A composition method for neat formulas of chromatic symmetric functions
Abstract.
We develop a composition method to unearth positive -expansions of chromatic symmetric functions , where the subscript stands for compositions rather than integer partitions. Using this method, we derive positive and neat -expansions for the chromatic symmetric functions of tadpoles, barbells and generalized bulls, and establish the -positivity of hats. We also obtain a compact ribbon Schur analog for the chromatic symmetric function of cycles.
Key words and phrases:
chromatic symmetric functions, -positivity, noncommutative symmetric functions, ribbon Schur functions, Schur positivity2020 Mathematics Subject Classification:
05E05, 05A151. Introduction
In his seminal paper [23], Stanley introduced the concept of the chromatic symmetric function for any graph , which tracks proper colorings of . It is a generalization of Birkhoff’s chromatic symmetric polynomial in the study of the -color problem. Chromatic symmetric functions encode many graph parameters and combinatorial structures, such like the number of vertices, edges and triangles, the girth, and the lattice of contractions, see Martin et al. [17] and [23, Page 167]. For any basis of the algebra of symmetric functions, a graph is said to be -positive if every -coefficient of is nonnegative. Stanley [23, Section 5] brought forward the question that which graphs are -positive, and asserted that a complete characterization of -positive graphs “appears hopeless.” He restated Stanley and Stembridge’s -free conjecture [27], which became a leading conjecture in the study of chromatic symmetric functions henceforth.
Conjecture 1.1 (Stanley and Stembridge).
The chromatic symmetric function of the incomparability graph of every -free poset is -positive.
Gasharov [8] confirmed the Schur positivity of the graphs in Conjecture 1.1, which are all claw-free. Stanley [24] then proposed the following Schur positivity conjecture and attributed it to Gasharov, see also Gasharov [9].
Shareshian and Wachs [21] introduced the notion of chromatic quasisymmetric functions, refined Gasharov’s Schur positivity result, and unveiled connections between Conjecture 1.1 and representation theory. By Guay-Paquet’s reduction [12], Conjecture 1.1 can be restated as that every unit interval graph, or equivalently, every claw-free interval graph, is -positive. These conjectures thereby charm graph theorists that are fascinated by claw-free graphs and interval graphs, see Faudree et al. [7] for an early survey on claw-free graphs, and Corneil et al. [2] for wide applications of interval graphs. The Schur positivity of interval graphs can be shown by using a result of Haiman [13]. Haiman’s proof used Kazhdan and Lusztig’s conjectures that were confirmed later, see [23, Page 187].
Technically speaking, to show that a graph is not -positive or not Schur positive is comparably undemanding, in the sense that the demonstration of a negative - or -coefficient for a particular partition is sufficient, which may call for a scrupulous selection of though. For instance, Wang and Wang [30] proved the non--positivity and non-Schur positivity of some spiders and brooms. Two common criteria for the non-positivity are Wolfgang III’s connected partition criterion and Stanley’s stable partition criterion, see [33] and [24] respectively.
In contrast, to confirm that a graph is -positive is seldom easy. Stanley [23] studied paths and cycles by displaying the generating functions of their chromatic symmetric functions, whose Taylor expansions indicate the -positivity as plain sailing. Gebhard and Sagan [10] lifted up to certain in the algebra of symmetric functions in noncommutative variables, so that equals the commutative image of . They developed a theory for certain -positivity of , which leads to the -positivity of . In particular, -chains are -positive. Tom [29] obtained an -expansion of the chromatic symmetric function of a general unit interval graph in terms of “forest triples,” and used it to reconfirm the -positivity of -chains. Dahlberg and van Willigenburg [4] classified when is a positive linear combination of the elementary symmetric functions in noncommuting variables. Via this -approach, Wang and Wang [31] uncovered the -positivity of two classes of cycle-chords. Aliniaeifard et al. [1] reinterpreted the equivalence idea for the -positivity in terms of the quotient algebra of and obtained the -positivity of kayak paddle graphs. An example of using chromatic quasisymmetric functions to show the -positivity can be found from Huh et al. [14] for melting lollipops.
We think the plainest way of confirming the -positivity of a graph is to compute out and make certain that the -coefficient for each partition is nonnegative. A variant idea is to recast as a linear combination of -positive chromatic symmetric functions with positive coefficients, see Dahlberg and van Willigenburg [3] for a treatment of lollipops for example. Up to late 2023, to the best of our knowledge, only complete graphs, paths, cycles, melting lollipops, -chains, and slightly melting -chains own explicit formulas of chromatic symmetric functions, see Section 2.3 and Tom [29]. In this paper, we conceive a new approach along this way, called the composition method.
We were inspired from Shareshian and Wachs’s discovery
(1.1) |
for paths , where the sum runs over compositions of , and
(1.2) |
They [22, Table 1] obtained Eq. 1.1 by using Stanley’s generating function for Smirnov words, see also Shareshian and Wachs [20, Theorem 7.2]. An equally engaging formula for cycles was brought to light by Ellzey [6], see Proposition 2.4.
The composition method is to expand a chromatic symmetric function in the elementary symmetric functions which are indexed by compositions . This idea can be best understood through Eq. 1.1. The -coefficients, taking Eq. 1.2 for example, are functions defined for compositions. See Section 2.5 for more examples. An ordinary -coefficient for any partition is the sum of “the -coefficients” over all compositions that can be rearranged as ; we write this property of as
(1.3) |
Here arises a potential ambiguity about the wording “the -coefficient”. Namely, when the parts of decrease weakly and so coincides with , it may be understood as either the coefficient of in some -expansion or the coefficient of in the unique -expansion of . This ambiguity comes from the unspecification of the background algebra, which leads us to the algebra of noncommutative symmetric functions, see Sections 2.2 and 2.4 for details.
