<<

arXiv:0911.5278v1 [math-ph] 27 Nov 2009 aclto fresultant of Calculation 3 Objects Basic 2 Introduction 1 Contents practical be resul could these which of , overview for brief formulas a explicit present We lo field. a this in studied: known well rather course of are resultants n research, with or forms non-quadratic with , () Conclusion 6 applications and Integral 5 of Calculation 4 ∗ † . omlso yvse ye...... type Sylvester of Formulas 3.1 eutnsaegtigicesnl motn nmdr th modern in important increasingly getting are Resultants . vlaino ra ...... series . . . hypergeometric . as . . . . . . . . Algebraic ...... . 5.4 . . of . . . Evaluation ...... 5.3 ...... Action-independence . identities . . . . Ward . . and . . . 5.2 . integrals . . . . Non-Gaussian ...... 5.1 ...... case . . . . . symmetric . . . The . . . . polytopes . . Newton . . 4.3 . . through . . Discriminant ...... invariants 4.2 . . through . Discriminant ...... 4.1 ...... type . Schur of . type Formulas Bezout . + . Sylvester 3.4 of Formulas type Bezout 3.3 of Formulas 3.2 TP ocw Russia Moscow, ITEP, [email protected] Russia; Moscow, ITEP, .. aalysfrua...... complex formula Koszul . Macaulay’s of . a 3.1.3 as Sylvester of determinant a 3.1.2 as Resultant 3.1.1 e n l eut nRslatTheory Resultant in Results Old and New and IT ogpun,Russia Dolgoprudny, MIPT, A.Morozov yueu naftr hsc research. physics future a in useful ly fepii omls euiu rprisadintriguin and properties beautiful formulas, explicit of t ABSTRACT nGusa nerl.Bigasbeto oeta three-h than more of subject a Being integrals. on-Gaussian ∗ s nldn ohrcn n led lsia.Epai i Emphasis classical. already and recent both including ts, n Sh.Shakirov and oeia hsc:te perweee n el ihnon- with deals one whenever appear they physics: eoretical 1 [email protected] ; 42 ...... 36 ...... † 6 ...... 33 ...... 27 ...... 8 ...... 19 ...... 6 ...... 38 ...... 39 ...... 22 . . . . . 15 ...... 33 ...... 14 ...... ITEP/TH-68/09 eainhp are relationships g undred-year aeon made s linear 48 36 27 6 4 2 1 Introduction

This paper is devoted to resultants – natural and far-going generalization of from conventional linear to generic non-linear systems of algebraic equations

f1(x1,...,xn)=0   f2(x1,...,xn)=0     . . .  f (x ,...,x )=0  n 1 n     1  where f1,...,fn are homogeneous of degrees r1,...,rn. This system of n polynomial equations in n variables is over-defined: in general position it does not have non-zero solutions (x1,...,xn) at all. A non-vanishing solution exists if and only if a single algebraic constraint is satisfied,

R f1,f2,...,fn =0 (1) where R is an called resultant, depending on coefficients of the non-linear system under consideration. Clearly, for linear equations resultant is nothing but the determinant – a familiar object of linear . For non-linear equations, resultant provides a natural generalization of the determinant and therefore can be considered as a central object of non- [1]. Historically, the foundations of resultant theory were laid in the 19th century by Cayley, Sylvester and Bezout [2]. In this period mostly the case of n = 2 variables was considered, with some rare exceptions. The theory of resultants for n > 2 and closely related objects, namely discriminants and , was significantly developed by Gelfand, Kapranov and Zelevinsky in their seminal book [3]. Let us emphasize, that determinant has a lot of applications in physics. Resultant is expected to have at least as many, though only a few are established yet. This is partly because today’s physics is dominated by methods and ideas of linear algebra. These methods are not always successful in dealing with problems, which require an essentially non-linear treatment. We expect that resultants and related algebraic objects will cure this problem and become an important part of the future non-linear physics and for entire string theory program in the sence of [4]. To illustrate this idea, let us briefly describe the relevance of resultants to non-Gaussian integrals – a topic of undisputable value in physics. It is commonly accepted, that quantum physics can be formulated in the language of functional integrals. It is hard to question the importance of this approach: since the early works of Wigner, Dirac and Feynman, functional integrals have been a source of new insights and puzzles. Originated in quantum mechanics, they have spread fastly in quantum field theory and finally developed into a widely used tool of modern theoretical physics. From the point of view of functional integration, any quantum theory deals with integrals of a form

f(x) = dx f(x) e−S(x) (2) 1In this paper we assume homogeneity of the polynomial equations to simplify our considerations; actually, the homogeneity condition can be relaxed and the whole theory can be formulated for non-homogeneous polynomials as well.

2 called correlators. The integration x belongs to a linear (which is often infinite-dimensional), and the choice of particular action S(x) specifies a particular theory or a model. In the simplest models, S(x) is merely a , which in some particular can be written as

S(x)= Sij xixj (3) i,j

If this is the case, the integral (2) is called Gaussian. In a sence, Gaussian integrals constitute a foundation of quantum theory: they describe free (non-interacting) particles and fields. These systems are relatively simple, and so are the Gaussian integrals. For example, the simplest two-dimensional Gaussian integral can be evaluated, say, by diagonalisation, and equals

2 2 const dx dy e−(ax +bxy+cy ) = (4) √b2 4ac − where the expression b2 4ac can be recognized as a determinant of the quadratic form ax2 + bxy + cy2. Other − Gaussian integrals, with more variables, with non-homogeneous action or with non-trivial insertion f(φ), are equally easy and similarly related to determinants – central objects of linear algebra, which describe consistency of linear systems of equations. Remarkably, the study of low-dimensional non-Gaussian integrals of degree three

S(x)= Sijkxixj xk (5) i,j,k and higher, reveals their relation to resultants: for example, a two-dimensional non-Gaussian integral equals

3 2 2 3 const dx dy e−(ax +bx y+cxy +dy ) = (6) √6 27a2d2 b2c2 18abcd +4ac3 +4b3d − − where the expression 27a2d2 b2c2 18abcd +4ac3 +4b3d can be recognized as a discriminant of the cubic − − form ax3 + bx2y + cxy2 + dy3, given by resultant of its derivatives:

R 3ax2 +2bxy + cy2, bx2 +2cxy +3dy2 = 27a2d2 b2c2 18abcd +4ac3 +4b3d (7) − −

In this way, resultants naturally appear in the study of low-dimensional non-Gaussian integrals. This connection between functional integration and non-linear algebra can be traced much further: section 5 of present paper provides a brief summary of this interesting direction of research. See also [5] for more details. In this paper we present a selection of topics in resultant theory in a more-or-less explicit manner, which is more convenient for a reader who needs to perform a practical calculation. This also makes our paper accessible to non-experts in the field. Most of statements are explained rather than proved, for some proofs see [3].

3 2 Basic Objects

Basic objects of study in non-linear algebra are generic finite-dimensional [1]. However, especially important in applications are symmetric tensors or, what is the same, homogeneous polynomials of degree r in n variables. In present paper, we restrict consideration to this case. It should be emphasized, that restriction to symmetric tensors is by no means an oversimplification: most of the essential features of non-linear algebra are present at this level. There are two different notations for homogeneous polynomials, which can be useful under different circumstances: notation

n

S(x1, x2,...,xn)= Si1,i2,...,ir xi1 xi2 ...xir i ,i ,...,i =1 1 2 r and notation

a1 a2 an S(x1, x2,...,xn)= sa1,a2...an x1 x2 ...xn a +a +...+a =r 1 2 n To distinguish between these notations, we denote coefficients by capital and small letters, respectively. A r homogeneous polynomial of degree r in n variables has Cn+r−1 independent coefficients. Resultant

A system of n polynomial equations, homogeneous of degrees r1,...,rn in variables x1,...,xn

f1(x1,...,xn)=0   f2(x1,...,xn)=0     . . .

 fn(x1,...,xn)=0    has non-zero solutions, if and only if its coefficients satisfy one algebraic constraint:

Rr ,...,r f ,...,fn =0 1 n { 1 }

The left hand side of this algebraic constraint is called a resultant of this system of equations. Resultant always exists and is unique up to rescaling. Resultant is a homogeneous polynomial of degree

1 1 deg R = r ...r + . . . + r1,...,rn 1 n r r 1 n in the coefficients of the system. Moreover, it is homogeneous in coefficients of each fi separately, and has degree r1 ...ri−1ri+1 ...rn in these coefficients. When all equations are linear, i.e, when ri = 1, resultant is equal to the determinant of this system: R ... f ,f ,...,fn = f f . . . fn. Resultant is defined 11 1{ 1 2 } 1 ∧ 2 ∧ ∧ only when the of homogeneous equations matches the number of variables: when these numbers differ, various more complicated quantities should be used to describe the existence and/or degeneracy of solutions. Some of these quantities are known as higher resultants or subresultants [6], some as complanarts [7].

