldsymbol {a})$
, let K denote the clique of size
$a_n$
attached to both
$v_{n-1}$
and
$v_n$
, and pick an arbitrary vertex u from K. We clearly have

where
$\sigma = \sum _{u' \in V(K) \setminus \left \{{u}\right \}} x_{u'}$
.The above identities imply that
$\lambda (x_u - x_n) = x_n - x_u - x_{n+1}$
, and so

Let
$R(\tilde {\boldsymbol {a}})$
be the rowing graph obtained from
$R(\boldsymbol {a})$
by removing u and attaching a clique, denoted
$\tilde {K}$
, to both
$v_n$
and
$v_{n+1}$
, where
$\tilde {\boldsymbol {a}} = (a_1, \dots , a_n-1,a_{n+1},0^{(\ell -1)})$
, and the order
$a_{n+1}$
of
$\tilde {K}$
is chosen later. In particular,
$V(R(\tilde {\boldsymbol {a}})) = V(R(\boldsymbol {a})) \setminus \left \{{u}\right \} \cup V(\tilde {K})$
.
We define a vector
$\tilde {\boldsymbol {x}}\colon V(R(\tilde {\boldsymbol {a}}))\to \mathbb {R}$
as follows:

Because
$\boldsymbol {x}$
is a nonzero vector, one can check that so is
$\tilde {\boldsymbol {x}}$
. We compute

Moreover we can compare
$\tilde {\boldsymbol {x}}^\intercal A_{R(\tilde {\boldsymbol {a}})} \tilde {\boldsymbol {x}}$
and
$\boldsymbol {x}^\intercal A_{R(\boldsymbol {a})} \boldsymbol {x}$
as follows

which simplifies via (7) to

The Rayleigh principle says that
$\lambda _1(R(\tilde {\boldsymbol {a}}))$
is at most

which, as
$a_{n+1} \to \infty $
, approaches

Here we assumed that the denominator in the limit is nonzero. Indeed, suppose on the contrary that
$\boldsymbol {x}^\intercal \boldsymbol {x} = x_n^2 + x_u^2 - (x_n + x_u)^2 = -2x_nx_u$
. Because
$-2x_nx_u \le x_n^2 + x_u^2 \le x_n^2 + x_u^2 + x_{n+1}^2 \le \boldsymbol {x}^\intercal \boldsymbol {x}$
, it must be the case that
$x_n + x_u = 0$
and
$x_{n+1} = 0$
. In view of (8), we have
$x_n = x_u = 0$
and hence
$\boldsymbol {x} = \boldsymbol {0}$
, which is a contradiction.
It suffices to prove that

which is equivalent to

Using (8), we know that the left hand side of the last inequality is equal to

From Lemma 2.22(a), we know that
$\lambda \ge -1-\sqrt {2}$
in addition to the assumption that
$\lambda \le -2$
. One can check

We are left to prove
$x_n^2 - \lambda x_nx_{n+1} + x_{n+1}^2 \ge 0$
. In fact, we prove inductively that
$x_i^2 - \lambda x_ix_{i+1} + x_{i+1}^2 \ge 0$
for
$i \in \left \{{n, \dots , n + \ell - 1}\right \}$
. The base case where
$i = n + \ell - 1$
follows immediately from
$\lambda x_{n+\ell } = x_{n+\ell -1}$
. For the inductive step, using
$\lambda x_{i+1} = x_i + x_{i+2}$
, we obtain

which is nonnegative by the inductive hypothesis.
Proof of (e).
Take
$m = \lceil 5/\varepsilon \rceil + 1$
. Let
$v_{-2} \dots v_n$
denote the central path of the rowing graph
$R(\boldsymbol {a})$
, where
$\boldsymbol {a} = (a_1, \dots , a_n)$
, and let
$\boldsymbol {x}\colon V(R(\boldsymbol {a})) \to \mathbb {R}$
be a unit eigenvector associated with the smallest eigenvalue of
$R(\boldsymbol {a})$
. Choose
$k \in \left \{{1, \dots , m}\right \}$
such that
$x_{k-1}$
reaches the minimum in absolute value. In particular,

Let
$v_{-2} \dots v_{k-1}v_* v_{k} \dots v_n$
denote the central path of
$R(\tilde {\boldsymbol {a}})$
, where
$\tilde {\boldsymbol {a}} = (a_1, \dots , a_{k-1},0,a_{k}, \dots , a_n)$
. We naturally view the vertex set of
$R(\tilde {\boldsymbol {a}})$
as
$V(R(\boldsymbol {a})) \cup \{ v_*\}$
, and we extend the unit eigenvector
$\boldsymbol {x}\colon V(R(\boldsymbol {a})) \to \mathbb {R}$
to
$\tilde {\boldsymbol {x}}\colon V(R(\tilde {\boldsymbol {a}})) \to \mathbb {R}$
by setting
$\tilde {x}_* = x_{k-1}$
. The Rayleigh principle says that
$\lambda _1(R(\tilde {\boldsymbol {a}}))$
is at most

From Lemma 2.22(a), we obtain

We now have all of the ingredients needed to establish Theorem 2.19.
Proof of Theorem 2.19.
In view of Theorem 2.20, we may assume that
$\lambda \in (\lambda ^*, \lambda ')$
. Fix
$\varepsilon> 0$
. We assume for the sake of contradiction that no rowing graph has its smallest eigenvalue in
$(-\lambda - \varepsilon , -\lambda )$
because otherwise we are done. Define

Claim. For every
$m \in \mathbb {N}^+$
there exists
$(a_1, \dots , a_n) \in A$
with exactly m nonzero entries such that

Proof of Claim.
We order the elements in S as follows:

Alternatively,
$(a_1, \dots , a_{n_1}) \prec (b_1, \dots , b_{n_2})$
if and only if there exists
$\delta> 0$
such that
$\sum _{i=1}^{n_1} a_ix^i < \sum _{i=1}^{n_2} b_ix^i$
for every
$x \in (0, \delta )$
. We shall repeatedly use the fact that for every
$\ell \in \mathbb {N}^+$
the set

does not contain an infinite descending chain, hence
$(S_\ell , \preceq )$
is well-founded, that is, every nonempty subset of
$S_\ell $
has a minimal element.Footnote
8
This fact can be established by a simple induction on
$\ell $
.
In particular, the ordering on S implies that
$(a_1, \dots a_{k-1}, 0, a_k, \dots , a_n) \prec (a_1, \dots , a_n)$
when both sequences are in S and
$k \in \left \{{1, \dots , n}\right \}$
. Thus it suffices to construct, for every
$m \in \mathbb {N}^+$
, a minimal element
$\boldsymbol {a}^{(m)}$
of
$A_m$
, where

Our inductive construction additionally requires that
$\boldsymbol {a}^{(1)}, \boldsymbol {a}^{(2)}, \boldsymbol {a}^{(3)},\dots $
form a descending chain.
Apply Lemma 2.22(c) to obtain
$\ell \in \mathbb {N}^+$
such that

In particular, when
$n_1 = 0$
, we have

For the base case where
$m = 1$
, note that
$R(0^{(\ell )})$
is just
$E_{2,\ell }$
defined in Proposition 2.1, which satisfies that

Since no rowing graph has its smallest eigenvalue in
$(-\lambda - \varepsilon , -\lambda )$
, the above two inequalities imply

which further implies that the length of every sequence in
$A_1$
is at most
$\ell $
, that is,
$A_1 \subseteq S_{\ell }$
. Moreover, Lemma 2.22(b) implies that
$A_1$
is nonempty. Therefore there exists a minimal element of
$A_1$
.
For the inductive step, suppose that
$m \in \mathbb {N}^+$
, and that
$\boldsymbol {a}^{(1)} \succ \dots \succ \boldsymbol {a}^{(m)}$
are minimal elements of
$A_1, \dots , A_m$
respectively. Say
$\boldsymbol {a}^{(m)} = (b_1, \dots , b_n)$
. Lemma 2.22(d) shows that there exists
$b_{n+1} \in \mathbb {N}^+$
such that
$\boldsymbol {b}^{(m+1)} := (b_1, \dots , b_{n-1}, b_n-1, b_{n+1}) \in A$
. Since
$\boldsymbol {b}^{(m+1)} \prec \boldsymbol {a}^{(m)}$
and
$\boldsymbol {a}^{(m)}$
is minimal in
$A_m$
, it must be the case that
$b_n> 1$
or equivalently
$\boldsymbol {b}^{(m+1)} \in A_{m+1}$
. It suffices to show that