In order to give a step by step instruction for applying the composition method, we need some basic knowledge of the algebra . First, the commutative images of the basis elements and of are the elementary and power sum symmetric functions and , respectively. Second, every symmetric function has an infinite number of noncommutative analogs in , in which only a finite number are -positive with integer coefficients. Third, a symmetric function is -positive if and only if it has a -positive noncommutative analog. For the purpose of showing the -positivity of a chromatic symmetric function , one may follow the steps below.
- Step 1:
-
Initiate the argument by deriving a noncommutative analog in its -expansion. We know two ways to achieve this. One is to start from the -expansion of by definition, which implies the -expansion of a noncommutative analog directly. Then we transform the analog to its -expansion by change-of-basis, see Appendix A for this approach working for cycles. The other way is to compute by applying Orellana and Scott’s triple-deletion property [19], and by using graphs with known -expansions, see Theorem 3.3 for this way working for tadpoles.
- Step 2:
-
Find a positive -expansion. Decompose the set of all compositions of as , such that
-
(1):
,
-
(2):
the compositions in each have the same underlying partition, say, , and
-
(3):
the inner sum for each has an -positive commutative image, i.e., .
It follows that
(1.4) is a positive -expansion.
-
(1):
- Step 3:
-
Produce a neat -expansion by shaping Eq. 1.4. One thing we can do is to simplify each of the coefficients for given composition functions . Another thing is to further merge the terms for distinct indices, say and , with the same underlying partition . Sign-reversing involutions, injections and bijections may help embellish expressions to make them compact and elegant.
One may catch a whiff of the combinatorial essence of the composition method from each of the steps. Besides suitably selecting a vertex triple to apply the triple-deletion property, a vast flexibility lies in both the process of decomposing and coefficient shaping. We wish that the -positivity of Eq. 1.4 is as transparent as the -positivity in Eq. 1.1. Step is not necessary for the sole purpose of positivity establishment, however, it would be computationally convenient if we make use of a neat -expansion in proving the -positivity of graphs that are of more complex.
In this paper, we start the journey of understanding the computing power of the composition method in proving the -positivity of graphs.
After making necessary preparations in Section 2, we apply the composition method for special families of graphs in Section 3. We work out neat formulas for tadpoles and barbells. The former are particular squids that were investigated by Martin et al. [17], see also Li et al. [15], while the latter contains lollipops, lariats and dumbbells as specializations. Using the composition method, we also establish the -positivity of hats. The family of hats contains both tadpoles and generalized bulls. Our result for hats induces a second -expansion for tadpoles. The family of generalized bulls was listed as an infinite collection of -positive claw-free graphs that are not claw-contractible-free by Dahlberg et al. [5, Section ]. We also consider the line graphs of tadpoles, since the line graph of any graph is claw-free, which is a key condition in both Conjectures 1.1 and 1.2.
An early try of the composition method towards Schur positivity is [28], in which Thibon and Wang obtained the ribbon Schur expansion of a noncommutative analog for spiders of the form . They are not ribbon positive. This analog yields a skew Schur expansion of . By the Littlewood–Richardson rule, the ordinary Schur coefficients are by that means multiset sizes of Yamanouchi words, and the Schur positivity then follows by injections. A similar proof for the Schur positivity of spiders of the form is beyond uncomplicated. We thereby expect more satisfying applications of the composition method in establishing the Schur positivity of graphs. In this paper, we give a compact ribbon Schur analog for the chromatic symmetric function of cycles, see Theorem 3.2.
2. Preliminaries
This section contains necessary notion and notation, basic results on commutative symmetric functions, chromatic symmetric functions, and noncommutative symmetric functions, that will be of use.
2.1. Compositions and partitions
We use terminology from Stanley [25]. Let be a positive integer. A composition of is a sequence of positive integers with sum , commonly denoted . It has size , length , and reversal . The integers are called parts of . For notational convenience, we write if all parts have the same value , and denote the th last part as ; thus . We consider the number to have a unique composition, denoted . Whenever a capital letter such like and is adopted to denote a composition, we use the small letter counterparts such as and respectively with integer subscripts to denote the parts. A factor of is a subsequence that consists of consecutive parts. A prefix (resp., suffix) of is a factor that starts from (resp., ends at ). Denote by the the number of parts in , namely,
(2.1) |
A partition of is a multiset of positive integers with sum , commonly denoted as
where . For any composition , there is a unique partition satisfying Eq. 1.3, i.e., the partition obtained by rearranging the parts of . As partitions have Young diagrams as graphic representation, one uses the terminology ribbons to illustrate compositions. In French notation, the ribbon for a composition is the collection of boxes such that
-
•
Row consists of consecutive boxes, and
-
•
the last box on Row and the first box on Row are in the same column.
In the theory of integer partitions, by saying a Young diagram one emphasizes the geometric shape of the partition . Being analogous in our composition calculus, we phrase the wording “a ribbon ” to call attention to the illustration of the composition .
Following MacMahon [16], the conjugate of a composition is the ribbon consisting of the column lengths of from right to left. This is different to the conjugate of a partition , whose Young diagram is obtained by turning rows into columns. For example, and . A refinement of is a composition such that
for some integers , where and . We say that is a coarsement of if is a refinement of . The reverse refinement order for compositions is the partial order defined by
The first parts of blocks of with respect to are the numbers , with product
The last parts of blocks of with respect to are the numbers , with product
By definition, one may derive directly that
(2.2) |
For any compositions and , the concatenation of and is the composition , and the near concatenation of and is the composition
In French notation, the ribbon (resp., ) is obtained by attaching the first box of immediately below (resp., to the immediate right of) the last box of .