4 Discriminant

A single homogeneous polynomial S(x1,...,xn) of degree r in n variables gives rise to a system of equations

∂S(x ,...,x ) 1 n =0 ∂x1    . . .   ∂S(x ,...,x ) 1 n =0  ∂x  n   which are all homogeneous of degree r 1 in variables x ,...,xn. Resultant of this system is called a discriminant − 1 of S and is denoted as Dn|r(S). Discriminant is a homogeneous polynomial of degree

deg D (S)= n(r 1)n−1 n|r − in the coefficients of S, what follows from the analogous formula for the generic resultant. When S is a quadratic form, i.e, when r = 2, discriminant is equal to its determinant: Dn|2(S) = det S. Invariant

A Lie of linear transformations

x1 G11 G12 ... G1n x1

 . . .   . . .   . . .  →              xn   Gn1 G12 ... Gnn   xn              with det G = 0 is called GL(n). Its with det G = 1 is called SL(n). In tensor notations, they act

xi Gij xj →

Thus they act on coefficients of homogeneous polynomials S(x1,...,xn) of degree r in n variables by the rule

Si ,...,i Gi j ...Gi j Sj ,...,j 1 r → 1 1 r r 1 r

Polynomials I(S) which map to theirselves under SL(n) transformations are called SL(n) invariants of S. In particular, discriminant Dn|r(S) is a SL(n) invariant of S. However, there are many more invariants which have no obvious relation to the discriminant. The number of functionally independent invariants of S is equal to the number of independent coefficients of S minus the dimension of SL(n), that is

# functionally independent invariants of S = Cr n2 +1 n+r−1 −

Therefore, invariants of given degree form a finite-dimensional linear space. All invariants form a .

5 3 Calculation of resultant 3.1 Formulas of Sylvester type

To find the resultant of n polynomial equations, homogeneous of degree r in variables x1,...,xn

f1(x1,...,xn)=0   f2(x1,...,xn)=0     . . .   fn(x1,...,xn)=0    it is possible to consider another system, multiplied by various :

s1 sn consider x1 ...xn fi = 0 instead of just fi =0

In the , coefficients of this system can be viewed as a rectangular nCq Cr+q matrix, q+n−1 × r+q+n−1 where q = s1 + . . . + sn is the multiplier’s degree. This matrix is called the (generalized) .

3.1.1 Resultant as a determinant of Sylvester matrix

q r+q In some cases it is possible to choose q in such a way, that Sylvester matrix is , i.e, nCq+n−1 = Cr+q+n−1. In these cases, resultant is equal to the determinant of Sylvester matrix. Looking closer at this numeric relation, it is easy to see that the only cases when it is possible are n = any, r = 1 and n = 2, r = any. In the first case, it suffices to choose q = 0 and Sylvester matrix equals the matrix of the system. This is the case of linear algebra. The second case is less trivial: it is necessary to choose q = r 1 to make the Sylvester matrix square. − For example, in the case n =2, r = 2 the system of equations has a form

2 2 f11x1 + f12x1x2 + f22x2 =0

 2 2  g11x1 + g12x1x2 + g22x2 =0  Multiplying the equations of this system by all monomials of degree q = 1, that is, by x1 and x2, we obtain

3 2 2 3 f11x1 + f12x1x2 + f22x1x2 +0x2 =0

 3 2 2 3  0x1 + f11x1x2 + f12x1x2 + f22x2 =0    3 2 2 3  g11x1 + g12x1x2 + g22x1x2 +0x2 =0  3 2 2 3  0x1 + g11x1x2 + g12x1x2 + g22x2 =0    These four polynomials are linear combinations of four monomials, i.e, give rise to a square Sylvester matrix:

f11 f12 f22 0

 0 f11 f12 f22  R2|2 f,g = det (8) { }  g11 g12 g22 0       0 g g g   11 12 22   

6 Formulas, which express resultant as a determinant of some auxillary matrix, are usually called determinantal. In the next case n =2, r = 3 the system of equations has a form

3 2 2 3 f111x1 + f112x1x2 + f122x1x2 + f222x2 =0

 3 2 2 3  g111x1 + g112x1x2 + g122x1x2 + g222x2 =0

 2 2 Multiplying the equations of this system by all monomials of degree q = 2, that is, by x1, x1x2 and x2, we get

5 4 3 2 2 3 4 5 f111x1 + f112x1x2 + f122x1x2 + f222x1x2 +0x1x2 +0x2 =0

 5 4 3 2 2 3 4 5  0x1 + f111x1x2 + f112x1x2 + f122x1x2 + f222x1x2 +0x2 =0    5 4 3 2 2 3 4 5  0x1 +0x1x2 + f111x1x2 + f112x1x2 + f122x1x2 + f222x2 =0    5 4 3 2 2 3 4 5  g111x1 + g112x1x2 + g122x1x2 + g222x1x2 +0x1x2 +0x2 =0  5 4 3 2 2 3 4 5  0x1 + g111x1x2 + g112x1x2 + g122x1x2 + g222x1x2 +0x2 =0    5 4 3 2 2 3 4 5  0x1 +0x1x2 + g111x1x2 + g112x1x2 + g122x1x2 + g222x2 =0    These four polynomials are linear combinations of four monomials, i.e, give rise to a square Sylvester matrix:

f111 f112 f122 f222 0 0

 0 f111 f112 f122 f222 0 

 0 0 f111 f112 f122 f222    R2|3 f,g = det   (9) { }  g111 g112 g122 g222 0 0       0 g111 g112 g122 g222 0       0 0 g111 g112 g122 g222      Note, that the matrix has a typical ladder form. The beauty and simplicity made this formula and corresponding resultant R2|r widely known. In our days, Sylvester method is the main apparatus used to calculate two- dimensional resultants and it is the only piece of resultant theory included in undergraduate text books and computer programs such as Mathematica and Maple. However, its generalization to n> 2 is not straightforward, since for n> 2 and r 2 Sylvester matrix is not square. ≥ It turns out that, despite Sylvester matrix in higher dimensions is rectangular, it is still related closely to the resultant. Namely, top-dimensional minors of the Sylvester matrix are always divisible by the resultant, if their size is bigger, than resultant’s degree (i.e, for big enough q). Various generalizations of the Sylvester method explicitly describe the other, non-resultantal, factor. Therefore, such generalized Sylvester formulas always express the resultant as a ratio of two polynomial quantities. This is certainly a drawback, at least from the computational point of view. Sylvester’s method posesses at least two different generalizations for n > 2: one homological, which leads to the theory of so-called Koszul complexes, another more combinatorial, which leads to the construction of so-called Macaulay matrices.