Assume for the sake of contradiction that there exists
$\boldsymbol {b} \in A_{m+1}$
such that
$\boldsymbol {b} \preceq \boldsymbol {b}^{(m+1)}$
and the length of
$\boldsymbol {b}$
is more than
$(m+1)\ell $
. By the pigeonhole principle, there exists an initial segment, denoted
$\boldsymbol {b}'$
, of
$\boldsymbol {b}$
such that
$(\boldsymbol {b}', 0^{(\ell )})$
is an initial segment of
$\boldsymbol {b}$
. We may require in addition that either
$\boldsymbol {b}'$
is the empty sequence or the last entry of
$\boldsymbol {b}'$
is positive. If
$\boldsymbol {b}'$
is the empty sequence, then we can argue similarly to the base case that

which contradicts
$\boldsymbol {b} \in A_{m+1}$
.
We may assume that the last entry of
$\boldsymbol {b}'$
is positive. Because
$(\boldsymbol {b}', 0^{(\ell )})$
is an initial segment of
$\boldsymbol {b}$
, and the last entry of
$\boldsymbol {b}$
is positive, the number of nonzero entries in
$\boldsymbol {b}'$
, denoted
$i \in \mathbb {N}^+$
, is at most m. Since
$\boldsymbol {b}' \prec \boldsymbol {b} \preceq \boldsymbol {b}^{(m+1)} \prec \boldsymbol {a}^{(m)} \preceq \boldsymbol {a}^{(i)}$
, and
$\boldsymbol {a}^{(i)}$
is minimal in
$A_i$
, we know that
$\boldsymbol {b}' \not \in A$
, and in particular

From (11), we know that

Since no rowing graph has its smallest eigenvalue in
$(-\lambda -\varepsilon , -\lambda )$
, the above two inequalities imply

which contradicts
$\boldsymbol {b} \in A_{m+1}$
.
Finally, let m be given by Lemma 2.22(e). The claim provides
$(a_1, \dots , a_n) \in A$
with m nonzero entries such that (10) holds. Let
$\ell \in \mathbb {N}$
be such that

Since the length of
$(a_1, \dots , a_n, 0^{(\ell )})$
is at least m, Lemma 2.22(e) says that there exists
$k \in \left \{{1, \dots , m}\right \}$
such that

However (10) asserts that
$(a_1, \dots , a_{k-1}, 0, a_k, \dots , a_n) \not \in A$
, which implies that

Combining the last three inequalities, we obtain that

3 Forbidden subgraphs for
$\mathcal {G}^\pm (\lambda )$
A useful tool in spectral graph theory for signed graphs is switching – two signed graphs are switching equivalent if one graph can be obtained from the other by reversing all the edges in a cut-set. An important feature of switching equivalence is that the switching equivalent signed graphs all have the same spectrum.
Hereinafter we adopt the following convention for an unsigned graph G. With a slight abuse of notation, we denote by G the all-positive signed graph with underlying graph G. We also denote by
$-G$
the all-negative signed graph with the same underlying graph.
To prove the second main theorem, we need to extend the concepts and results from the previous section. We recommend that the reader go through the rest of the section alongside Section 2.
3.1 Proof of Theorem 1.5 for
$\lambda < 2$
We first generalize path extensions, path-clique extensions, clique extensions, and Lemma 2.3(c1, c2).
Definition 3.1. Given a nonempty vertex subset A of a signed graph
$F^\pm $
, a signed vertex subset
$A^\pm $
of
$F^\pm $
is the vertex subset A together with an assignment of signs (positive or negative). Given, in addition,
$\ell \in \mathbb {N}$
and
$m \in \mathbb {N}^+$
,
-
(a) the path extension
$(F^\pm , A^\pm , \ell )$ is obtained from
$F^\pm $ by adding an all-positive path
$v_0 \dots v_\ell $ of length
$\ell $ and connecting
$v_0$ to every vertex v in A by an edge signed according to the sign of v in
$A^\pm $ ;
-
(b) the path-clique extension
$(F^\pm , A^\pm , \ell , K_m)$ is further obtained from
$(F^\pm , A^\pm , \ell )$ by adding an all-positive clique of size m and connecting every vertex in the clique to
$v_\ell $ by a positive edge;
-
(c) the clique extension
$(F^\pm , A^\pm , K_m)$ is obtained from
$F^\pm $ by adding a clique of order m and connecting every vertex in the clique to every vertex v in A by an edge signed according to the sign of v in
$A^\pm $ .
Lemma 3.2. For every signed cycle
$C^\pm _n$
of length n and every signed set
$A^\pm $
of two adjacent vertices of
$C^\pm _n$
, the path-clique extensions and the clique extensions of
$C^\pm _n$
satisfy:
-
(c1)
$\lim _{m\to \infty } \lambda _1(C_n^\pm , A^\pm , \ell , K_m) \le -2$ for fixed
$\ell \in \mathbb {N}$ ;
-
(c2)
$\lim _{m\to \infty } \lambda _1(C_n^\pm , A^\pm , K_m) \le -2$ .
Proof. Let
$v_0,v_1,\dots , v_{n-1}$
be the vertices of
$C_n^\pm $
, and let
$\sigma \colon E(C^\pm _n) \to \left \{{\pm 1}\right \}$
be the signing of
$C^\pm _n$
. Suppose that
$A^\pm $
is a signed set of
$\left \{{v_0,v_{n-1}}\right \}$
. By switching a suitable subset of
$\left \{{v_0, v_{n-1}}\right \}$
, we may assume that
$A^\pm = \{v_0^+, v_{n-1}^+\}$
. If
$C_n^\pm $
has an even number of positive edges, then
$C^\pm _n$
is already switching equivalent to the all-negative cycle of length n, whose smallest eigenvalue is
$-2$
. If the edge
$v_0v_{n-1}$
is negative, then both
$(C_n^\pm , A^\pm , \ell , K_m)$
and
$(C_n^\pm , A^\pm , K_m)$
contain a subgraph that is switching equivalent to the all-negative triangle
$-K_3$
, and so their smallest eigenvalues are at most
$\lambda _1(-K_3) = -2$
.
Hereafter we assume that
$C_n^\pm $
has an odd number of positive edges and the edge
$v_0v_{n-1}$
is positive. Furthermore, by switching a suitable subset of
$\left \{{v_1, \dots , v_{n-2}}\right \}$
, we may assume that
$v_0v_{n-1}$
is the only positive edge of the signed cycle
$C_n^\pm $
.
For (c1), we define a vector
$\boldsymbol {x}\colon V(C_n^\pm , A^\pm , \ell , K_m) \to \mathbb {R}$
as follows. Let
$v_{n}v_{n+1}\dots v_{n+\ell }$
and
$K_m$
be the path and the clique added to
$C_n^\pm $
to obtain the path-clique extension
$(C_n^\pm , A^\pm , \ell , K_m)$
, where
$v_{n}$
is adjacent to
$v_0$
and
$v_{n-1}$
, and
$v_{n+\ell }$
is adjacent to every vertex in
$K_m$
. Assign
$x_{v_i} = -1$
for
$i \in \left \{{0, \dots , n-1}\right \}$
,
$x_{v_{n+i}} = 2(-1)^i$
for
$i \in \left \{{0, \dots , \ell }\right \}$
, and
$x_u = -x_{n+\ell }/m$
for every
$u \in V(K_m)$
. We abuse notation and write
$x_i$
in place of
$x_{v_i}$
for
$i \in \left \{{0, \dots , n+\ell }\right \}$
. The Rayleigh principle says that the smallest eigenvalue of
$(C_n^\pm , A^\pm , \ell , K_m)$
is at most

which is equal to
$(-2n - 8(\ell +1) - 4/m)/(n + 4(\ell +1) + 4/m)$
, and approaches
$-2$
as
$m \to \infty $
.
For (c2), we define a vector
$\boldsymbol {x} \colon V(C_n^\pm , A^\pm , K_m) \to \mathbb {R}$
as follows. Let
$K_m$
be the clique added to
$C_n^\pm $
to obtain the clique extension
$(C_n^\pm , A^\pm , K_m)$
. Assign
$x_{v_i} = -1$
for
$i \in \left \{{0, \dots , n-1}\right \}$
and
$x_u = -(x_0 + x_{n-1})/m$
for every
$u \in V(K_m)$
. We abuse notation and write
$x_i$
in place of
$x_{v_i}$
for
$i \in \left \{{0,\dots , n-1}\right \}$
. The Rayleigh principle says that the smallest eigenvalue of
$(C_n^\pm , A^\pm , K_m)$
is at most

which is equal to
$(-2n - 4/m)/(n + 4/m)$
, and approaches
$-2$
as
$m \to \infty $
.
We naturally generalize extension families as well as Lemma 2.5.
Definition 3.3 (Signed extension family).
Given a signed graph
$F^\pm $
and
$\ell , m \in \mathbb {N}^+$
, the signed extension family
$\mathcal {X}^\pm (F^\pm , \ell , m)$
of
$F^\pm $
consists of the path-extension
$(F^\pm , A^\pm , \ell )$
, the path-clique extensions
$(F^\pm , A^\pm , \ell _0, K_m)$
, and the clique extension
$(F^\pm , A^\pm , K_m)$
, where
$A^\pm $
ranges over the nonempty signed vertex subsets of
$F^\pm $
, and
$\ell _0$
ranges over
$\left \{{0, \dots , \ell -1}\right \}$
.
Note that for an unsigned graph F and
$\ell , m \in \mathbb {N}^+$
, the extension family
$\mathcal {X}(F, \ell , m)$
(defined in Definition 2.4) is a subfamily of the signed extension family
$\mathcal {X}^\pm (F, \ell , m)$
.
Lemma 3.4. For every
$\lambda < 2$
, there exist
$\ell , m \in \mathbb {N}^+$
such that both the signed extension family
$\mathcal {X}^\pm (C, \ell , m)$
of the claw graph C and the signed extension family
$\mathcal {X}^\pm (D, \ell , m)$
of the diamond graph D are disjoint from
$\mathcal {G}^\pm (\lambda )$
.
Proof. From Lemma 2.3(n1, n2, p) and Lemma 2.5, we obtain
$\ell , m \in \mathbb {N}^+$
such that
$P_\ell \not \in \mathcal {G}^\pm (\lambda )$
, none of the following graphs