The decomposition of a ribbon relatively to a composition is the unique expression
where , each is a ribbon of size , and each symbol stands for either the concatenation or the near concatenation. For instance,
We call the ribbons blocks of . In the language of ribbons, the block consists of the first boxes of the ribbon that is obtained from by removing the previous blocks .
A hook is a ribbon of the form for some and . Every hook appears as the English letter L or a degenerate one, that is, a horizontal ribbon or a vertical ribbon . Here we recognize the ribbon as horizontal. Denote by the set of ribbons such that every block in the decomposition is a hook. Then
is the set of hooks of the composition consisting of a single part. Moreover, since every factor of a hook is still a hook, we have for all . For example, , , and . Let . By definition, the set is in a bijection with the set
where the symbol stands for the concatenation operation. As a consequence, one may calculate .
2.2. Commutative symmetric functions
We give an overview of necessary notion and notation for the theory of commutative symmetric functions. For comprehensive references, one may refer to Stanley [26] and Mendes and Remmel [18]. Let be a commutative ring with identity. A symmetric function of homogeneous degree over is a formal power series
such that for any permutation . Denote by the field of rational numbers. Define , and define to be the vector space of homogeneous symmetric functions of degree over . Common bases of include the elementary symmetric functions , the complete homogeneous symmetric functions , the power sum symmetric functions , and the Schur symmetric functions . The first three ones are multiplicatively defined by
where
The Schur symmetric function can be defined combinatorially by , where is the set of column strict tableaux of shape , and the weight is the product of for all entries in . Here a tableau of shape is said to be column strict if
-
•
the entries in each row weakly increase, and
-
•
the entries in each column strictly increase starting from the longest row; this is to say from bottom to top in French notation.
The Schur symmetric functions are said to be “the most important basis for with respect to its relationship to other areas of mathematics” and “crucial in understanding the representation theory of the symmetric group,” see [18, Page 37].
For any basis of and any symmetric function , the -coefficient of is the unique number such that , denoted . The symmetric function is said to be -positive if every -coefficient of is nonnegative. For instance, every elementary symmetric function is Schur positive since , where are Kostka numbers, see [18, Exercise 2.12].
With the aid of the function defined by Eq. 1.3, one may extend the domain of these basis symmetric functions from partitions to compositions. Precisely speaking, one may define for any composition and any basis . With this convention, we are safe to write instead of the redundant expression . Since is not a basis of , the notation is undefined.
2.3. Chromatic symmetric functions
Stanley [23] introduced the chromatic symmetric function for a graph as
where is a countable list of indeterminates, and runs over proper colorings of . Chromatic symmetric functions are particular symmetric functions, and it is a generalization of Birkhoff’s chromatic polynomials , since . For instance, the chromatic symmetric function of the complete graph is
(2.3) |
We will need the -expansion of , see [23, Theorem 2.5].
Proposition 2.1 (Stanley).
The chromatic symmetric function of a graph is
where is the partition consisting of the component orders of the spanning subgraph .
By [18, Theorem 2.22], every -coefficient in a power sum symmetric function is an integer. It then follows from Proposition 2.1 that every -coefficient of is integral. Stanley [23, Corollary 3.6] presented the following quick criterion for the -positivity.
Proposition 2.2 (Stanley).
Any graph whose vertices can be partitioned into two cliques is -positive.
Such graphs have several characterizations, such as the complements of bipartite graphs and the incomparability graphs of -free posets, see Guay-Paquet [12, Theorem 5.3]. Stanley [23, Propositions 5.3 and 5.4] confirmed the -positivity of paths and cycles.
Proposition 2.3 (Stanley).
Let and . Denote by the -vertex path and by the -vertex cycle. Then
As a consequence, paths and cycles are -positive.
Explicit formulas for the -coefficients of and were obtained by extracting the coefficients of these generating functions, see Wolfe [32, Theorem 3.2]. Shareshian and Wachs [22] obtained the much simpler Eq. 1.1 for paths. Ellzey [6, Corollary 6.2] gave a formula for the chromatic quasisymmetric function of cycles, whose specialization is an equally simple one.
Proposition 2.4 (Ellzey).
For , .
We provide a proof for Proposition 2.4 using the composition method in Appendix A. Orellana and Scott [19, Theorem 3.1, Corollaries 3.2 and 3.3] established the triple-deletion property for chromatic symmetric functions.
Theorem 2.5 (Orellana and Scott).
Let be a graph with a stable set of order . Denote by , and the edges linking the vertices in . For any set , denote by the graph with vertex set and edge set . Then
2.4. Noncommutative symmetric functions
For an introduction and basic knowledge on noncommutative symmetric functions, see Gelfand et al. [11]. Let be a field of characteristic zero. The algebra of noncommutative symmetric functions is the free associative algebra generated by an infinite sequence of indeterminates over , where . It is graded by the weight function . The homogeneous component of weight is denoted . Let be an indeterminate that commutes with all indeterminates . The elementary symmetric functions are themselves, whose generating function is denoted by
The complete homogeneous symmetric functions are defined by the generating function
The power sum symmetric functions of the first kind are defined by the generating function
For any composition , define
The algebra is freely generated by any one of these families. Here the superscript notation are adopted to indicate that the functions are multiplicative with respect to composition concatenations. The sign of is defined by
(2.4) |
It is direct to check that
(2.5) |
Another linear basis of is the ribbon Schur functions , which can be defined by
see [11, Formula (62)]. We list some transition rules for these bases, see [11, Propositions 4.15 and 4.23, and Note 4.21].