7 3.1.2 Resultant as a determinant of Koszul complex

In construction of the Koszul complex, one complements the commuting variables

x , x ,...,xn : xixj xj xi =0 1 2 − with anti-commuting variables

θ1,θ2,...,θn : θiθj + θj θi =0

and considers polynomials, depending both on x1,...,xn and θ1,...,θn. Denote through Ω(p, q) the space of such polynomials of degree p in x-variables and degree q in θ-variables. The degree q can not be greater than n, since θ are anti-commuting variables. Dimensions of these spaces are

p q dim Ω(p, q)= Cp+n−1Cn (10)

Koszul differential, built from f, is a linear operator which acts on these spaces by the rule

∂ ∂ dˆ= f1(x1,...,xn) + . . . + fn(x1,...,xn) (11) ∂θ1 ∂θn and is automatically nilpotent:

∂ ∂ ∂ ∂ ∂ ∂ dˆdˆ= f (x)f (x) = f (x)f (x) + =0 j k ∂θ ∂θ j k ∂θ ∂θ ∂θ ∂θ j k j k k j

It sends

dˆ: Ω(p, q) Ω(p + r, q 1) → − giving rise to the following Koszul complex:

dˆ dˆ dˆ dˆ Ω(p, q) Ω(p + r, q 1) Ω(p +2r, q 2) ... Ω(0, R) → − → − → →

Thus, for one and the same operator dˆthere are many different Koszul complexes, depending on a single parameter R – the x-degree of the rightmost space. Using the formula (10) for dimensions, it is easy to write down all of them. For example, the tower of Koszul complexes for the case 3 2 is |

8 R Spaces Dimensions Euler 0 Ω(0, 0) 1 1 1 Ω(1, 0) 3 3 2 Ω(0, 1) Ω(2, 0) 3 6 3 3 Ω(1, 1) → Ω(3, 0) 9 →10 1 3 2 → → | 4 Ω(0, 2) Ω(2, 1) Ω(4, 0) 3 18 15 0 5 Ω(1, 2) → Ω(3, 1) → Ω(5, 0) 9 → 30 → 21 0 6 Ω(0, 3) Ω(2→, 2) Ω(4→, 1) Ω(6, 0) 1 18→ 45→ 28 0 7 Ω(1, 3) → Ω(3, 2) → Ω(5, 1) → Ω(7, 0) 3 → 30 → 63 → 36 0 ...→ ...→ → → → ...→ ... where Euler characteristic is the alternating combination of dimensions. The rightmost differential

dˆ dˆ: Ω(R r, 1) Ω(R, 0) − → acts on the corresponding linear spaces as

i1 i2 in i1 i2 in x x ...x θj x x ...x fj x ,...,xn 1 2 n → 1 2 n 1 and is represented by nCR−r CR matrix. It is easy to see, that this is exactly the rectangular R−r+n−1 × R+n−1 Sylvester matrix. Therefore, Koszul complex in a sence complements the Sylvester matrix with many other matrices, which carry additional information. Resultant is equal to certain characteristic of this complex

R f ,...,fn = DET (d , d ,...,dp ) (12) n|r{ 1 } 1 2 −1

which is also called determinant, not to be confused with ordinary determinant of a . Here di are the rectangular matrices, representing particular terms of the complex. The ”determinantal” terminology was introduced by Cayley, who first studied DET (d1, d2,...,dp−1), and is rather unusual for modern literature, where DET (d1, d2,...,dp−1) is called a (Reidemeister) torsion of the complex d1,...,dp−1. Since the quantity itself is a rather natural generalisation of determinants from linear maps to complexes, we prefer to use the historical name ”determinant of a complex”. Calculation of determinants of complexes is described in some detail in [3, 8]. Given a complex

d1 d2 dp−1 L L ... Lp (13) 1 −→ 2 −→ −→ with vanishing Euler characteristic

p χ = dim L dim L + dim L . . . + ( 1) dim Lp =0 1 − 2 3 − −

one needs to select a basis in each linear space Li and label the basis vectors by elements of a set

9 1, 2,...,li

After the basis is chosen, the complex is represented by a sequence of rectangular matrices. The li li matrix × +1 di has li rows and li+1 coloums. To calculate the determinant of this complex, one needs to select arbitrary subsets σi 1, 2,...,li which consist of ⊂ i−1 i+k+1 σi = ( 1) lk − k=1 elements, i.e,

σ =0, σ = l , σ = l l , σ = l l + l , ... 1 2 1 3 2 − 1 4 3 − 2 1 and define conjugate subsets σi = 1, 2,...,li /σi which consist of σi = li σi = σi − +1 elements. Finally, one needs to calculate minors

Mi = the of di, which occupies rows σi and coloumns σi+1

The determinant of complex is then given by

p−1 + +1 (−1)p i Mp−1Mp−3 . . . DET (d1, d2,...,dp−1)= Mi = (14) Mp Mp . . . i=1 −2 −4

In fact, this answer does not depend on the choice of σi [8] and thus gives a practical way to calculate resultants of polynomial systems. Given a system of equations, one straightforwardly constructs the differential (11) and the linear spaces (10), calculates the matrices and their minors according to the above procedure and, finally, uses the formula (14) to find the resultant. The answer, obtained in this way, is excessively complicated in the following two senses: (i) the ratio of various determinants at the r.h.s. is actually reduced to a polynomial, but this cancelation is not explicit in (12); (ii) the r.h.s. is actually independent of the choice of Koszul complex, but this independence is not explicit in (12). Still, Koszul complex provides an explicit tool of calculation of resultants, which is reasonably useful for particular low n and r.

10 To illustrate the Koszul complex method, let us consider in detail the case n = 3, r = 2, a system of three quadratic equations in three unknowns. Let us denote their coefficients by a,b,c instead of usual f,g,h:

2 2 2 f1(x1, x2, x3)= a11x1 + a12x1x2 + a13x1x3 + a22x2 + a23x2x3 + a33x3

 2 2 2  f2(x1, x2, x3)= b11x1 + b12x1x2 + b13x1x3 + b22x2 + b23x2x3 + b33x3   2 2 2 f3(x1, x2, x3)= c11x1 + c12x1x2 + c13x1x3 + c22x2 + c23x2x3 + c33x3    By definition, Koszul differential is given by

∂ ∂ ∂ dˆ= f1 + f2 + f3 ∂θ1 ∂θ2 ∂θ3

Let us describe in detail the simplest case R = 4 and next-to-the-simplest case R = 5. The complex with R = 4, is the first determinant-possessing complex (i.e. has a vanishing Euler characteristics):

Ω(0, 2) Ω(2, 1) Ω(4, 0) → → with dimensions 3 18 15. If we select a basis in Ω(0, 2) as → →

θ θ ,θ θ ,θ θ { 2 3 1 3 1 2} a basis in Ω(2, 1) as

x2θ , x2θ , x2θ , x x θ , x x θ , x x θ , x x θ , x x θ , { 1 3 1 2 1 1 1 2 3 1 2 2 1 2 1 1 3 3 1 3 2

x x θ , x2θ , x2θ , x2θ , x x θ , x x θ , x x θ , x2θ , x2θ , x2θ 1 3 1 2 3 2 2 2 1 2 3 3 2 3 2 2 3 1 3 3 3 2 3 1} and, finally, a basis in Ω(4, 0) as

x4, x3x , x3x , x2x2, x2x x , x2x2, x x3, x x2x , x x x2, x x3, x4, x3x , x2x2, x x3, x4 { 1 1 2 1 3 1 2 1 2 3 1 3 1 2 1 2 3 1 2 3 1 3 2 2 3 2 3 2 3 3} then Koszul differential is represented by a pair of matrices: one 3 18 ×

11 and another 18 15 ×

By selecting some 3 columns in dˆ1 and complementary 15 rows in dˆ2, we obtain the desired resultant

This particular formula corresponds to the choice of columns 1, 2 and 5 in the first matrix, but the resultant, of course, does not depend on this choice: any other three columns will do. One only needs to care that determinant in the denominator does not vanish. If it is non-zero, then the upper minor is divisible by the lower minor, and the fraction is equal (up to sign) to the resultant.