is in
$\mathcal {G}^\pm (\lambda )$
, and
$\mathcal {X}(C, \ell , m) \cup \mathcal {X}(D, \ell , m)$
is disjoint from
$\mathcal {G}^\pm (\lambda )$
. Since
$P_\ell \not \in \mathcal {G}^\pm (\lambda )$
, no path extension in
$\mathcal {X}^\pm (C, \ell , m) \cup \mathcal {X}^\pm (D, \ell , m)$
is in
$\mathcal {G}^\pm (\lambda )$
.
We are left to deal with the path-clique extensions and the clique extension

where F is C or D, and
$A^\pm $
is the signed set of a nonempty vertex subset A of F. We break the rest of the proof into three cases.
Case 1:
$A^\pm $
is all-positive or all-negative. This case follows since the signed graphs in (13), after switching the cut-set between
$V(F)$
and its complement in case
$A^\pm $
is all-negative, become respectively
$(F, A, 0, K_m), \dots , (F, A, \ell -1, K_m)$
, and
$(F, A, K_m)$
from the extension family
$\mathcal {X}(F, \ell , m)$
, which is disjoint from
$\mathcal {G}^\pm (\lambda )$
.
Case 2: There exists an edge
$uv$
of F such that
$\left \{{u^-, v^+}\right \} \subseteq A^\pm $
. Since the all-negative triangle
$-K_3$
, whose smallest eigenvalue is
$-2$
, is switching equivalent to a subgraph of
$(F, A^\pm , 0)$
, no signed graph in (13) is in
$\mathcal {G}^\pm (\lambda )$
.
Case 3:
$A^\pm $
is neither all-positive nor all-negative, and no edge
$uv$
of F satisfies that
$\left \{{u^-, v^+}\right \} \subseteq A^\pm $
. Label the vertices of C and D as in Figure 3. We may assume without loss of generality that
$\left \{{1^-, 3^+}\right \} \subseteq A^\pm $
. One can check that, for
$F \in \left \{{C, D}\right \}$
, the signed graphs in (12), after switching at the vertex labeled by
$1$
, are respectively subgraphs of those in (13), and hence no signed graph in (13) is in
$\mathcal {G}^\pm (\lambda )$
.
Next, generalizing Lemma 2.6, we show that forbidding a star, an all-negative complete graph, and a signed extension family of
$F^\pm $
effectively forbids
$F^\pm $
itself in every sufficiently large connected signed graph.
Lemma 3.5. For every signed graph
$F^\pm $
and
$k_1, k_2, \ell , m \in \mathbb {N}^+$
, there exists
$N \in \mathbb {N}$
such that for every connected signed graph
$G^\pm $
with more than N vertices, if
$G^\pm $
does not contain any subgraph that is switching equivalent to any member in
$\left \{{S_{k_1}, -K_{k_2}}\right \} \cup \mathcal {X}^\pm (F^\pm , \ell , m)$
, then
$G^\pm $
does not contain any subgraph that is switching equivalent to
$F^\pm $
either.
We leave the proof to the reader as one can make the proof of Lemma 2.6 into that for Lemma 3.5 by mutatis mutandis. To get started, one might want to take
$N = vd^\ell $
, where d is the Ramsey number
$R(k_1, k_2, 3^{v+1} m + v)$
.
The last ingredient is a sufficient condition for signed line graphs, which generalizes Theorem 2.7.
Definition 3.6 (Bidirected graph and signed line graph).
A bidirected graph is a graph in which each vertex-edge incidence has a positive or negative sign. We decorate variables for bidirected graphs with an arrow on the top. A signed graph
$G^\pm $
is the signed line graph of a bidirected graph
$\vec {H}$
if the underlying graph of
$G^\pm $
is the line graph of
$\vec {H}$
, and moreover for every two distinct edges
$e_1$
and
$e_2$
of
$\vec {H}$
having a vertex v in common, the sign of the edge
$e_1 e_2$
in
$G^\pm $
is the product of the signs of the two incidences
$(v, e_1)$
and
$(v, e_2)$
in
$\vec {H}$
.

Figure 7 A bidirected graph and its signed line graph. In a signed graph, the positive edges are represented by solid segments and the negative edges are represented by dashed segments.
Pictorially we place a sign close to each incidence in a bidirected graph. See Figure 7 for an example of a bidirected graph and its signed line graph.
Remark. Our definition of a signed line graph is different from the existing literature – the conventional definition (e.g., Zaslavsky [Reference Zaslavsky34]) reverses all the edge signs of
$G^\pm $
in Definition 3.6. Our choice of the definition is based on aesthetic reasons: the signed line graph of an all-positive bidirected graph is still all-positive, and the smallest eigenvalue of a signed line graph is at least
$-2$
.
Lemma 3.7 (Proposition 2.2 of Vijayakumar [Reference Vijayakumar31]).
For every signed graph
$G^\pm $
, if
$G^\pm $
does not contain any subgraph that is switching equivalent to the claw graph, the diamond graph, or the all-negative triangle, then
$G^\pm $
is a signed line graph.
Although the conclusion in [Reference Vijayakumar31, Proposition 2.2] is weaker than that in Lemma 3.7, Vijayakumar’s proof already gives Lemma 3.7. Below we reproduce his proof in terms of signed line graphs.
Proof. Let G be the underlying graph of
$G^\pm $
. One can check that G contains neither the claw graph nor the diamond graph as a subgraph. Thus Theorem 2.7 implies that G is a line graph of another graph, denoted H, without isolated vertices. We may identify the vertex set of
$G^\pm $
with the edge set of H.
For every
$v \in V(H)$
, let
$E_v(H)$
be the set of edges that are incident to v in H. Define a signing
$\sigma \colon \left \{{(v, e)}\colon {v \in V(H), e \in E_v(H)}\right \} \to \left \{{\pm 1}\right \}$
such that for every
$v \in V(H)$
,

Such a signing
$\sigma $
can be chosen as follows: for every
$v\in V(H)$
, define
$\sigma (v, e_0) = 1$
for some arbitrary
$e_0 \in E_v(H)$
, and define
$\sigma (v, e) = A_{G^\pm }(e_0, e)$
for any other
$e \in E_v(H)$
. Since
$G^\pm $
does not contain any subgraph that is switching equivalent to the all-negative triangle, it can be seen that (14) holds. Assigning
$\sigma (v, e)$
to every incidence
$(v, e)$
of H, we obtain a bidirected graph
$\vec {H}$
whose signed line graph is
$G^\pm $
.
Similar to Lemma 2.8, we obtain a finite set of forbidden subgraphs for
$\mathcal {G}^\pm (\lambda )$
that forces every sufficiently large connected signed graph to be the singed line graph of a bidirected tree whose complexity is uniformly bounded.
Lemma 3.8. For every
$\lambda < 2$
, there exist
$N \in \mathbb {N}$
and a finite family
$\mathcal {F}_1^\pm $
that is disjoint from
$\mathcal {G}^\pm (\lambda )$
such that for every connected signed graph
$G^\pm $
with more than N vertices, if
$G^\pm $
contains no member in
$\mathcal {F}_1^\pm $
as a subgraph, then there exists a bidirected rooted tree
$\vec {H} \in \vec {\mathcal {T}}_N$
such that
$G^\pm $
is the signed line graph of
$\vec {H}$
, where
$\vec {\mathcal {T}}_N$
is the family of bidirected rooted tree such that every connected component obtained from removing the root has at most N vertices.
Proof. Denote
$\mathcal {C}^\pm _n$
the set of signed cycles of length n. We obtain
$\ell , m \in \mathbb {N}^+$
from Lemmas 2.3, 3.2 and 3.4 such that the following family

is disjoint from
$\mathcal {G}^\pm (\lambda )$
. Let
$\mathcal {F}_1^\pm $
be the smallest family, that is closed under switching, containing
$\widetilde {\mathcal {F}}_1^\pm $
. Clearly
$\mathcal {F}_1^\pm $
is also disjoint from
$\mathcal {G}^\pm (\lambda )$
.
We omit the rest of the proof as it, mutatis mutandis, is very much like the one in the proof of Theorem 1.3 for
$\lambda < 2$
on Page 7. We point out that one should make use of Lemmas 3.5 and 3.7 in place of Lemma 2.6 and Theorem 2.7 respectively.
We are ready to present the proof of the second main theorem for
$\lambda < 2$
.
Proof of Theorem 1.5 for
$\lambda < 2$
.
Let
$N \in \mathbb {N}$
and
$\mathcal {F}_1^\pm $
be given by Lemma 3.8, and set