Proposition 2.6 (Gelfand et al.).
For any composition , we have
(2.6) | ||||
(2.7) | ||||
(2.8) |
where are the composition blocks of the decomposition .
Equation 2.7 is true by virtue of Eq. 2.2, though it was expressed in terms of the product in [11]. Recall from Eq. 1.3 that maps a composition to its underlying partition. We use the same notation to denote the projection map defined by and by extending it linearly. By definition, for any composition ,
where is the skew partition of shape . For instance,
When for some and , we say that is the commutative image of , and that is a noncommutative analog of . For instance, Eqs. 1.1 and 2.4 imply that and have the noncommutative analogs
(2.9) | ||||
(2.10) |
respectively. If a chromatic symmetric function has a noncommutative analog , then for any partition ,
The aforementioned ambiguity issue is solved naturally in the language of the algebra . Indeed, since is a basis of , we talk about the well defined -coefficients instead of the undefined “-coefficients”.
By definition, any chromatic symmetric function has an infinite number of noncommutative analogs, among which only a finite number with integer coefficients are -positive. In particular, if a symmetric function is -positive, then the analog is -positive. Therefore, a symmetric function is -positive if and only if it has a -positive noncommutative analog. Therefore, in order to prove that a graph is -positive, it suffices to find a -positive analog of . The algebra plays the role of providing theoretical support for the composition method. As a consequence, we display only positive -expansions in theorem statements. We would not write in terms of noncommutative analogs except when arguing -coefficients is convenient.
2.5. Warming up for the composition method
This section consists of a property of the function defined by Eq. 1.2, some other composition functions and their interrelations, as well as some practices of using these functions.
From definition, it is straightforward to see that for any composition that is obtained by rearranging the non-first parts of . Another five-finger exercise is as follows.
Lemma 2.7.
Let and be nonempty compositions such that . Then
for any composition that is obtained by rearranging the parts of such that .
Proof.
Direct by Eq. 1.2. ∎
For any number , we define the surplus partial sum of with respect to to be the number
(2.11) |
Define the -surplus of to be the number
(2.12) |
Then . The function will appear in Theorem 3.3. Here is a basic property.
Lemma 2.8.
Let and . If , then .
Proof.
This is transparent if one notices . ∎
Lemma 2.8 will be used in the proof of Theorem 3.10. Similarly, for any number , we define the deficiency partial sum of with respect to to be the number
(2.13) |
and define the -deficiency of to be the number
(2.14) |
Then . The function (resp., ) can be expressed in terms of (resp., ).
Lemma 2.9.
Let and . Then
(2.15) |
or equivalently,
(2.16) |
Proof.
We shall show Eq. 2.15 first. If , then and , satisfying Eq. 2.15. Suppose that , and
(2.17) |
Then . By Eq. 2.13,
Subtracting from by each sum in the above inequality, we obtain
which reads, . Adding it up with Eq. 2.17, we obtain the sum as desired. This proves Eq. 2.15. Using Eqs. 2.12 and 2.14, one may infer Eq. 2.16 from Eq. 2.15. This completes the proof. ∎
Lemma 2.9 will be used in the proof of Theorem 3.9. Let us express the product in terms of the functions and .
Lemma 2.10.
For and ,
(2.18) | ||||
(2.19) |
Lemma 2.11.
For ,
(2.20) |
Dually, for ,
(2.21) |
Proof.
By Eq. 2.18, the convolution on the left hand of Eq. 2.20 has a noncommutative analog
Combining it with Proposition 2.4, we obtain
The coefficient of of the first sum on the right side is the partial sum such that
that is, the sum . This proves Eq. 2.20. In the same fashion, one may show Eq. 2.21. ∎
We need the noncommutative setting in the proof above since the coefficient of is considered. Lemma 2.11 will be used in the proof of Theorem 3.3 for tadpoles.
Corollary 2.12.
For , the average of the full convolution of chromatic symmetric functions of paths and cycles with total order is the chromatic symmetric function of the path of order , i.e.,
It can be shown alternatively by taking in Eq. 2.21 and using Proposition 2.4 and the identity , or, by Proposition 2.3.
3. Neat formulas for some chromatic symmetric functions
In this section, we use the composition method to produce neat formulas for the chromatic symmetric functions of several families of graphs, including tadpoles and their line graphs, barbells, and generalized bulls. We also establish the -positivity of hats.
3.1. The ribbon expansion for cycles
In view of Eq. 2.6, if a noncommutative symmetric function is -positive, then it is -positive. Thibon and Wang [28] discovered that the analog has the rather simple ribbon expansion
We present a -expansion for a noncommutative analog of cycles.
Lemma 3.1.
Proof.
Let be the cycle with vertices arranged counterclockwise. Let . The contribution of the edge set in Proposition 2.1 is . When , the graph consists of paths. Let be the order of the path containing . Then . Let be the orders of paths counterclockwise in the sequel. Since the path containing has possibilities:
we can deduce by Proposition 2.1 that
Since , has the desired analog. ∎
Now we can produce a ribbon Schur analog of .
Theorem 3.2.
The chromatic symmetric function of cycles has a noncommutative analog
where and are the last and second last part of respectively, is defined by Eq. 2.1, and is the maximum number of parts that start .
Proof.