12 Alternative complex is the second complex with vanishing χ, that corresponds to R = 5:

Ω(1, 2) Ω(3, 1) Ω(5, 0) → → It has dimensions 9 30 21. If we select a basis in Ω(1, 2) as → → x θ θ , x θ θ , x θ θ , x θ θ , x θ θ , x θ θ , x θ θ , x θ θ , x θ θ { 1 2 3 1 1 3 1 1 2 2 2 3 2 1 3 2 1 2 3 2 3 3 1 3 3 1 2} a basis in Ω(3, 1) as

x3θ , x3θ , x3θ , x2x θ , x2x θ , x2x θ , x2x θ , x2x θ , x2x θ , { 1 3 1 2 1 1 1 2 3 1 2 2 1 2 1 1 3 3 1 3 2 1 3 1 2 2 2 2 2 2 3 3 3 x1x2θ3, x1x2θ2, x1x2θ1, x1x2x3θ3, x1x2x3θ2, x1x2x3θ1, x1x3θ3, x1x3θ2, x1x3θ1, x2θ3, x2θ2, x2θ1,

2 2 2 2 2 2 3 3 3 x2x3θ3, x2x3θ2, x2x3θ1, x2x3θ3, x2x3θ2, x2x3θ1, x3θ3, x3θ2, x3θ1 x x θ , x2θ , x2θ , x2θ , x x θ , x x θ , x x θ , x2θ , x2θ , x2θ 1 3 1 2 3 2 2 2 1 2 3 3 2 3 2 2 3 1 3 3 3 2 3 1} and, finally, a basis in Ω(5, 0) as

x5, x4x , x4x , x3x2, x3x x , x3x2, x2x3, x2x2x , x2x x2, x2x3, { 1 1 2 1 3 1 2 1 2 3 1 3 1 2 1 2 3 1 2 3 1 3 x x4, x x3x , x x2x2, x x x3, x x4, x5, x4x , x3x2, x2x3, x x4, x5 1 2 1 2 3 1 2 3 1 2 3 1 3 2 2 3 2 3 2 3 2 3 3} then Koszul differential is represented by a pair of matrices: one 9 30 ×

and another 30 21 ×

13 By selecting some 9 columns in dˆ1 and complementary 21 rows in dˆ2, we obtain the desired resultant

This particular formula corresponds to the choice of columns 1, 3, 5, 7, 12, 14, 16, 19 and 21 in the first matrix. Of course, any other choice will give the the same (up to a sign) result, if only the minor in denominator is non-zero. Direct calculation (division of one minor by another) should demonstrate, that the two complexes 3 18 15 and 9 30 21 give one and the same expression for the resultant. It should also be reproduced → → → → by determinants of all other complexes from the 3 2 list with R> 3. | 3.1.3 Macaulay’s formula

The second, rather combinatorial, generalization of the Sylvester method is due to Macaulay. This method allows to express the resultant as a ratio of just two determinants (for Koszul complexes (14) both the numerator and denominator are products of several determinants). For this reason, Macaulay formula is computationally more economical, than eq. (14). We do not go into details here, they can be found in [10].

14 3.2 Formulas of Bezout type

Consider a system of n polynomial equations, homogeneous of degree r in variables x1,...,xn:

f1(x1,...,xn)=0   f2(x1,...,xn)=0     . . .   fn(x1,...,xn)=0    where

a1 a2 an fi(x1,...,xn)= fi|a1a2...an x1 x2 ...xn a +a +...+a =r 1 2 n

In the monomial basis, coefficients of this system can be viewed as forming an n Cr rectangular matrix. × n+r−1 The set of all top-dimensional minors of this matrix is naturally a tensor with n indices:

Mi ...i = n n minor of f, situated in coloumns i ,...,in 1 n × 1

n−1 Resultant R f ,...,fn is expected to be a homogeneous polynomial of variables Mi ...i of degree r . Such { 1 } 1 n expressions are called formulas of Bezout type. The simplest example of Bezout formulas is the case n =2, r = 1, where the system of equations is just a simple linear system:

f1x1 + f2x2 =0   g1x1 + g2x2 =0  and the only top-dimensional minor is

M = f g f g 12 1 2 − 2 1

In this case, resultant is just equal to this minor:

R f,g = M 2|1{ } 12

In the next-to-simplest case n =2, r = 2 the system of equations has a form

2 2 f11x1 + f12x1x2 + f22x2 =0

 2 2  g11x1 + g12x1x2 + g22x2 =0  and there are three top-dimensional minors:

15 M = f g f g 12 11 12 − 12 11 M = f g f g 13 11 22 − 22 11 M = f g f g 23 12 22 − 22 12

In this case, resultant is a homogeneous polynomial of degree 2 in minors , which is moreover a determinant

M12 M13 R f,g = M M M 2 = det 2|2{ } 12 23 − 13   M13 M23   For n =2, r = 3 the system of equations has a form

3 2 2 3 f111x1 + f112x1x2 + f122x1x2 + f222x2 =0

 3 2 2 3  g111x1 + g112x1x2 + g122x1x2 + g222x2 =0  and there are six top-dimensional minors:

M = f g f g M = f g f g 12 111 112 − 112 111 13 111 122 − 122 111 M = f g f g M = f g f g 14 111 222 − 222 111 23 112 122 − 122 112 M = f g f g M = f g f g 24 112 222 − 222 112 34 122 222 − 222 122

In this case, resultant is a homogeneous polynomial of degree 3 in minors, which is again a determinant

R f,g = M M M + M M M M M 2 M 2 M +2M M M M 3 M 2 M (15) 2|3{ } 12 34 14 12 34 23 − 12 24 − 13 34 13 14 24 − 14 − 14 23

M12 M13 M14

 M M + M M  R f,g = det 13 23 14 23 (16) 2|3{ }      M M M   14 24 34       

Note, that six variables Mij are not independent: they are subject to one constraint

M M M M + M M = 0 (17) 12 34 − 13 24 14 23 which is well-known as Plucker relation. For this reason, the polynomial expression is actually not unique: there are many different polynomial expressions for resultant through minors, whose difference is proportional to the Plucker relation. Generally (for higher n and r) there are several Plucker relations, not a single one. Variables

16 Mi1...in are sometimes called Plucker coordinates. Bezout formulas, which express resultant through Plucker coordinates, exist in higher dimensions: for example, for n =3, r = 3 we have

2 2 2 f(x1, x2, x3)= f11x1 + f12x1x2 + f13x1x3 + f22x2 + f23x2x3 + f33x3 =0 2 2 2  g(x1, x2, x3)= g11x1 + g12x1x2 + g13x1x3 + g22x2 + g23x2x3 + g33x3 =0   2 2 2 h(x1, x2, x3)= h11x1 + h12x1x2 + h13x1x3 + h22x2 + h23x2x3 + h33x3 =0   and there are 20 top-dimensional minors:

M = f g h f g h g f h + g f h + h f g h f g 123 11 12 13 − 11 13 12 − 11 12 13 11 13 12 11 12 13 − 11 13 12 M = f g h f g h g f h + g f h + h f g h f g 124 11 12 22 − 11 22 12 − 11 12 22 11 22 12 11 12 22 − 11 22 12 M = f g h f g h g f h + g f h + h f g h f g 125 11 12 23 − 11 23 12 − 11 12 23 11 23 12 11 12 23 − 11 23 12 M = f g h f g h g f h + g f h + h f g h f g 126 11 12 33 − 11 33 12 − 11 12 33 11 33 12 11 12 33 − 11 33 12 M = f g h f g h g f h + g f h + h f g h f g 134 11 13 22 − 11 22 13 − 11 13 22 11 22 13 11 13 22 − 11 22 13 M = f g h f g h g f h + g f h + h f g h f g 135 11 13 23 − 11 23 13 − 11 13 23 11 23 13 11 13 23 − 11 23 13 M = f g h f g h g f h + g f h + h f g h f g 136 11 13 33 − 11 33 13 − 11 13 33 11 33 13 11 13 33 − 11 33 13 M = f g h f g h g f h + g f h + h f g h f g 145 11 22 23 − 11 23 22 − 11 22 23 11 23 22 11 22 23 − 11 23 22 M = f g h f g h g f h + g f h + h f g h f g 146 11 22 33 − 11 33 22 − 11 22 33 11 33 22 11 22 33 − 11 33 22 M = f g h f g h g f h + g f h + h f g h f g 156 11 23 33 − 11 33 23 − 11 23 33 11 33 23 11 23 33 − 11 33 23 M = f g h f g h g f h + g f h + h f g h f g 234 12 13 22 − 12 22 13 − 12 13 22 12 22 13 12 13 22 − 12 22 13 M = f g h f g h g f h + g f h + h f g h f g 235 12 13 23 − 12 23 13 − 12 13 23 12 23 13 12 13 23 − 12 23 13 M = f g h f g h g f h + g f h + h f g h f g 236 12 13 33 − 12 33 13 − 12 13 33 12 33 13 12 13 33 − 12 33 13 M = f g h f g h g f h + g f h + h f g h f g 245 12 22 23 − 12 23 22 − 12 22 23 12 23 22 12 22 23 − 12 23 22 M = f g h f g h g f h + g f h + h f g h f g 246 12 22 33 − 12 33 22 − 12 22 33 12 33 22 12 22 33 − 12 33 22 M = f g h f g h g f h + g f h + h f g h f g 256 12 23 33 − 12 33 23 − 12 23 33 12 33 23 12 23 33 − 12 33 23 M = f g h f g h g f h + g f h + h f g h f g 345 13 22 23 − 13 23 22 − 13 22 23 13 23 22 13 22 23 − 13 23 22 M = f g h f g h g f h + g f h + h f g h f g 346 13 22 33 − 13 33 22 − 13 22 33 13 33 22 13 22 33 − 13 33 22 M = f g h f g h g f h + g f h + h f g h f g 356 13 23 33 − 13 33 23 − 13 23 33 13 33 23 13 23 33 − 13 33 23 M = f g h f g h g f h + g f h + h f g h f g 456 22 23 33 − 22 33 23 − 22 23 33 22 33 23 22 23 33 − 22 33 23

This time there is again a single Plucker relation

17 M M M M + M M M M + M M 123 456 − 124 356 125 346 − 126 345 134 256

M M + M M + M M M M + M M = 0 (18) − 135 246 136 245 145 236 − 146 235 156 234

In this respect, the case n = 3, r = 2 is analogous to the case n = 2, r = 4. Be it coincidence or not, this is not the only parallel between these two cases: see the chapter 5 about integral discriminants. Resultant is a homogeneous polynomial of degree 4 in minors:

R f,g,h = M M M M M M M M M M M M + M M M 2 3|2{ } 123 145 156 456 − 123 145 236 456 − 123 145 345 356 123 145 346 − M M 2 M + M M M M + M M M M 2M M M M M M M M + 123 146 456 123 146 235 456 123 146 245 356 − 123 146 246 346 − 123 156 234 456 M M M 2 + M M M M + M M M M M M M 2 M M M M + 123 156 246 123 234 246 356 123 234 345 356 − 123 234 346 − 123 235 245 356 M M M M + M M M M M M M 2 M M M M M M M M + 123 235 246 346 123 236 245 346 − 123 236 246 − 123 236 246 345 − 124 126 156 456 M M M M + M M M M M M M 2 + M M M M M M M M + 124 126 246 356 124 126 345 356 − 124 126 346 124 135 236 456 − 124 135 246 356 M M M M M M M M + M M M M M M 2 M + M M M M + 124 136 146 456 − 124 136 235 456 124 136 246 346 − 124 146 356 124 146 156 346 M M M M M M M M M M 2 M M M M M + M M M M + 124 146 235 356 − 124 146 236 346 − 124 156 246 − 124 156 235 346 124 156 236 246 M M M M +M M M M M M M M +M M M M +M M M M 124 156 236 345 125 126 146 456− 125 126 245 356 125 126 246 346 125 134 156 456− M M M M M M M M + M M M 2 M M M M + M M M M 125 134 236 456 − 125 134 345 356 125 134 346 − 125 135 146 456 125 135 245 356 − M M M M + M M M M M M M M + M M M M + M M 2 M 125 135 246 346 125 136 234 456 − 125 136 245 346 125 136 246 345 125 146 346 − M M M M M M M M M M M M + M M M M + M M 2 M + 125 146 156 246 − 125 146 156 345 − 125 146 234 356 125 146 236 246 125 156 245 M M M M M M M M M 2 M M + M 2 M M M 2 M 2 M 2 M M 125 156 234 346 − 125 156 236 245 − 126 145 456 126 245 346 − 126 246 − 126 246 345 − M M M M +M M M M M M M M +M M M M M M M M + 126 134 146 456 126 134 235 456− 126 134 246 346 126 135 145 456− 126 135 234 456 M M M 2 + M M 2 M 3M M M M + M M M M + M M M M + 126 135 246 126 145 356 − 126 145 146 346 126 145 156 246 126 145 156 345 2M M 2 M + M M 2 M 2M M M M + M M M M M M M M 126 146 246 126 146 345 − 126 146 156 245 126 146 234 346 − 126 146 235 246 − M M M M M M M M + M M M M M 2 M M + M 2 M M + 126 156 234 246 − 126 156 234 345 126 156 235 245 − 134 156 456 134 246 356 M 2 M M M 2 M 2 + M M M M M M M M + M M M M 134 345 356 − 134 346 134 135 146 456 − 134 135 245 356 134 135 246 346 − M M M M + M M M M M M M 2 M M M M +2M M M M 134 136 145 456 134 136 245 346 − 134 136 246 − 134 136 246 345 134 145 146 356 − M M M M M M M M + M M M M 2M M 2 M +3M M M M + 134 145 156 346 − 134 145 235 356 134 145 236 346 − 134 146 346 134 146 156 246 M M M M + M M M M M M M M M M M M M M 2 M 134 146 156 345 134 146 235 346 − 134 146 236 246 − 134 146 236 345 − 134 156 245 − M M 2 M + M M M M + M M M M M M M M M M 2 M 135 145 356 135 145 146 346 135 145 234 356 − 135 145 236 246 − 135 146 246 − M M M M + M M M M + M M 2 M M M M M M M M M 135 146 234 346 135 146 236 245 136 145 346 − 136 145 146 246 − 136 145 146 345 − M M M M + M M M M + M M 2 M + M M M M + M M M M 136 145 234 346 136 145 235 246 136 146 245 136 146 234 246 136 146 234 345 − M M M M M 2 M 2 +M 2 M M +2M M 2 M M M 2 M M M M M + 136 146 235 245 − 145 156 145 156 236 145 146 156 − 145 146 236 − 145 146 156 235 M M 2 M M 4 + M 3 M M 2 M M + M M M M M M M M 145 156 234 − 146 146 235 − 146 156 234 256 123 146 345 − 256 123 156 245 − M M M M + M M M M M M M M + M M M M 256 123 234 346 256 123 236 245 − 256 124 126 346 256 124 135 346 − M M M M + M M M M M M M M + M M 2 M + M M M M 256 124 136 345 256 124 146 156 − 256 124 146 236 256 126 245 256 126 134 345 − M M M M M M M M + M M M M M M 2 M + M M M M 256 126 135 245 − 256 126 145 146 256 126 146 234 − 256 134 346 256 134 136 245 − M M M M + M M M M M M M 2 + M M M M M M M M 256 134 145 156 256 134 145 236 − 256 134 146 256 135 145 146 − 256 136 145 234 This time it is not a determinant of some matrix (or, at least, such matrix is not yet found). Instead, it is a Pfaffian of the following 8 8 antisymmetric matrix, made of the Plucker coordinates: ×

18 0 M356 M456 M246 M156 M146 M256 M346   −M356 0 −M346 M146 M136 M126 M236 M156 − M236      −M456 M346 0 M245 M146 M145 M246 M345         −M246 −M146 −M245 0 M134 M124 M234 − M145 −M234    A =   (19)    −M156 −M136 −M146 −M134 0 M123 −M126 M126 − M135         −M146 −M126 −M145 −M124 −M123 0 M134 − M125 −M134         −M256 −M236 −M246 −M234 + M145 M126 −M134 + M125 0 M146 − M235         −M346 −M156 + M236 −M345 M234 −M126 + M135 M134 −M146 + M235 0      The Pfaffian is equal to

R f,g,h = pfaffA = ǫi ...i Ai i Ai i Ai i Ai i (20) 3|2{ } 1 8 1 2 3 4 5 6 7 8

where ǫ is the completely antisymmetric tensor (simply the antisymmetrisation over i1,...,i8). The Pfaffian is also the same, as the of determinant, since for antisymmetric matrices determinant is always an exact square. Observation (20) was done in [13], where the variables Mijk are being interpreted as coordinates on a Grassmanian manifold and then, using advanced homological algebra, this and similar formulas are being obtained. Generally speaking, resultant is expected to be a polynomial of degree rn−1 in the Plucker coordinates

Mi1...in . It is an open problem to find an explicit formula for this polynomial.