where
$\vec {\mathcal {T}}_N$
is the family of bidirected rooted tree such that every connected component obtained from removing the root has at most N vertices. Setting
$\mathcal {F}_2^\pm $
to be the family of signed graphs that are minimal in
$\widetilde {\mathcal {F}}_2^\pm $
under taking subgraphs, one can check that
$\mathcal {F}_0^\pm \cup \mathcal {F}_1^\pm \cup \mathcal {F}_2^\pm $
is a forbidden subgraph characterization of
$\mathcal {G}^\pm (\lambda )$
.
It suffices to prove that
$\mathcal {F}_2^\pm $
is finite. Let
$\vec {T}_1, \dots , \vec {T}_n$
be an enumeration of bidirected rooted trees
$\vec {T}$
on at most
$N+1$
vertices. We encode
$G^\pm \in \mathcal {F}_2^\pm $
by
$t_{G^\pm } \in \mathbb {N}^n$
as follows. Let
$\vec {H}$
be the bidirected rooted tree in
$\vec {\mathcal {T}}_N$
such that
$G^\pm $
is the signed line graph of
$\vec {H}$
. After removing the root u from
$\vec {H}$
, let
$U_1, U_2, \dots , U_m$
be the vertex sets of the connected components. For
$i \in \left \{{1, \dots , m}\right \}$
, we view the subgraph
$\vec {H}_i := \vec {H}[\left \{{u}\right \} \cup U_i]$
as a bidirected tree rooted at u. Set
$t_{G^\pm } = (t_1, \dots , t_n)$
, where
$t_i$
is the number of occurrences of
$\vec {T}_i$
in
$\vec {H}_1, \dots , \vec {H}_m$
. Because no member of
$\mathcal {F}_2^\pm $
is a subgraph of any other, one can deduce that
$\left \{{t_{G^\pm }}\colon {G^\pm \in \mathcal {F}_2^\pm }\right \}$
is an antichain in
$(\mathbb {N}^n, \le )$
, and so
$\mathcal {F}_2^\pm $
is finite by Lemma 2.9.
3.2 Proof of Theorem 1.5 for
$\lambda \in [2, \lambda ^*)$
We prove a generalization of Theorem 2.10, from which the second main theorem for
$\lambda \in [2, \lambda ^*)$
follows.
Theorem 3.9. For every
$\lambda \in [2,\lambda ^*)$
, the number of connected signed graphs in
$\mathcal {G}^\pm (\lambda ) \setminus \mathcal {G}^\pm (2)$
is finite.
Proof of Theorem 1.5 for
$\lambda \in [2, \lambda ^*)$
.
Using the fact that every minimal forbidden subgraph for
$\mathcal {G}^\pm (2)$
has at most
$10$
vertices (see [Reference Vijayakumar31]), and Theorem 3.9, one can prove this case by following the argument in the proof of Theorem 1.3 for
$\lambda \in [2, \lambda ^*)$
on Page 9.
Recall from Theorem 2.17(a) that a graph is a generalized line graph if and only if it is represented by a subset of the root system
$D_n$
defined in Definition 2.16. For signed graphs, we work directly with those that are represented by a subset of
$D_n$
.
Definition 3.10 (Representation of signed graphs).
Given a signed graph
$G^\pm $
and a subset V of
$\mathbb {R}^n$
, we say that
$G^\pm $
is represented by V if the Gram matrix of V is equal to
$A_{G^\pm } + 2I$
, where
$A_{G^\pm }$
is the signed adjacency matrix of
$G^\pm $
.
These signed graphs that are represented by a subset of
$D_n$
, like generalized line graphs, also have a finite forbidden subgraph characterization.
Theorem 3.11 (Chawathe and Vijayakumar [Reference Chawathe and Vijayakumar8]).
The minimal forbidden subgraphs for the family
$\mathcal {D}_\infty ^\pm $
of signed graphs that are represented by a subset of
$D_n$
are listed in Figures 5 and 8 up to switching equivalence.

Figure 8 Additional minimal forbidden subgraphs for
$\mathcal {D}_\infty ^\pm $
(up to switching equivalence).
Next, generalizing Lemma 2.14, we carry out the following computation.
Lemma 3.12. For every minimal forbidden subgraph
$F^\pm $
for the family
$\mathcal {D}^\pm _\infty $
, and every nonempty signed vertex subset
$A^\pm $
of
$F^\pm $
, the path extensions and the clique extensions of
$F^\pm $
satisfy

Moreover, the equality holds in the first inequality if and only if
$F^\pm = G_4$
and
$A^\pm \in \left \{{\left \{{3^\pm }\right \}, \left \{{4^\pm }\right \}}\right \}$
.
We prove Lemma 3.12 under computer assistance in Appendix A. Lastly, the proof of Lemma 2.15 can be taken almost verbatim to show the following generalization.
Lemma 3.13. Suppose that
$A^\pm $
is a nonempty vertex subset of a signed graph
$F^\pm $
and
$\lambda \ge 2$
. If the path extensions of
$F^\pm $
satisfy

then there exists
$m \in \mathbb {N}^+$
such that the path-clique extensions of
$F^\pm $
satisfy

We are ready to prove Theorem 3.9.
Proof of Theorem 3.9.
Suppose that
$\lambda \in [2, \lambda ^*)$
. Let
$\mathcal {F}^\pm $
denote the set of minimal forbidden subgraphs for the family
$\mathcal {D}^\pm _\infty $
. According to Theorem 3.11, the family
$\mathcal {F}^\pm $
is finite. Combining Lemmas 3.12 and 3.13, we choose
$\ell , m \in \mathbb {N}^+$
such that for every
$F^\pm \in \mathcal {F}^\pm $
, the signed extension family
$\mathcal {X}^\pm (F^\pm , \ell , m)$
is disjoint from
$\mathcal {G}^\pm (\lambda )$
. We know that
$S_5 \not \in \mathcal {G}^\pm (\lambda )$
and
$-K_4 \not \in \mathcal {G}^\pm (\lambda )$
. In particular, no graph in
$\mathcal {G}^\pm (\lambda )$
contains a subgraph that is switching equivalent to any member in the following family

Using Lemma 3.5, we obtain
$N \in \mathbb {N}$
such that for every connected signed graph
$G^\pm $
with more than N vertices, if no member in (15) is switching equivalent to a subgraph of
$G^\pm $
, then neither is any
$F^\pm \in \mathcal {F}^\pm $
, and so
$G^\pm $
is represented by a subset of
$D_n$
and is clearly in
$\mathcal {G}^\pm (2)$
. This implies that every connected graph in
$\mathcal {G}^\pm (\lambda ) \setminus \mathcal {G}^\pm (2)$
has at most N vertices.
3.3 Proof of Theorem 1.5 for
$\lambda \ge \lambda ^*$
Proof of Theorem 1.5 for
$\lambda \ge \lambda ^*$
.
If
$\mathcal {F}^\pm $
is a finite forbidden subgraph characterization of
$\mathcal {G}^\pm (\lambda )$
, then the all-positive signed graphs in
$\mathcal {F}^\pm $
would form a finite forbidden subgraph characterization of
$\mathcal {G}(\lambda )$
, which contradicts Theorem 1.3 for
$\lambda \ge \lambda ^*$
.
4 Spherical two-distance sets
We introduce the definition of chromatic number for signed graphs.
Definition 4.1 (Chromatic number).
A valid p-coloring of a signed graph
$G^\pm $
is a coloring of the vertices using p colors such that the endpoints of every negative edge receive different colors, and the endpoints of every positive edge receive identical colors. (See Figure 9 for an example.) The chromatic number
$\chi (G^\pm )$
of a signed graph
$G^\pm $
is the smallest p for which
$G^\pm $
has a valid p-coloring. If
$G^\pm $
does not have a valid p-coloring for any p, we write
$\chi (G^\pm ) = \infty $
.

Figure 9 A valid
$3$
-coloring of a signed graph.
Recall from Section 1 the following spectral graph theoretic quantity,

We say that
$k_p(\lambda )$
is achievable if it is finite and the infimum can be attained. The spectral graph theoretic quantity
$k_p(\lambda )$
can be seen as a generalization of the spectral radius order
$k(\lambda )$
defined by

Indeed,
$k_1(\lambda ) = k_2(\lambda ) = k(\lambda )$
. However, the behavior of
$k_p(\lambda )$
is far more mysterious when
$p \ge 3$
(see the remark after [Reference Jiang, Tidor, Yao, Zhang and Zhao22, Definition 1.10] for more details). The technique developed in the current paper enables us to certify values of
$k_p(\lambda )$
whenever
$\lambda < \lambda ^*$
.
Theorem 4.2. For every
$p \ge 3$
and
$\lambda < \lambda ^*$
, there exists
$N \in \mathbb {N}$
such that

and in particular,
$k_p(\lambda )$
is achievable whenever it is finite. Moreover, for
$\lambda = 2$
, there exists
$p_0 \ge 3$
such that
$k_p(2) = p^2/(p-1)^2$
for all
$p \ge p_0$
.
We will return to the proof of Theorem 4.2 towards the end of the subsection. The next result constructs large spherical codes with two fixed angles.
Proposition 4.3 (Proposition 2.2 of Jiang et al. [Reference Jiang, Tidor, Yao, Zhang and Zhao22]).
Fix
$-1 \le \beta < 0 \le \alpha < 1$
. Then
$N_{\alpha ,\beta }(d) \ge d$
for every
$d \in \mathbb {N}^+$
. Moreover if
$k_p(\lambda ) < \infty $
, where
$\lambda =(1-\alpha )/(\alpha - \beta )$
and
$p = \lfloor -\alpha /\beta \rfloor + 1$
, then