Recall that is the set of ribbons such that every block in the decomposition is a hook. By Eq. 2.8, we can rewrite the formula in Lemma 3.1 as
(3.1) |
where is the set of decompositions such that every block in is a hook. Here each bullet is either the concatenation or the near concatenation. It is direct to compute
Below we consider such that .
We introduce a sign-reversing involution to simplify the inner sum in Eq. 3.1. Let
For any box in the ribbon , denote
-
•
by the hook in that contains , and
-
•
by the box lying to the immediate right of , if it exists.
We call the right neighbor of . We say that a box of is an active box of if
-
•
its right neighbor exists,
-
•
, and
-
•
the union of boxes is a hook.
Let be the set of decompositions that contain an active box. We define a transformation on as follows. Let . Let be the last active box of . Define to be the decomposition obtained from by
-
•
dividing into two hooks which contain and respectively, if ;
-
•
merging and into a single hook, if .
From definition, we see that is an involution. In view of the sign of the inner sum in Eq. 3.1, we define the sign of to be . Then becomes sign-reversing as
As a result, the contribution of decompositions in to the inner sum in Eq. 3.1 is zero, and for the inner sum can be replaced with the set
of decompositions of without active boxes.
First of all, we shall show that
Let be a hook and . Let . Then has no active boxes. In particular, the second last box of is not active. It follows that and
where . Therefore, by Eq. 3.1,
Below we can suppose that is not a hook. Then the subtrahend in Eq. 3.1 vanishes, and Eq. 3.1 implies that
(3.2) |
Second, we claim that unless . In fact, if , then the second last box of is active for any decomposition . Thus
This proves the claim. It follows that
Denote the last box on the horizontal part by . We say that a box of is a leader of a decomposition if it is the first box of some hook of length at least in .
Third, we claim that
Let . If , then the third last box in is active for any , which implies as before. This proves the claim. Moreover, if is not a leader for some , then the second last box in is active in , contradicting the choice of . Therefore, by Eq. 3.2,
(3.3) |
Fourth, we shall show that
Suppose that and . Let be the th last box in . In particular, . We observe that since otherwise it would be active. Moreover, if ends with , then must be a leader of , since otherwise would be active. To sum up, we are left to cases:
-
(1)
ends with , , and is a leader,
-
(2)
ends with ,
-
(3)
ends with .
Let . The classification above allows us to transform Eq. 3.3 to
(3.4) |
For , let be the column of boxes in that contains . Then
For , we observe that is the union of several blocks in . Conversely, since is a leader, , and there are ways to decompose to form the blocks of some . Computing various cases for in the same vein, we can deduce from Eq. 3.4 that
Note that each of the terms in the parenthesis holds true even for when .
Fifth, let us compute the -coefficient for
If , then must be a leader, since otherwise would be active. Since every vertical hook has sign , we can deduce from Eq. 3.3 that
Computing the number of decompositions in for each of the sums above, we derived that
which is true even for . Note that , and
holds as an identity. Therefore,
Finally, collecting the coefficients above, we obtain
(3.5) |
which can be recast as the desired formula. ∎
In view of Eq. 3.5, every -coefficient is nonnegative. For instance,
3.2. Tadpoles and their line graphs
For and , the tadpole is the graph obtained by connecting a vertex on the cycle and an end of the path . By definition,
See Fig. 1 for the tadpole and its line graph .
Li et al. [15, Theorem 3.1] pointed out that tadpoles possess Gebhard and Sagan’s -positivity, which implies the -positivity. They gave the chromatic symmetric function
(3.6) |
in their formula (3.11). By investigating the analog , Wang and Wang [31, Theorem 3.2] obtained the -positivity of the line graphs , which implies the -positivity of the graphs and . They [31, Formulas (3.2) and (3.3)] also obtained the formulas
(3.7) | ||||
(3.8) |
One may deduce Theorem 3.3 alternatively by using Eqs. 2.21, 3.6 and 2.15. The tadpole is called an -pan. For example, the -pan has the chromatic symmetric function
We remark that Theorem 3.3 reduces to Eq. 1.1 when , and to Proposition 2.4 when .
A lariat is a tadpole of the form . Dahlberg and van Willigenburg [3] resolved conjectures of Wolfe [32] on by analyzing Eq. 3.6. We now bring out a neat formula for , which implies effortless resolutions of the conjectures.
Corollary 3.4 (Lariats).
For , we have .
Proof.
Direct by taking in Theorem 3.3. ∎
The line graphs of tadpoles also admit simple analogs.
Theorem 3.5 (The line graphs of tadpoles).
Proof.
For example, we have
The noncommutative setting in the proof above is adopted since -coefficients are considered.
3.3. Barbells
In other words, the -chain can be obtained from a sequence , , , , of cliques such that and share one vertex, and that the shared vertices are distinct. The number of vertices and edges of are respectively
For instance, and . The family of -chains contains many special graphs.
-
(1)
A lollipop is a -chain of the form . A lariat is a lollipop of the form .
-
(2)
A barbell is a -chain of the form . A dumbbell is a barbell of the form .
-
(3)
A generalized bull is a -chain of the form .
Tom [29, Theorem 2] gave a formula for the chromatic symmetric function of melting lollipops, with lollipops as a specialization.
Theorem 3.6 (Lollipops, Tom).
Let . Then .
Theorem 3.6 covers Eqs. 1.1, 2.3 and 3.4. It can be derived alternatively by using Eqs. 2.3 and 1.1 and
which is due to Dahlberg and van Willigenburg [3, Proposition 9].
Using Dahlberg and van Willigenburg’s method of discovering a recurrence relation for the chromatic symmetric functions of lollipops, we are able to handle barbells.
Theorem 3.7 (Barbells).