3.3 Formulas of Sylvester + Bezout type

By combining the techniques of Sylvester and Bezout, it is possible to sufficiently enlarge the number of cases, where resultant is expressed as some polynomial (usually determinant) without any divisions. The basic example here is n =3, r = 2. In this case, the system of quadratic equations has a form

2 2 2 f(x1, x2, x3)= f11x1 + f12x1x2 + f13x1x3 + f22x2 + f23x2x3 + f33x3 =0 2 2 2  g(x1, x2, x3)= g11x1 + g12x1x2 + g13x1x3 + g22x2 + g23x2x3 + g33x3 =0 (21)   2 2 2 h(x1, x2, x3)= h11x1 + h12x1x2 + h13x1x3 + h22x2 + h23x2x3 + h33x3 =0   It can be represented as a 3 6 rectangular matrix. Obviously, another 3 6 matrix is necessary to complete × × a square 6 6 matrix. This complementary matrix is constructed quite elegantly, making use of Jacobian × ∂f J(x , x , x ) = det i , f f,f g,f h 1 2 3 ∂x 1 ≡ 2 ≡ 3 ≡ j which is a homogeneous polynomial of degree 3 in x1, x2, x3 and at the same time of degree 3 in coefficients of f,g and h. Jacobian can be expressed through Plucker variables Mijk as follows:

19 J(x x x )= M x3 + ( 2M + M )x2x + (2M M )x2x + ( M +2M )x x2 + ( M + 1 2 3 123 1 − 134 125 1 2 126 − 135 1 3 − 234 145 1 2 − 235 4M )x x x + ( M +2M )x x2 + M x3 + (2M + M )x2x + (M +2M )x x2 + M x3 146 1 2 3 − 236 156 1 3 245 2 246 345 2 3 256 346 2 3 356 3

The derivatives of J are homogeneous quadratic polynomials in x1, x2, x3, just like the equations f,g and h:

f11 f12 f13 f22 f23 f33 f 2 x1    g  g11 g12 g13 g22 g23 g33   x1x2            h   h11 h12 h13 h22 h23 h33         x1x3        ∂J =       3M123 −4M134 + 2M125 −2M135 + 4M126 −M234 + 2M145 −M235 + 4M146 −M236 + 2M156      2   ∂x1     x2               ∂J   −2M134 + M125 4M145 − 2M234 −M235 + 4M146 3M245 2M345 + 4M246 M256 + 2M346       x2x3   ∂x2              2M126 − M135 −M235 + 4M146 −2M236 + 4M156 M345 + 2M246 2M256 + 4M346 3M356  ∂J     2       x3  ∂x3               The matrix in the right hand side is half Sylvester, half Bezout. Its determinant is equal to the resultant:

f11 f12 f13 f22 f23 f33 g11 g12 g13 g22 g23 g33 h11 h12 h13 h22 h23 h33 R3|2 f,g,h = (22) { } 3M123 −4M134 + 2M125 −2M135 + 4M126 −M234 + 2M145 −M235 + 4M146 −M236 + 2M156 −2M134 + M125 4M145 − 2M234 −M235 + 4M146 3M245 2M345 + 4M246 M256 + 2M346 2M126 − M135 −M235 + 4M146 −2M236 + 4M156 M345 + 2M246 2M256 + 4M346 3M356 In a compact notation this formula can be written as

∂J ∂J ∂J R3|2 f,g,h = det f,g,h, , , (23) { } 6×6 ∂x ∂x ∂x 1 2 3 where det . . . stands for the determinant of the matrix, obtained by expanding the polynomials inside brackets { } in the monomial basis. The other possibility to design a is to take 9 polynomials xifj, which are homogeneous polynomials of degree 3, and complement them with the Jacobian itself, not its derivatives. In this case one gets a 10 10 matrix in the monomial basis: ×

R3|2 f,g,h = det x1f, x1g, x1h, x2f, x2g, x2h, x3f, x3g, x3h, J (24) { } 10×10 { }

According to this method, it is necessary to make the matrix square by adding a number of certain polynomials

20 to the Sylvester block. Of course, the most non-trivial part of the method is construction of these polynomials. For higher r> 2, these polynomials are constructed with the help of

2 3 ∂fi 1 ∂ fi 1 ∂ fi (x1,...,xn,p1,...,pn) = det + pk + pkpl + . . . J ∂xj r 1 ∂xj ∂xk (r 1)(r 2) ∂xj ∂xk∂xl − − −

Jacobian-like quantity can be viewed as a series in the additional variables p ,...,pn: J 1 1 (x ,...,xn,p ,...,pn)= J(x ,...,xn)+ Ji(x ,...,xn) + Jij (x ,...,xn)pipj + . . . J 1 1 1 1 2! 1

Note, that J(x1,...,xn) is nothing but the ordinary Jacobian and Ji(x1,...,xn) are its derivatives, while quantities Jij (x1,...,xn) are already new and non-trivial (in particular, they are not equal to the second derivatives of J). In the next case n =3, r = 3 the Sylvester-Bezout formulas take form

R3|3 f1,f2,f3 = det xixj fk, Jl (25) { } 21×21 { } and

R3|3 f1,f2,f3 = det xifj , Jkl (26) { } 15×15 { }

Just like in the previous case, there exist two different formulas. In the next case n =3, r = 4 we have

R3|4 f1,f2,f3 = det xixj xkfl, Jpq (27) { } 36×36 { } and

R3|4 f1,f2,f3 = det xixj fk, Jlpq (28) { } 28×28 { }

Generally, for n = 3 and r = any, we have two formulas of Sylvester-Bezout type:

R3|r f1,f2,f3 = det xi1 ...xir−1 fj , Jk1...kr−2 { } r(2r+1)×r(2r+1) (29) and

R3|r f1,f2,f3 = det xi1 ...xir−2 fj , Jk1...kr−1 { } r(2r−1)×r(2r−1) (30) As for higher dimensions n> 3, an explicit formula is known only for n =4, r = 2 where it has a form

R4|2 f1,f2,f3,f4 = det xifj , Jk (31) { } 20×20 { }

For higher n and r, such formulas are still an open direction of research, see [11] for more details.