Conjecture 1.8 asserts that the constructions in Proposition 4.3 are optimal up to an error sublinear in d. A framework to bound
$N_{\alpha ,\beta }(d)$
from above was formulated in [Reference Jiang, Tidor, Yao, Zhang and Zhao22].
Definition 4.4 (Definition 5.2 of Jiang et al. [Reference Jiang, Tidor, Yao, Zhang and Zhao22]).
Given
$p \in \mathbb {N}^+$
and a family
$\mathcal {H}$
of signed graphs, let
$M_{p, \mathcal {H}}(\lambda , N)$
be the maximum possible value of
$\operatorname {mult}(\lambda , G^\pm )$
over all signed graphs
$G^\pm $
on at most N vertices that do not contain any member of
$\mathcal {H}$
as a subgraph and satisfy
$\chi (G^\pm ) \le p$
and
$\lambda ^{p+1}(G^\pm ) \le \lambda $
.
Theorem 4.5 (Theorem 5.3 of Jiang et al. [Reference Jiang, Tidor, Yao, Zhang and Zhao22]).
Fix
$-1 \le \beta < 0 \le \alpha < 1$
. Set
$\lambda = (1-\alpha )/(\alpha - \beta )$
and
$p = \lfloor -\alpha /\beta \rfloor + 1$
. Let
$\mathcal {H}$
be a finite family of signed graphs with
$\lambda ^1(H^\pm )> \lambda $
for each
$H^\pm \in \mathcal {H}$
. Then

A proper choice of the finite family
$\mathcal {H}$
in the last theorem establishes Conjecture 1.8 for
$\lambda < \lambda ^*$
.
Corollary 4.6. Fix
$-1 \le \beta < 0 \le \alpha < 1$
. Set
$\lambda = (1-\alpha )/(\alpha -\beta )$
and
$p = \lfloor -\alpha / \beta \rfloor + 1$
. If
$\lambda < \lambda ^*$
, then

Proof. According to Corollary 1.6, we can take a finite forbidden subgraph characterization, denoted
$\mathcal {H}$
, of the family
$\mathcal {G}^\mp (\lambda )$
of signed graphs with largest eigenvalue at most
$\lambda $
. Thus
$M_{p, \mathcal {H}}(\lambda , N)$
is the maximum possible value of
$\operatorname {mult}(\lambda , G^\pm )$
over all signed graphs
$G^\pm $
on at most N vertices that satisfy
$\chi (G^\pm ) \le p$
and
$\lambda ^1(G^\pm ) \le \lambda $
. Note that
$M_{p, \mathcal {H}}(\lambda , N)$
is a nondecreasing function of N. We break the rest of the proof into two cases.
Case 1:
$M_{p, \mathcal {H}}(\lambda , N) = 0$
for every
$N \in \mathbb {N}^+$
. In other words, there is no signed graph
$G^\pm $
that satisfy
$\chi (G^\pm ) \le p$
and
$\lambda ^1(G^\pm ) = \lambda $
. Thus this case corresponds to the case where
$k_p(\lambda ) = \infty $
. According to Proposition 4.3 and Theorem 4.5, we have
$N_{\alpha ,\beta }(d) = d + O_{\alpha ,\beta }(1)$
.
Case 2:
$M_{p, \mathcal {H}}(\lambda , N)> 0$
for every
$N \ge d_0$
. This case corresponds to the case where
$k_p(\lambda ) < \infty $
. Suppose that
$d \ge d_0$
. Proposition 4.3 says that
$N_{\alpha ,\beta }(d) \ge d$
, and so
$M_{p, \mathcal {H}}(\lambda , N_{\alpha ,\beta }(d))> 0$
. Let
$G^\pm $
be a signed graph with at most
$N_{\alpha ,\beta }(d)$
vertices that satisfy
$\chi (G^\pm ) \le p$
,
$\lambda ^1(G^\pm ) \le \lambda $
and
$\operatorname {mult}(\lambda , G^\pm ) = M_{p, \mathcal {H}}(\lambda , N_{\alpha ,\beta }(d))$
. Thus

Theorem 4.5 then implies that

which, in view of Theorem 4.2 and Proposition 4.3, gives
$N_{\alpha ,\beta }(d) = k_p(\lambda ) d / (k_p(\lambda ) - 1) + O_{\alpha ,\beta }(1)$
.
Lastly, we come back to the achievability of the spectral graph theoretic quantity
$k_p(\lambda )$
. Actually we may assume in the definition (16) of
$k_p(\lambda )$
that
$G^\pm $
is connected by passing to a suitable connected component of
$G^\pm $
in case
$G^\pm $
is not connected. Furthermore, as we shall mainly work with the signed graph that reverses all the edge signs of
$G^\pm $
in (16), we redefine
$k_p(\lambda )$
as follows:

where
$-G^\pm $
reverses all the edge signs of
$G^\pm $
.
We break the proof of Theorem 4.2 into three cases
$\lambda < 2$
,
$\lambda \in (2, \lambda ^*)$
, and
$\lambda = 2$
.
Proof of Theorem 4.2 for
$\lambda < 2$
.
We obtain
$\ell \in \mathbb {N}^+$
from Lemma 2.3(p) such that
$P_{\ell +1} \not \in \mathcal {G}^\pm (\lambda )$
. Applying Lemma 3.5 to
$F^\pm = K_1$
(the
$1$
-vertex graph),
$k_1 = 4$
,
$k_2 = 3$
and
$m = 2p$
, we obtain
$N \in \mathbb {N}$
such that for every connected signed graph
$G^\pm $
, if
$G^\pm $
does not contain any subgraph that is switching equivalent to any member in
$\left \{{S_4, -K_3}\right \} \cup \mathcal {X}^\pm (K_1, \ell , 2p)$
, then
$G^\pm $
contains at most N vertices.
It suffices to show that for every connected signed graph
$G^\pm $
with more than N vertices, either
$\chi (-G^\pm )> p$
or
$G^\pm \not \in \mathcal {G}^\pm (\lambda )$
. By our choice of N, the signed graph
$G^\pm $
contains a subgraph that is switching equivalent to a member in
$\left \{{S_4, -K_3}\right \} \cup \mathcal {X}^\pm (K_1, \ell , 2p)$
. We break the rest of the proof into two cases.
Case 1:
$G^\pm $
contains a subgraph that is switching equivalent to
$S_4$
,
$-K_3$
, or a path extension in
$\mathcal {X}^\pm (K_1, \ell , 2p)$
. Since
$\lambda _1(S_4) = \lambda _1(-K_3) = -2$
and every path extension in
$\mathcal {X}^\pm (K_1, \ell , 2p)$
is switching equivalent to
$P_{\ell +1}$
, the signed graph
$G^\pm $
is not in
$\mathcal {G}^\pm (\lambda )$
.
Case 2:
$G^\pm $
contains a subgraph that is switching equivalent to a path-clique extension or a clique extension in
$\mathcal {X}^\pm (K_1, \ell , 2p)$
. Observe that every path-clique extension and every clique extension in
$\mathcal {X}^\pm (K_1, \ell , 2p)$
contains a subgraph that is switching equivalent to
$K_{2p+1}$
, and moreover every signed graph that is switching equivalent to
$K_{2p+1}$
always contains
$K_{p+1}$
as a subgraph. Thus the signed graph
$G^\pm $
contains
$K_{p+1}$
as a subgraph. Therefore
$-G^\pm $
contains
$-K_{p+1}$
as a subgraph, and so
$\chi (-G^\pm ) \ge p+1$
.
The proof of Theorem 4.2 for
$\lambda \in (2, \lambda ^*)$
follows immediately from Theorem 3.9.
Proof of Theorem 4.2 for
$\lambda \in (2, \lambda ^*)$
.
Theorem 3.9 implies that there exists
$N\in \mathbb {N}$
such that every connected signed graph
$G^\pm $
with
$\lambda _1(G^\pm ) = -\lambda $
has at most N vertices.
The proof of Theorem 4.2 for
$\lambda = 2$
is trickier. We need the following generalization of Theorem 2.17. This generalization was explicitly recognized by Chawathe and Vijayakumar [Reference Chawathe and Vijayakumar8, Theorem 1.1], who attributed the result to Witt [Reference Witt33], and referred the reader to [Reference Cameron, Goethals, Seidel and Shult7] for a combinatorial proof.
Theorem 4.7 (Witt [Reference Witt33] and Cameron et al. [Reference Cameron, Goethals, Seidel and Shult7]).
For every connected signed graph
$G^\pm $
, the smallest eigenvalue of
$G^\pm $
is at least
$-2$
if and only if
$G^\pm $
is represented by a subset of
$D_n$
or
$E_8$
.
Recall that
$\mathcal {D}_\infty ^\pm $
denotes the family of signed graphs that are represented by a subset of
$D_n$
for some
$n \in \mathbb {N}^+$
, where