Proof.
Fix and . See Fig. 3.
For , the graph reduces to a lollipop, and the desired formula reduces to Theorem 3.6. Below we can suppose that . We consider a graph family
defined as follows. Define . For , define to be the graph obtained from by removing the edges , , . In particular,
-
•
, and
-
•
is the disjoint union of the lollipop and the complete graph .
By applying Theorem 2.5 for the vertex triple in , we obtain
Therefore, one may deduce iteratively that
Then we can deduce by bootstrapping that
By Eqs. 2.3 and 3.6, we obtain
(3.10) |
We can split it as
(3.11) |
where is the part containing , and the part without . Let
(3.12) | ||||
(3.13) |
Then and
From Eq. 3.10, we obtain
Considering in the negative part. When runs from to and runs over compositions such that , runs over all compositions in . Since
we can deduce that
(3.14) |
On the other hand, by Eq. 3.10, we find
Similarly, we consider in the negative part. When runs from to and runs over compositions such that , runs over all compositions in . Note that
where . Therefore,
where
Note that the involution defined for the compositions such that by exchanging the first two parts satisfies . Therefore,
In view of Eq. 3.12, the last sum can be recast by considering the possibility of as
in which the negative part is exactly by Eq. 3.14. Therefore,
Hence by Eqs. 3.11 and 3.13, we obtain the formula as desired. ∎
For example,
We remark that Theorem 3.7 reduces to Theorem 3.6 when . In view of the factor in Theorem 3.7, we do not think it easy to derive Theorem 3.7 by applying Tom’s -chain formula to barbells. The next two formulas for the graphs and dumbbells are particular cases of Tom’s -chain formula. They are straightforward from Theorem 3.7.
Corollary 3.8 (Tom).
Let and . Then
Proof.
In Theorem 3.7, taking and yields the first formula, while taking and yields the second. ∎
We remark that the -positivity of the graphs and are clear from Proposition 2.2. On the other hand, in Corollary 3.8, taking in the first formula and taking in the second result in the same formula
3.4. Hats and generalized bulls
A hat is a graph obtained by adding an edge to a path. Let
The hat is the graph obtained from the path by adding the edge , see Fig. 4. It is a unicyclic graph with the cycle length . By definition,
It is clear that is isomorphic to . In particular, the hat is the tadpole , the hat is a path with a repeated edge, and the hat is the generalized bull .
Computing , we encounter the chromatic symmetric function of spiders with legs. For any partition , the spider is the tree of order obtained by identifying an end of the paths , , , see Fig. 5 for an illustration of .
Zheng [34, Lemma 4.4] showed that for any multiset and ,
(3.15) |
For proving the -positivity of hats, we introduce a special composition bisection defined as follows. For any composition of size at least , we define a bisection by
It is possible that is empty. A key property of this bisection is the implication
(3.16) |
Theorem 3.9.
Every hat is -positive.
Proof.
Let . Since is -positive, we can suppose that . Let . When , applying Theorem 2.5 for the triangle in Fig. 6, we obtain
(3.17) |
By adding Eq. 3.17 for the parameter from to the value , we obtain
Substituting Eq. 3.15 for spiders into the formula above, we deduce that
Substituting Eq. 1.1 for paths and Theorem 3.3 for tadpoles into it, we obtain
(3.18) |
Note that the upper (reps., lower) bound for in the second (resp., third) sum can be replaced with (resp., ). As a consequence, one may think the last two sums run as for the same set of pairs . By Lemma 2.9, we can merge their coefficients of as
Therefore, we can rewrite Eq. 3.18 as
(3.19) |
where , ,
One should note the following facts:
-
•
When , it is possible that is the empty composition.
-
•
for any .
-
•
for any . Moreover, together with Eq. 3.18, one may infer that
(3.20)
We will deal with the cases and respectively. Let
Let . We shall show that the map defined by
is a bijection from to the set
Before that, it is direct to check by definition that
(3.21) | ||||
Therefore, if the bijectivity is proved, then we can simplify Eq. 3.19 to
(3.22) |
where .
In order to establish the bijectivity of , we need to prove that
-
(1)
,
-
(2)
is injective, and
-
(3)
is surjective. for any , there exists such that .
We proceed one by one. Item 1 If we write , then by the implication (3.16),
(3.23) |
Let us check by definition:
-
•
since ;
-
•
since ;
-
•
; and
-
•
.
Item 3 Let . Consider . By the implication (3.16), we obtain Eq. 3.23. Thus . It remains to check that :
-
•
since .
-
•
.
-
•
.
- •
This proves that is bijective.
It remains to deal with the case . Continuing with Eq. 3.22, we decompose as
where
We remark that the bound restriction in is to guarantee that is not trivial:
In fact, the restriction implies ; conversely, if , then has no prefix such that . This proves the equivalence relation.
Now, fix . Let
Then the sets for are disjoint. In fact, if
then and have the same prefix by the implication (3.16), the same suffix , and the same middle part ; thus . The pairs for the first sum in Eq. 3.22 that we do not use to cancel the second sum form the set
Since for any and , Eq. 3.22 can be recast as
(3.24) |
where
Hence it suffices to show that .
Let . Then for all since . For , we define
Then by the implication (3.16),
(3.25) | ||||
We observe that
-
•
, since ; and
-
•
, since .
Therefore,
(3.26) |
where
Let us compare with , and compare with , respectively.