21 3.4 Formulas of Schur type

Resultants also admit a very different kind of formulas, related to certain analytic structure of the of resultant, described in [12]. By an analytic structure we mean that logarithm of resultant of a system

f1(x1,...,xn)=0   f2(x1,...,xn)=0     . . .   fn(x1,...,xn)=0    depends on polynomials f1,...,fn in a very specific way, which is much simpler to understand than for the resultant itself. Two formulas are used to express this statement. The first is the operator formula

∞ ∞ r1 ˆ k rn ˆ k log Rr1,...,rn (I f)= (f1) . . . (fn) log det(I A) (32) − (r1k)! (rnk)! − k=0 k=0 A=0 where A is an auxillary n n matrix, and fˆ , fˆ ,..., fˆn are differential operators × 1 2

∂ ∂ ∂ fˆ = f , ,..., 1 1 ∂A ∂A ∂A 11 12 1n ∂ ∂ ∂ fˆ = f , ,..., 2 2 ∂A ∂A ∂A 21 22 2n (33) . . . ∂ ∂ ∂ fˆ = f , ,..., n n ∂A ∂A ∂A n1 n2 nn associated with polynomials f1,f2,...,fn. The second is the contour integral formula:

log R f1,f2,...,fn = . . . log f1 d log f2 . . . d log fn + r1 log R f2,...,fn (34) ∧ ∧ x1=0 where the contour integral is (n 1)-fold and functions fi are taken with arguments fi(1,z ,...,zn ). The − 1 −1 contour of integration is assumed to encircle all the common roots of the system

fk(1,z1,...,zn−1)=0, k =2,...,n (35) see [12] for more details. An explicit formula for log R follows after several iterations of (34): for n =2

log Rr r f ,f = log f (1,z)d log f (1,z)+ r log f (0, 1) (36) 1 2 { 1 2} 1 2 1 2

22 for n =3

log Rr r r f ,f ,f = log f (1,z ,z )d log f (1,z ,z ) d log f (1,z ,z )+ 1 2 3 { 1 2 3} 1 1 2 2 1 2 ∧ 3 1 2 r1 log f2(0, 1,z)d log f3(0, 1,z)+ r1r2 log f3(0, 0, 1) and so on. In this way, logarithm of the resultant is expressed through the simple contour integrals. To obtain R from log R, we exponentiate the latter: for example, for n =2

Rr r f ,f = exp log f (1,z)d log f (1,z)+ r log f (0, 1) (37) 1 2 { 1 2} 1 2 1 2

Using this formula, it is straightforward to express the resultant of two polynomials through their roots: if

r1 f (1,z)= a (z αi) (38) 1 − i=1

r2 f (1,z)= b (z βi) (39) 2 − i=1 then

r1 r2 r1 r2 r2 r1 Rr r f ,f = exp log(z αi)d log(z βj )+ r log a + r log b = a b (βj αi) (40) 1 2 { 1 2}  − − 2 1  − i=1 j=1 i=1 j=1   Exponential formula (37) and its direct analogues for n > 2 have several advantages. The most important are simplicity and explicitness: such formulas have a transparent structure, which is easy to understand and memorize. Another one is universality: they can be written for any n, not just for low values. However, there is one serious drawback as well. Resultant is known to be a homogeneous polynomial of degree

r1r2 ...rn di = ri in coefficients of the i-th equation fi, and of total degree

1 1 1 d = d + . . . + d = r r ...r + + . . . + 1 n 1 2 n r r r 1 2 n in coefficients of all equations, but in exponential formula the polynomial nature of the resultant is not explicit. Therefore, it is not fully convenient for practical calculations of resultants (although it is conceptually adequate and can be used in theoretical considerations).

23 To turn these formulas into a computational tool, one can apply (34) to a shifted system

f˜ (x) = (x )r1 λ f (x) 1 1 − 1 1  f˜ (x) = (x )r2 λ f (x) 2 2 − 2 2  (41)  . . .  r f˜n(x) = (xn) n λnfn(x)  −   and then expand the shifted in the integrands into Taylor series in powers of the ”spectral parameters”

λi. Then, one obtains the following series expansion

∞ ∞ k1 k2 kn log R f˜ ,..., f˜n = . . . Tk k ...k f ,...,fn λ λ ...λ (42) 1 − 1 2 n 1 1 2 n k1=0 k =0 n

(a minus sign is just a convention) where particular Taylor components Tk1k2...kn are homogeneous polynomial expressions of degree ki in coefficients of fi. We call them traces of a non-linear system f1(x),...,fn(x). By exponentiating, one obtains the shifted resultant:

∞ ∞ k1 k2 kn R f˜ ,..., f˜n = exp . . . Tk k ...k λ λ ...λ 1 − 1 2 n 1 2 n k1=0 k =0 n

d d1 d2 d The original resultant R f ,...,fn is equal to ( 1) times the coefficient of λ λ ...λ n in R f˜ ,..., f˜n . 1 − 1 2 n 1 We can extract it, expanding the right hand side in powers of λi:

∞ ∞ ∞ ∞ k1 k2 kn k1 k2 kn exp . . . Tk k ...k λ λ ...λ = . . . k k ...k λ λ ...λ − 1 2 n 1 2 n P 1 2 n 1 2 n k k k k 1=0 n=0 1=0 n=0

Such power series expansion of exp (S(x)), where S(x) is itself a power series (perhaps, of many variables), is often called a Schur expansion. It implies the following relation between Taylor components:

k +...+k 1 n ( 1)m k k ...k = − Tv Tv ...Tv P 1 2 n m! 1 2 m m=1 v1+v2+...+vm=k

where the sum is taken over all ordered partitions of a a vector k = (k1, k2,...,kn) into m vectors, denoted as v1,...,vm. An ordered partition is a way of writing a vector with integer components as a sum of vectors with integer components, where the of the items is significant. Polynomials k k ...k (T ) are often called P 1 2 n multi-Schur polynomials. Several first multi-Schur polynomials are given by

24 , = T , P1 0 − 1 0 2 , = T , + T /2 P2 0 − 2 0 1,0 2 , = T , + T , T , + T , T , T T , /2 (43) P2 1 − 2 1 2 0 0 1 1 0 1 1 − 1,0 0 1 2 , , = T , , + T , , T , , + T , , T , , T T , , /2 P2 1 0 − 2 1 0 2 0 0 0 1 0 1 0 0 1 1 0 − 1,0,0 0 1 0

, , = T , , + T , , T , , + T , , T , , + T , , T , , T , , T , , T , , P1 1 1 − 1 1 1 1 0 0 0 1 1 0 1 0 1 0 1 1 1 0 0 0 1 − 0 1 0 0 0 1 1 0 0

Resultant is a certain multi- of traces:

d m+d d ( 1) R f ,...,fn = ( 1) d ,...,d = − Tv Tv ...Tv (44) 1 − P 1 n m! 1 2 m m=1 v1+v2+...+vm=d

The explicit formula for traces, obtained in [12], has a form

n 1 ki Tk = Tk1k2...kn = det δij riki rij (fi)ri1,ri2,...,rin (45) k1k2 ...kn 2≤i,j≤n − rij i=1 where

k j1 j2 jn k (f)j1,j2,...,jn = coefficient of x1 x2 ...xn in f(x1, x2,...,xn)

and the sum goes over all non-negative integer n n matrices rij such that the sum of entries in any i-th row × or i-th coloumn is fixed and equals riki. Formula (45) is valid for positive k1,...,kn. The cases when some ki = 0, also make no difficulty. If some ki = 0, then

Tk1,k2,...,kn f1,f2,...,fn = ri Tk1,k2,...,ki 1,ki+1,...,kn f1,f2,...,fi−1,fi+1,...,fn (46) { } − { } xi=0 where the trace in the right hand side is taken in variables x1,...,xi−1, xi+1,...,xn. It is an easy exercise to prove this statement, making a series expansion in (34). Similarly, operator approach gives [12]

k1 kn r1k1+...+rnkn r1(fˆ1) rn(fˆn) trA Tk1,k2,...,kn f1,f2,...,fn = . . . (47) { } (r1k1)! (rnkn)! r1k1 + . . . + rnkn A=0 Together, formulas 44 and 45 (or, alternatively, 47) give an explicit to calculate resultants without any polynomial divisions. The absence of polynomial divisions should increase computational speed of the algorithm, if compared to with division such as Koszul complex.