We shall focus on the signed graphs in
$\mathcal {D}_\infty ^\pm $
. To that end, we generalize bidirected graphs to multigraphs as follows.
Definition 4.8 (Bidirected multigraph).
A bidirected multigraph is a multigraph (allowing parallel edges, but no loops) in which each vertex-edge incidence
$(v, e)$
has a positive or negative sign, denoted
$\sigma (v, e) \in \left \{{\pm 1}\right \}$
, such that every pair of parallel edges
$e_1$
and
$e_2$
satisfies the following condition: for one of the shared endpoints, say
$v_1$
, the incidences
$(v_1, e_1)$
and
$(v_1, e_2)$
have the same sign, while for the other shared endpoint
$v_2$
, the incidences
$(v_2, e_1)$
and
$(v_2, e_2)$
have opposite signs. In particular, there are at most two edges between any two vertices in a bidirected multigraph.
As we transfer relevant properties of a signed graph
$G^\pm \in \mathcal {D}_\infty ^\pm $
to a bidirected multigraph
$\vec {H}$
, a vertex coloring of
$G^\pm $
will correspond to an edge coloring of
$\vec {H}$
.
Definition 4.9 (Intersecting edges and proper edge colorings).
As opposed to parallel edges, we say that two edges of a bidirected multigraph intersect at a vertex v if v is the only vertex they have in common. A proper p-edge-coloring of a bidirected multigraph
$\vec {H}$
is a coloring of the edges using p colors such that for each pair of edges
$e_1, e_2$
that intersect at one vertex, say v, the edges
$e_1$
and
$e_2$
receive different colors when the incidences
$(v, e_1)$
and
$(v, e_2)$
have the same sign, whereas
$e_1$
and
$e_2$
receive identical colors when
$(v, e_1)$
and
$(v, e_2)$
have opposite signs. See Figure 10 for an example.

Figure 10 A proper
$3$
-edge-coloring of a bidirected multigraph.
Definition 4.10 (Associated vectors and uniform vertices).
Suppose that
$v_1, \dots , v_n$
are the vertices of a bidirected multigraph
$\vec {H}$
. For every edge
$e \in E(\vec {H})$
with endpoints
$v_i$
and
$v_j$
, define the associated vector of the edge e by

We say that a vertex v of
$\vec {H}$
is uniform if the sign
$\sigma (v, e)$
is the same for every edge e incident to v.
Proposition 4.11. There is a one-to-one correspondence between signed graphs
$G^\pm $
in
$\mathcal {D}_\infty ^\pm $
and bidirected multigraphs
$\vec {H}$
without isolated vertices such that the following properties hold.
-
(1) The underlying graph of
$G^\pm $ is the line graph of
$\vec {H}$ , in particular, the number of vertices of
$G^\pm $ is equal to the number of edges of
$\vec {H}$ .
-
(2) The signed graph
$G^\pm $ is connected if and only if
$\vec {H}$ is connected except when
$\vec {H}$ consists of a single pair of parallel edges.
-
(3) The signed graph
$-G^\pm $ has a valid p-coloring if and only if
$\vec {H}$ has a proper p-edge-coloring.
-
(4) The multiplicity
$\operatorname {mult}(-2, G^\pm )$ is equal to
$m - \operatorname {\mathrm {rank}} V$ , where m is the number of edges of
$\vec {H}$ and
$V = \{\vec {e} \in D_n \colon e \in E(\vec {H})\}$ .
-
(5) The signed graph
$G^\pm $ is all-positive if and only if every vertex of
$\vec {H}$ is uniform.
Proof. Suppose that
$G^\pm $
is represented by
$V \subseteq D_n$
. Construct a bidirected multigraph
$\vec {H}$
with vertices
$v_1, \dots , v_n$
and incidence signing
$\sigma $
as follows. For each vector in V of the form
$\varepsilon _1 \boldsymbol {e}_i + \varepsilon _2 \boldsymbol {e}_j$
with
$i < j$
, we put an edge e connecting
$v_i$
and
$v_j$
, and we assign
$\sigma (v_i, e) = \varepsilon _1$
and
$\sigma (v_j, e) = \varepsilon _2$
. For every pair of parallel edges
$e_1$
and
$e_2$
, which connect
$v_i$
and
$v_j$
, since
$\sigma (v_i, e_1)\boldsymbol {e}_i + \sigma (v_j, e_1)\boldsymbol {e}_j$
and
$\sigma (v_i, e_2)\boldsymbol {e}_i + \sigma (v_j, e_2)\boldsymbol {e}_j$
are two distinct vectors from V, their inner product satisfies

and so their inner product must be
$0$
. Therefore
$\vec {H}$
is a bona fide bidirected multigraph. We may remove all the isolated vertices (if any) from
$\vec {H}$
.
Conversely one can read off the vectors in V from
$\vec {H}$
– for every edge e, include the associated vector
$\vec {e} \in D_n$
in V. Furthermore one can reconstruct
$G^\pm $
from
$\vec {H}$
as follows. We identify the vertex set of
$G^\pm $
with the edge set of
$\vec {H}$
, and for each pair of distinct vertices
$e_1$
and
$e_2$
of
$G^\pm $
, they are adjacent in
$G^\pm $
if and only if
$e_1$
and
$e_2$
intersect at one vertex, say v, in
$\vec {H}$
, and the sign of the edge
$e_1e_2$
in
$G^\pm $
is
$\sigma (v, e_1)\sigma (v, e_2)$
. One can then check that the reconstructed
$G^\pm $
is indeed represented by V.
Finally, it is routine to transfer properties back and forth between
$G^\pm $
and
$\vec {H}$
.
Now we are ready to strengthen the last case of Theorem 4.2.
Lemma 4.12. For every
$p \ge 3$
, every connected signed graph
$G^\pm $
in
$\mathcal {D}_\infty ^\pm $
with
$\chi (-G^\pm ) \le p$
and
$\lambda _1(G^\pm ) = -2$
satisfies

and equality holds if and only if
$G^\pm $
is the all-positive line graph of the complete bipartite graph
$K_{p,p}$
.
Proof. According to Proposition 4.11, it suffices to show that for every connected bidirected multigraph
$\vec {H}$
with n vertices and m edges, if
$\vec {H}$
has a proper p-edge-coloring, then

and equality holds if and only if
$\vec {H}$
is the complete bipartite graph
$K_{p,p}$
with uniform vertices.
Suppose that
$c\colon E(\vec {H}) \to \left \{{c_1, \dots , c_p}\right \}$
is a proper p-edge-coloring of
$\vec {H}$
. One can deduce from Definition 4.9 the following properties.
-
(a) For two edges
$e_1$ and
$e_2$ that intersect at a vertex v, if
$\sigma (v, e_1) = \sigma (v, e_2)$ , then
$c(e_1) \neq c(e_2)$ .
-
(b) For two edges
$e_1$ and
$e_2$ that intersect at a vertex v, if
$\sigma (v, e_1) \neq \sigma (v, e_2)$ , then
$c(e_1) = c(e_2)$ , and any other edge incident to v has to be parallel to either
$e_1$ or
$e_2$ .
We break the rest of the proof on the upper bound of
$m / \operatorname {\mathrm {rank}} V$
into two cases.
Case 1:
$\vec {H}$
has no parallel edges. We claim that the maximum degree of
$\vec {H}$
is at most p. Suppose on the contrary that k edges
$e_1, \dots , e_k$
pairwise intersect at a vertex v, where
$k> p$
. Because
$k> p \ge 3$
, the contrapositive of (b) implies that
$\sigma (v, e_i)$
is the same for
$i \in \left \{{1, \dots , k}\right \}$
. Moreover (a) says that
$c(e_1), \dots , c(e_k)$
are distinct, which contradicts the assumption that c is a proper p-edge-coloring. In particular, the claim implies that

Since
$\vec {H}$
is connected, one can check that the set of vectors
$\left \{{\vec {e} \in D_n}\colon {e \in E(T)}\right \}$
, where T is a spanning tree of
$\vec {H}$
, is linearly independent. Therefore
$\operatorname {\mathrm {rank}} V \in \left \{{n - 1, n}\right \}$
. We may assume that

Indeed, in the case where
$n \le 3$
, we have

in the case where
$n> 2p$
, we have

and in the case where
$\operatorname {\mathrm {rank}} V = n$
, we also have

Let U be the vertex set of
$\vec {H}$
, and define the vertex subset

We claim that
$\vec {H}[U \setminus U_0]$
is bipartite. To specify its bipartition, let
$\boldsymbol {x}\colon U \to \mathbb {R}$
be a nonzero vector in the orthogonal complement of V. For every edge
$e\in E(\vec {H})$
with endpoints
$u_1$
and
$u_2$
, since
$\boldsymbol {x}$
is orthogonal to the vector
$\vec {e}$
associated to e, we know that

hence
$ \left \lvert {\boldsymbol {x}(u_1)} \right \rvert = \left \lvert {\boldsymbol {x}(u_2)} \right \rvert $
. Because
$\vec {H}$
is connected, we deduce that
$ \left \lvert {\boldsymbol {x}(u)} \right \rvert $
is constant for every
$u \in U$
, and in particular,
$\boldsymbol {x}(u) \neq 0$
for every
$u \in U$
. Since every vertex in
$U \setminus U_0$
is uniform, we partition
$U \setminus U_0$
as follows:

In view of (18), we deduce that
$\vec {H}[U \setminus U_0]$
is a bipartite graph with parts
$U_1$
and
$U_2$
, and so the number of edges in
$\vec {H}[U \setminus U_0]$
is at most
$ \left \lvert {U_1} \right \rvert \left \lvert {U_2} \right \rvert $
. According to (b), every vertex in
$U_0$
has degree at most
$2$
. Thus, we can estimate m using
$n_0 := \left \lvert {U_0} \right \rvert $
as follows:

where the last inequality assumes that
$0 \le n_0 \le n-2$
. In the corner case where
$n_0 \ge n-1$
, we can estimate m by

which implies
$m \le 3(n-1)/2 < 2n-3$
. Therefore, we always have

which, as a function of n, increases on
$(1,\infty )$
. As
$n \le 2p$
and
$p \ge 3$
, the function reaches its maximum
$p^2/(2p-1)$
at
$n = 2p$
, and so
$m/\operatorname {\mathrm {rank}} V \le p^2/(2p-1)$
. It is easy to see that equality holds if and only if
$\vec {H}$
is the complete bipartite graph
$K_{p,p}$
with uniform vertices.
Case 2:
$\vec {H}$
has parallel edges. Let
$e_1$
and
$e_2$
be a pair of parallel edges. Since
$\vec {H}$
is connected, there is a spanning tree T that contains
$e_1$
. One can check that the set of vectors
$\left \{{\vec {e} \in D_n}\colon {e \in E(T) \cup \left \{{e_2}\right \}}\right \}$
is linearly independent. Therefore

We next use the discharging method to bound the average degree
$2m / n$
of
$\vec {H}$
. Each vertex v starts with the charge
$d(v)$
, that is, the degree of v. We claim that for every vertex v with
$d(v)> p$
, exactly one of the following two situations happens.
-
(I) The vertex v is uniform, that is, the sign
$\sigma (v, e)$ is the same for every edge e incident to v.
-
(II) The edges incident to v are precisely two pairs
$(e_1, e_1')$ and
$(e_2, e_2')$ of parallel edges satisfying
$\sigma (v, e_1) \neq \sigma (v, e_2)$ , and in particular
$d(v) = 4$ and
$p = 3$ .
Indeed, suppose that
$d(v)> p$
, and suppose that situation (I) does not happen, that is, there exist two edges
$e_1$
and
$e_2$
incident to v such that
$\sigma (v, e_1) \neq \sigma (v, e_2)$
. Because
$d(v)> p \ge 3$
, we may further assume that
$e_1$
and
$e_2$
intersect at v by replacing
$e_1$
or
$e_2$
with another edge incident to v in case
$e_1$
and
$e_2$
are parallel. Because
$d(v) \ge 4$
, according to (b), situation (II) must happen.
To describe the discharging rules, we say that two edges
$e_1$
and
$e_2$
are twin if they are parallel and
$c(e_1) = c(e_2)$
, and for every vertex v with
$d(v)> p$
, we call v a type I vertex when (I) happens, and a type II vertex when (II) happens.
-
• The first discharging rule sends
$1$ from every type I vertex v to each of its neighbors
$v'$ that are connected to v through twin edges.
-
• The second discharging rule sends
$1/2$ from every type II vertex v to each of its two neighbors.
One can further deduce from Definition 4.9 the following property of twin edges.
-
(c) For twin edges
$e_1$ and
$e_2$ incident to a vertex v, if
$\sigma (v, e_1) \neq \sigma (v, e_2)$ , then
$e_1$ and
$e_2$ are the only two edges incident to v.
We first consider the charge of each vertex after the first discharging rule is applied. Suppose that v is a type I vertex, and
$v'$
is an arbitrary vertex that is connected to v through a pair of parallel edges
$e_1$
and
$e_2$
. Since
$\sigma (v, e_1) = \sigma (v, e_2)$
, it must be the case that
$\sigma (v', e_1) \neq \sigma (v', e_2)$
, and so
$v'$
is not a type I vertex. Thus v does not receive any charge from its neighbors. According to (a), among all the edges incident to v, only the parallel ones can receive identical colors. Thus after v sends out charges to its neighbors, its charge decreases exactly to the number of distinct colors assigned to the edges incident to v, and so the charge of v is at most p now. Suppose in addition that
$e_1$
and
$e_2$
are twin edges, or in other words,
$v'$
receives
$1$
from v. According to (c), the only edges incident to
$v'$
are
$e_1$
and
$e_2$
, and so the charge of
$v'$
is at most
$3$
, which is at most p, now.
We then consider the charge of each vertex after the second discharging rule is applied. Notice that the second discharging rule only kicks in when
$p = 3$
. Suppose that v is a type II vertex, and two pairs
$(e_1, e_1')$
and
$(e_2, e_2')$
of parallel edges connect v respectively to
$v_1$
and
$v_2$
such that
$\sigma (v, e_1) \neq \sigma (v, e_2)$
. Without loss of generality, we may assume that
$\sigma (v, e_1) = +1$
and
$\sigma (v, e_2) = -1$
. We have the four possibilities for
$\sigma (v, e_1')$
and
$\sigma (v, e_2')$
. Once
$\sigma (v, e_1')$
and
$\sigma (v, e_2')$
are chosen, we can determine, according to (a) and (b), which edges incident to v receive identical or distinct colors. We summarize the four possibilities in the following figure. For each possibility, it is easy to check that v did not receive any charge from
$v_1$
or
$v_2$
when the first discharging rule was applied. Therefore the final charge of v is
$3$
, which is equal to p.

We are left to decide the final charge of
$v_1$
and
$v_2$
. Observe that for
$i \in \left \{{1,2}\right \}$
there are, au fond, only three possible ways, as shown below, to color
$e_i$
and
$e_i'$
and to assign
$\sigma (v_i, e_i)$
and
$\sigma (v_i, e_i')$
. For the first possibility, according to (c), the only edges incident to
$v_i$
are
$e_i$
and
$e_i'$
, and so the final charge of
$v_i$
is
$5/2$
, which is less than p. For the other two possibilities, the contrapositive of (b) implies that
$v_i$
must be uniform, and so its charge was at most p after the first discharging rule was applied. Assume for the sake of contradiction that
$v_i$
is connected to another type II vertex w through edges
$f_i$
and
$f_i'$
. Since
$v_i$
is uniform, in view of the three possibilities above, we know that
$c(e_i) \neq c(e_i')$
and
$c(f_i) \neq c(f_i')$
. In view of (a), we know that
$\left \{{c(e_i), c(e_i')}\right \} \cap \left \{{c(f_i), c(f_i')}\right \} = \varnothing $
. Thus the four edges
$e_i$
,
$e_i'$
,
$f_i$
and
$f_i'$
receive distinct colors under c, which contradicts the assumption that
$p = 3$
. Therefore v is the only type II neighbor of
$v_i$
, and the final charge of
$v_i$
is at most
$p + 1/2$
.

Since the discharging rules preserve the total charge, the average degree
$2m/n$
is at most the maximum final charge, which is at most
$p + 1/2$
. Finally, we obtain that

Proof of Theorem 4.2 for
$\lambda = 2$
.
From Theorem 4.7 and Lemma 4.12, we know that
$k_p(2)$
is the smaller of the two quantities
$p^2/(p-1)^2$
and

Because there are only finitely many signed graphs that are represented by a subset of
$E_8$
, the infimum in the definition of
$k_p^*$
is in fact a minimum, hence
$k_p(2)$
is achievable. Moreover
$k_p^* \ge k^*$
, where

Since
$k^*> 1$
, there exists
$p_0 \ge 3$
such that
$p^2 / (p-1)^2 \le k^* \le k_p^*$
for every
$p \ge p_0$
.
5 Concluding remarks
Extending Smith’s classification [Reference Smith29] of connected graphs with spectral radius at most
$2$
, Cvetković, Doob and Gutman [Reference Cvetković, Doob and Gutman9, Theorem 3.8] characterized the connected graphs with spectral radius in
$(2, \lambda ')$
, which were later explicitly classified by Brouwer and Neumaier [Reference Brouwer and Neumaier4]. Recall that
$\lambda ' = \sqrt {2 + \sqrt 5} \approx 2.05817$
. Retrospectively, perhaps both Theorem 1.2 and the fact that
$\lambda '$
is the smallest
$\lambda \in \mathbb {R}$
such that the set of graph spectral radii is dense in
$(\lambda , \infty )$
(cf. [Reference Hoffman16, Reference Shearer28]) indicated that it is possible to classify connected graphs with spectral radius at most
$\lambda '$
.
A similar line of research was carried out for signed graphs. McKee and Smyth [Reference McKee and Smyth24] determined all the connected signed graphs with spectral radius at most
$2$
– they are switching equivalent to subgraphs of
$S_{14}, S_{16}$
, and
$T_{2n}$
for
$n \ge 3$
in Figure 11. When investigating Lehmer’s Mahler measure problem, McKee and Smyth [Reference McKee and Smyth25, Theorem 3] further determined the
$17$
connected signed graphs with spectral radius in
$(2, 2.019)$
. Belardo, Cioabă, Koolen and Wang [Reference Belardo, Cioabă, Koolen and Wang3, Problem 3.11] raised the question on the classification of connected signed graphs with spectral radius at most
$\lambda '$
. This question was very recently resolved by Wang, Dong, Hou and Li [Reference Wang, Dong, Hou and Li32].