It follows that
(3.28) | ||||
(3.29) |
This sum in Eq. 3.29 can be simplified by telescoping. Precisely speaking, since th the negative term and the th positive term have sum
we can simplify the sum in Eq. 3.29 by keeping the first positive term and the last negative term as
Together with Eq. 3.28, we can infer that when ,
(3.30) |
In view of Eq. 3.28, we see that Eq. 3.30 holds for as well. Note that
Here we have two cases to deal with. If , then
(3.31) |
If , then by Eqs. 3.30 and 3.27,
It follows that
(3.32) |
This completes the proof. ∎
By carefully collecting all terms of along the proof of Theorem 3.9, and combinatorially reinterpreting the coefficients and bound requirements, we can assemble a positive -expansion for the chromatic symmetric function of hats.
Theorem 3.10 (Hats).
Let , where and . Then
(3.33) | ||||
where , and if we write as the bisection of such that ,
Proof.
We keep notion and notation in the proof of Theorem 3.9. Let . Then
The numerator and denominator in Eq. 3.31 can be recast as
(3.34) | ||||
(3.35) |
respectively. By Eqs. 3.31 and 3.32,
(3.36) | ||||
where
We claim that the right side of Eq. 3.36 can be simplified to and
(3.37) |
In fact, Eq. 3.37 is one of the original bound requirements. It suffices to show that also holds. Assume to the contrary that . Then
by the definition Eq. 2.12 of , contradicting Eq. 3.37. This proves the claim.
In view of Eq. 3.24, it remains to simplify
in which the summands have the same form . If a pair appears as in the first sum, then the requirement is equivalent to say that
and the requirement is equivalent to . Thus the set of pairs for the first sum is
On the other hand,
Since the product vanishes when , we can replace the conditions for with . Furthermore,
The sum for over this subset can be merged into the second sum as
this is because when ,
Collecting all the contributions to , we obtain Eq. 3.33 as desired. ∎
For example,
Particular hats are special graphs that we explored previously.
-
(1)
For , Theorem 3.10 reduces to Theorem 3.3 since only the second sum in Eq. 3.33 survives.
-
(2)
For , Theorem 3.10 reduces to Eq. 1.1, since only the first two sums in Eq. 3.33 survive, and they are the sum of terms for and for respectively.
-
(3)
For , Theorem 3.10 may give a noncommutative analog for the tadpole that is different from the one given by Theorem 3.3. For instance, these two analogs for are respectively
For , the hat is the generalized bull . We produce for generalized bulls a neat formula, which is not a direct specialization of Theorem 3.10.
Theorem 3.11 (Generalized bulls).
For and ,
Proof.
We shall simplify and separately. For , we proceed in steps. First, we claim that
(3.40) |
In fact, for the forward direction, if , then by Lemma 2.8, a contradiction. For the backward direction, we have two cases to deal with:
-
•
If , then holds trivially.
-
•
If , since , we then find and .
This proves the claim. It allows us to change the sum range for to
Second, for , we have and
Thirdly, we claim that
(3.41) |
In fact, recall from Eq. 3.35 that is a factor of . By the definition Eq. 3.25 of , the part is the part of such that
Since , we find . For any , define to be the composition obtained from by moving the part to the end, i.e., . Then , , and
Thus , and . Since , we find . Therefore, can be recovered from by moving the last part to the position immediately after . Hence is a bijection on , and
This proves the claim. We can strengthen by requiring without loss of generality.
We remark that Theorem 3.11 reduces to Corollary 3.4 when .
Appendix A A proof of Proposition 2.4 using the composition method
By Eqs. 2.7 and 2.5, we can deduce from Lemma 3.1 that
Let . Then any composition of length that is finer than can be written as
for some indices , where and . Therefore,
This proves Proposition 2.4.
Acknowledgment
We are grateful to Jean-Yves Thibon, whose expectation for noncommutative analogs of chromatic symmetric functions of graphs beyond cycles encourages us to go on the discovery voyage for neat formulas. We would like to thank Foster Tom for pointing us to Ellzey’s paper, and for presenting his signed combinatorial formula in an invited talk. We also thank the anonymous reviewer for sharing his/her understanding of the motivation and for the writing suggestions.
References
- Aliniaeifard et al. [2024] F. Aliniaeifard, V. Wang, and S. van Willigenburg. The chromatic symmetric function of a graph centred at a vertex. Electron. J. Combin., 31(4):#P4.22, 2024.
- Corneil et al. [2010] D. G. Corneil, S. Olariu, and L. Stewart. The LBFS structure and recognition of interval graphs. SIAM J. Discrete Math., 23(4):1905–1953, 2010.
- Dahlberg and van Willigenburg [2018] S. Dahlberg and S. van Willigenburg. Lollipop and lariat symmetric functions. SIAM J. Discrete Math., 32(2):1029–1039, 2018.
- Dahlberg and van Willigenburg [2020] S. Dahlberg and S. van Willigenburg. Chromatic symmetric functions in noncommuting variables revisited. Adv. in Appl. Math., 112:101942, 25, 2020.
- Dahlberg et al. [2020] S. Dahlberg, A. Foley, and S. van Willigenburg. Resolving Stanley’s -positivity of claw-contractible-free graphs. J. European Math. Soc., 22(8):2673–2696, 2020.
- Ellzey [2017] B. Ellzey. Chromatic quasisymmetric functions of directed graphs. Sém. Lothar. Combin., 78B:Art. 74, 12, 2017.
- Faudree et al. [1997] R. Faudree, E. Flandrin, and Z. Ryjáček. Claw-free graphs — a survey. Discrete Math., 164(1):87–147, 1997. The second Kraków conference of graph theory.