Figure 11 Maximal connected signed graphs with spectral radius at most
$2$
up to switching equivalence. The number of vertices in
$T_{2n}$
is
$2n$
.
With regard to smallest eigenvalues, we would like to extend Theorem 2.17 of Cameron et al. [Reference Cameron, Goethals, Seidel and Shult7] beyond
$\mathcal {G}(2)$
.
Problem 5.1. Classify all the connected graphs with smallest eigenvalue in
$(-\lambda ^*, -2)$
. In particular, classify such graphs that have sufficiently many vertices.
It is worth mentioning that Bussemaker and Neumaier [Reference Bussemaker and Neumaier6, Theorem 2.5] showed that
$E_{2,6}$
defined in Proposition 2.1 is the only connected graph with smallest eigenvalue in
$[-\lambda ^1(E_{2,6}), -2)$
, where
$\lambda ^1(E_{2,6}) \approx 2.00659$
. Very recently, Acharya and Jiang [Reference Acharya and Jiang1] provided a complete solution to Problem 5.1 – there are 794 infinite families of graphs and 4,752 exceptional graphs.
We can ask the same question for signed graphs, extending Theorem 4.7 beyond
$\mathcal {G}^\pm (2)$
.
Problem 5.2. Classify all the connected signed graphs with smallest eigenvalue in
$(-\lambda ^*, -2)$
. In particular, classify such signed graphs that have sufficiently many vertices.
Turning to spherical two-distance sets with two fixed angles, it is plausible to establish more instances of Conjecture 1.8 by taking advantage of
$\chi (G^\pm )$
in Definition 4.4. Denote by
$\mathcal {G}^\mp _p(\lambda )$
the family of signed graphs
$G^\pm $
with
$\chi (G^\pm ) \le p$
and
$\lambda ^1(G^\pm ) \le \lambda $
. Observe that
$\mathcal {G}^\mp _p(\lambda )$
, just like
$\mathcal {G}^\mp (\lambda )$
, is still closed under taking subgraphs. We raise the following question, whose solution could possibly establish more instances of Conjecture 1.8.
Problem 5.3. For every
$p \in \mathbb {N}^+$
, determine the set of
$\lambda \in \mathbb {R}$
for which
$\mathcal {G}^\mp _p(\lambda )$
has a finite forbidden subgraph characterization.
Notice that
$\mathcal {G}^\mp _1(\lambda )$
consists of unsigned graphs only, and
$\mathcal {G}^\mp _2(\lambda )$
consists of signed graphs that are switching equivalent to unsigned graphs. Thus Theorem 1.2 essentially answers Problem 5.3 when
$p \in \left \{{1,2}\right \}$
.
Finally, we point out that the proof of Theorem 4.2 for
$\lambda = 2$
actually shows that
$k_p(2) = p^2/(p-1)^2$
for every
$p \ge 1/(1-1/\sqrt {k^*})$
, where
$k^*$
is defined as in (19). We compute
$k^* = 15/14$
hence
$1/(1-1/\sqrt {k^*}) \approx 29.49$
. Indeed, Stanić noted in [Reference Stanić30] that, up to switching equivalence, there is a unique maximal signed graph, denoted
$M_8^\pm $
, that is represented by a subset of
$E_8$
, and moreover, the spectrum of
$M_8^\pm $
is
$\left \{{28^8, (-2)^{112}}\right \}$
. By the Cauchy interlacing theorem, every signed graph
$G^\pm $
that is represented by a subset of
$E_8$
satisfies
$\lambda ^1(G^\pm ) \le 28$
, and thus

which implies that
$ \left \lvert {G^\pm } \right \rvert /\operatorname {mult}(-2,G^\pm ) \ge 15/14$
. Since equality holds when
$G^\pm = M_8^\pm $
, we obtain
$k^* = 15/14$
. We leave the determination of
$k_p(2)$
for small p as an open problem.
Problem 5.4. For every
$p \in \left \{{3,4,\dots ,29}\right \}$
, determine the value of
$k_p(2)$
.
A Computer-assisted proofs
Notice that the path extensions
$(G_4, \left \{{3^\pm }\right \}, \ell )$
and
$(G_4, \left \{{4^\pm }\right \}, \ell )$
are switching equivalent to
$E_{2, \ell +3}$
(see Figure 2), and so Proposition 2.1 implies that equality holds for
$F = G_4$
and
$A \in \left \{{\left \{{3}\right \},\left \{{4}\right \}}\right \}$
in Lemma 2.14, and for
$F^\pm = G_4$
and
$A \in \left \{{\left \{{3^\pm }\right \},\left \{{4^\pm }\right \}}\right \}$
in Lemma 3.12.
The termination of a program that solves the following computational problem is a proof of the strict inequalities in Lemmas 2.14 and 3.12. In fact, we strengthen these strict inequalities by replacing
$\lambda ^* \approx 2.0198$
with
$101/50 = 2.02$
. Note that
$-101/50$
is not an algebraic integer, and hence it cannot be an eigenvalue of any signed graph.
Input. The first line of the input gives the number N of minimal forbidden subgraphs for
$\mathcal {D}_\infty ^\pm $
(up to switching equivalence). Each of the N lines that follow represents a signed graph in Figures 5 and 8 by two strings. The first string is the label of the signed graph. The second string is of the form u[1]u[2]…u[2e-1]u[2e] possibly followed by -v[1]v[2]…v[2f-1]v[2f], which lists the positive edges u[1]u[2],…,u[2e-1]u[2e] and the negative edges v[1]v[2],…,v[2f-1]v[2f].
Output. For each of the N signed graphs
$F^\pm $
, output one line containing x y z, where x is the label of
$F^\pm $
, y and z are both -1 if
$\lambda _1(F^\pm ) < -101/50$
already, otherwise y is, in all but one case, the minimum
$\ell $
such that
$\lambda _1(F^\pm , A^\pm , \ell ) < -101/50$
for every nonempty signed vertex subset
$A^\pm $
of
$F^\pm $
, and z is the minimum m such that
$\lambda _1(F^\pm , A^\pm , K_m) < -101/50$
for every nonempty signed vertex subset
$A^\pm $
of
$F^\pm $
. In the exceptional case where the label of
$F^\pm $
is G4, y is defined similarly but
$A^\pm $
cannot be
$\left \{{3^\pm }\right \}$
or
$\left \{{4^\pm }\right \}$
, and an extra
$\texttt {*}$
is appended to y.
Our implementation is straightforward. We iterate through the minimal forbidden subgraphs
$F^\pm $
for
$\mathcal {D}_\infty ^\pm $
. In each iteration, we compute y and z as follows. We first check whether
$A_{F^\pm } + (101/50)I$
is positive definite – if not we output -1 -1 for y z and continue to the next iteration. We test whether a matrix is positive definite by checking whether its leading principal minors are all positive. If the smallest eigenvalue of
$F^\pm $
turns out to be more than
$-101/50$
, we then go through all possible nonempty signed vertex subsets
$A^\pm $
of
$F^\pm $
. For each
$A^\pm $
, we increase
$\ell $
from
$0$
until the determinant of
$A_{(F^\pm , A^\pm , \ell )} + (101/50)I$
is negative, and we increase m from
$1$
until the determinant of
$A_{(F^\pm , A^\pm , K_m)} + (101/50)I$
is negative. We record the largest
$\ell $
and m when we go through
$A^\pm $
, and output them as y and z. In the exceptional case where the label of
$F^\pm $
is G4, we skip the computation of
$\ell $
when
$A^\pm $
is
$\left \{{3^\pm }\right \}$
or
$\left \{{4^\pm }\right \}$
. We avoid floating-point errors by representing every number as a rational number.
The actual code, written in Ruby, is available as the ancillary file maximum_extensions.rb in the arXiv version of this paper. As our code halts, we obtain a proof of Lemmas 2.14 and 3.12. We provide the input in Table 1 for the convenience of anyone who wants to program independently, and we provide in addition the output in Table 2 for cross-check.
Table 1 Input

Table 2 Output

Acknowledgments
We thank Sebastian Cioabă for sending us a copy of [Reference Hoffman18], and Jing Zhang for pointing us to [Reference Dickson13]. We are very grateful to Zhuang Xiong and two anonymous referees for their valuable feedback on the earlier versions of the paper. One referee noted a gap in the proof of a byproduct result of Theorem 1.5 on symmetric hollow integer matrices in an earlier version; that result was removed from the present manuscript and later proved in [Reference Jiang19].
Competing interest
The authors have no competing interests to declare.
Financial support
ZJ was supported in part by an AMS Simons Travel Grant, and by U.S. taxpayers through NSF grant DMS-2127650. AP was supported in part by U.S. taxpayers through NSF grant DMS-2349045.