- Gasharov [1996] V. Gasharov. Incomparability graphs of -free posets are -positive. In Proceedings of the 6th Conference on Formal Power Series and Algebraic Combinatorics (New Brunswick, NJ, 1994), volume 157, pages 193–197, 1996.
- Gasharov [1999] V. Gasharov. On Stanley’s chromatic symmetric function and clawfree graphs. Discrete Math., 205(1–3):229–234, 1999.
- Gebhard and Sagan [2001] D. D. Gebhard and B. E. Sagan. A chromatic symmetric function in noncommuting variables. J. Alg. Combin., 13(3):227–255, 2001.
- Gelfand et al. [1995] I. Gelfand, D. Krob, A. Lascoux, B. Leclerc, V. Retakh, and J. Thibon. Noncommutative symmetrical functions. Adv. Math., 112(2):218–348, 1995.
- Guay-Paquet [2013] M. Guay-Paquet. A modular relation for the chromatic symmetric functions of (3+1)-free posets. arXiv: 1306.2400, 2013.
- Haiman [1993] M. Haiman. Hecke algebra characters and immanant conjectures. J. Amer. Math. Soc., 6(3):569–595, 1993.
- Huh et al. [2020] J. Huh, S.-Y. Nam, and M. Yoo. Melting lollipop chromatic quasisymmetric functions and Schur expansion of unicellular LLT polynomials. Discrete Math., 343(3):111728, 2020.
- Li et al. [2021] E. Y. H. Li, G. M. X. Li, D. G. L. Wang, and A. L. B. Yang. The twinning operation on graphs does not always preserve -positivity. Taiwanese J. Math., 25(6):1089–1111, 2021.
- MacMahon [1915, 1916] P. A. MacMahon. Combinatory Analysis, volume I and II. Camb. Univ. Press, Cambridge, 1915, 1916.
- Martin et al. [2008] J. L. Martin, M. Morin, and J. D. Wagner. On distinguishing trees by their chromatic symmetric functions. J. Combin. Theory Ser. A, 115(2):237–253, 2008.
- Mendes and Remmel [2015] A. Mendes and J. Remmel. Counting with Symmetric Functions, volume 43 of Dev. Math. Springer, Cham, 2015.
- Orellana and Scott [2014] R. Orellana and G. Scott. Graphs with equal chromatic symmetric functions. Discrete Math., 320:1–14, 2014.
- Shareshian and Wachs [2010] J. Shareshian and M. L. Wachs. Eulerian quasisymmetric functions. Adv. Math., 225(6):2921–2966, 2010.
- Shareshian and Wachs [2012] J. Shareshian and M. L. Wachs. Chromatic quasisymmetric functions and Hessenberg varieties. In Configuration spaces, volume 14 of CRM Series, pages 433–460. Ed. Norm., Pisa, 2012.
- Shareshian and Wachs [2016] J. Shareshian and M. L. Wachs. Chromatic quasisymmetric functions. Adv. Math., 295:497–551, 2016.
- Stanley [1995] R. P. Stanley. A symmetric function generalization of the chromatic polynomial of a graph. Adv. Math., 111(1):166–194, 1995.
- Stanley [1998] R. P. Stanley. Graph colorings and related symmetric functions: ideas and applications: a description of results, interesting applications, & notable open problems. Discrete Math., 193(1-3):267–286, 1998.
- Stanley [2011. It is a thoroughly revised version of the 1st edition published in 1986.] R. P. Stanley. Enumerative Combinatorics, Vol. 1, volume 49 of Cambridge Stud. in Adv. Math. Camb. Univ. Press, Cambridge, 2nd edition, 2011. It is a thoroughly revised version of the 1st edition published in 1986.
- Stanley [2023. The main difference with the 1st ed. published in 1999 is 159 additional exercises and solutions on symmetric functions.] R. P. Stanley. Enumerative Combinatorics. Vol. 2, volume 62 of Cambridge Stud. in Adv. Math. Camb. Univ. Press, Cambridge, 2nd edition, 2023. The main difference with the 1st ed. published in 1999 is 159 additional exercises and solutions on symmetric functions.
- Stanley and Stembridge [1993] R. P. Stanley and J. R. Stembridge. On immanants of Jacobi-Trudi matrices and permutations with restricted position. J. Combin. Theory Ser. A, 62(2):261–279, 1993.
- Thibon and Wang [2023] J.-Y. Thibon and D. G. L. Wang. A noncommutative approach to the schur positivity of chromatic symmetric functions. arXiv:2305.07858, 2023.
- Tom [2024] F. Tom. A signed -expansion of the chromatic symmetric function and some new -positive graphs. In Proceedings of the 36th Conference on Formal Power Series and Algebraic Combinatorics, Sém. Lothar. Combin., volume 91B, Article #48, 12 pp., Bochum, 2024.
- Wang and Wang [2023a] D. G. L. Wang and M. M. Y. Wang. The -positivity and schur positivity of some spiders and broom trees. Discrete Appl. Math., 325:226–240, 2023a.
- Wang and Wang [2023b] D. G. L. Wang and M. M. Y. Wang. The -positivity of two classes of cycle-chord graphs. J. Alg. Combin., 57(2):495–514, 2023b.
- Wolfe [1998] M. Wolfe. Symmetric chromatic functions. Pi Mu Epsilon J., 10(8):643–657, 1998.
- Wolfgang III [1997] H. L. Wolfgang III. Two Interactions between Combinatorics and Representation Theory: Monomial Immanants and Hochschild Cohomology. PhD thesis, MIT, 1997.
- Zheng [2022] K. Zheng. On the -positivity of trees and spiders. J. Combin. Theory Ser. A, 189:105608, 2022.