25 To illustrate these general formulas, let us consider a particular calculation for n = 3 and r = 2, where the system of equations again has the form 21. The traces are calculated with (47), using differential operators

∂ 2 ∂2 ∂2 ∂ 2 ∂2 ∂ 2 fˆ = f + f + f + f + f + f 11 ∂A 12 ∂A ∂A 13 ∂A ∂A 22 ∂A 23 ∂A ∂A 33 ∂A 11 11 12 11 13 12 12 13 13 ∂ 2 ∂2 ∂2 ∂ 2 ∂2 ∂ 2 gˆ = g + g + g + g + g + g 11 ∂A 12 ∂A ∂A 13 ∂A ∂A 22 ∂A 23 ∂A ∂A 33 ∂A 21 21 22 21 23 22 22 23 23 ∂ 2 ∂2 ∂2 ∂ 2 ∂2 ∂ 2 hˆ = h + h + h + h + h + h 11 ∂A 12 ∂A ∂A 13 ∂A ∂A 22 ∂A 23 ∂A ∂A 33 ∂A 31 31 32 31 33 32 32 33 33 and a few first traces are

T1 =4f11 +4g22 +4h33

2 2 2 T2 =4f11 +4g12f12 +4h13f13 +8g11f22 +8h11f33 +4g22 +8g33h22 +4g23h23 +4h33

3 2 2 T3 =4f11 +6g12f11f12 +6h13f11f13 + 12g11f11f22 + 12h11f11f33 +6g11f12 +6h11f13 +3g23h13f12 + 2 6g33h12f12 +6g12g22f12 +3g13h23f12 +3g23h12f13 +6h13h33f13 +6g13h22f13 +3g12h23f13 +6g12f22 + 2 12g11g22f22 +6g13h13f22 + 12g33h11f22 +6g23h11f23 +3g12h13f23 +9g13h12f23 +6g11h23f23 +6h13f33 + 3 6g12h12f33 + 12g11h22f33 + 12h11h33f33 + 12g22g33h22 +4g22 +6g23h23h33 +6g22g23h23 + 2 2 3 6g23h22 +6g33h23 + 12g33h22h33 +4h33 and so on. The resultant has degree 12, and is expressed through traces as

Tk(f) R f,g,h = P − = 3|2 12 k 1 1 1 1 1 1 1 1 1 1 T + T 2 + T T + T T + T T + T T + T T T 3 T T T T 2T −12 12 72 6 35 5 7 32 4 8 27 3 9 20 2 10 11 1 11 − 384 4 − 60 3 4 5 − 108 3 6 − 1 1 1 1 1 1 1 1 1 T T 2 T T T T T T T 2T T T T T T T T T T T T T T 2T + 100 2 5 − 48 2 4 6 − 42 2 3 7 − 64 2 8 − 30 1 5 6 − 28 1 4 7 − 24 1 3 8 − 18 1 2 9 − 20 1 10 1 1 1 1 1 1 1 1 T 4 + T T 2T + T 2T 2 + T 2T T + T 3T + T T T 2 + T T 2T + T T T T + 1944 3 144 2 3 4 256 2 4 120 2 3 5 288 2 6 96 1 3 4 90 1 3 5 40 1 2 4 5 1 1 1 1 1 1 1 1 T T T T + T T 2T + T 2T 2 + T 2T T + T 2T T + T 2T T + T 3T T 3T 2 36 1 2 3 6 56 1 2 7 100 1 5 48 1 4 6 42 1 3 7 32 1 2 8 54 1 9 − 864 2 3 − 1 1 1 1 1 1 1 T 4T T T T 3 T T 2T T T T 3T T 2T 2T T 2T T 2 T 2T T T 1536 2 4 − 324 1 2 3 − 96 1 2 3 4 − 240 1 2 5 − 144 1 3 4 − 128 1 2 4 − 60 1 2 3 5 − 1 1 1 1 1 1 1 1 T 2T 2T T 3T T T 3T T T 3T T T 4T + T 6 + T T 4T + T 2T 2T 2 + 96 1 2 6 − 120 1 4 5 − 108 1 3 6 − 84 1 2 7 − 192 1 8 46080 2 1152 1 2 3 288 1 2 3 1 1 1 1 1 1 1 1 T 2T 3T + T 3T 3 + T 3T T T + T 3T 2T + T 4T 2 + T 4T T + T 4T T + T 5T 384 1 2 4 972 1 3 144 1 2 3 4 240 1 2 5 768 1 4 360 1 3 5 288 1 2 6 840 1 7 − 1 1 1 1 1 1 1 T 2T 5 T 3T 3T T 4T T 2 T 4T 2T T 5T T T 5T T T 6T + 7680 1 2 − 864 1 2 3 − 864 1 2 3 − 768 1 2 4 − 1440 1 3 4 − 1200 1 2 5 − 4320 1 6 1 1 1 1 1 1 1 T 4T 4 + T 5T 2T + T 6T 2 + T 6T T + T 7T T 6T 3 T 7T T 9216 1 2 2880 1 2 3 12960 1 3 5760 1 2 4 25200 1 5 − 34560 1 2 − 30240 1 2 3 − 1 1 1 1 1 T 8T + T 8T 2 + T 9T T 10T + T 12 161280 1 4 322560 1 2 1088640 1 3 − 7257600 1 2 479001600 1

This is a rather compact formula for R3|2. By substituting the expressions for traces, we obtain a polynomial expression in f, g and h, which contains 21894 monomials and takes several dozens of A4 pages.

26 Figure 1: Determinant of matrix Sij , represented as a diagram of tensor contraction. Black 2-valent vertices represent tensor S, white n-valent vertices represent tensor ǫ.

4 Calculation of discriminant

Discriminant is a particular case of resultant, associated with gradient systems of equations. Since discriminant is a function of a single homogeneous polynomial (instead of n homogeneous polynomials in the resultant case) it is slightly simpler than resultant, has bigger symmetry and some specific methods of calculation. These discriminant-specific methods are the topic of this section.

4.1 Discriminant through invariants

Let S(x1,...,xn) be a homogeneous polynomial of degree r in n variables. Discriminant of S is a polynomial SL(n) invariant of S. As such, it should be expressible as a polynomial in some simpler basis invariants, given by elegant diagram technique suggested in [1]. The simplest example is of course the determinant, given by a single diagram at Fig.1. The simplest non-trivial (i.e. non-Gaussian) example is a 3-form in 2 variables:

3 2 2 2 S(x1, x2)= S111x1 +3S112x1x2 +3S122x1x2 + S222x2

By dimension counting, there is only one elementary invariant I4 in this case, given by a diagram at Fig. 2. The subscript ”4” stands for the degree of this invariant. In this paper we find it convenient to denote the elementary invariants of degree k as Ik. It is straightforward to write the for the diagram:

i1j1 i2j2 k1l1 k2l2 i3k3 j3l3 I4 = Si1i2i3 Sj1j2j3 Sk1k2k3 Sl1l2l3 ǫ ǫ ǫ ǫ ǫ ǫ

Evaluating this sum, one gets the following explicit formula for I4

I =2S2 S2 12S S S S +8S S3 +8S3 S 6S2 S2 4 111 222 − 111 112 122 222 111 122 112 222 − 112 122

which is nothing but the algebraic discriminant D2|3 of S:

D2|3 = I4 (48)

27 Figure 2: The degree 4 invariant I4 of a 3-form in 2 variables, represented as a diagram of tensor contraction. Black 3-valent vertices represent tensor S, white 2-valent vertices represent tensor ǫ.

The next-to-simplest example is a 4-form in 2 variables, which can be written as

4 3 2 2 3 4 S(x1, x2)= S1111x1 +4S1112x1x2 +6S1122x1x2 +4S1222x1x2 + S2222x2

By dimension counting, there are two elementary invariants in this case. They have relatively low degrees 2 and 3, denoted as I2 and I3 and given by diagrams at Fig. 3 and Fig. 4, respectively. Looking at the diagrams, it is straightforward to write algebraic expressions for I2, I3:

Figure 3: The degree 2 invariant I2 of a 4-form in 2 variables, represented as a diagram of tensor contraction. Black 4-valent vertices represent tensor S, white 2-valent vertices represent tensor ǫ.

i1j1 i2j2 i3j3 i4j4 I2 = Si1i2i3i4 Sj1j2j3j4 ǫ ǫ ǫ ǫ

i1j1 i2j2 i3k1 i4k2 j3k3 j4k4 I3 = Si1i2i3i4 Sj1j2j3j4 Sk1k2k3k4 ǫ ǫ ǫ ǫ ǫ ǫ

28 Figure 4: The degree 3 invariant I3 of a 4-form in 2 variables, represented as a diagram of tensor contraction. Black 4-valent vertices represent tensor S, white 2-valent vertices represent tensor ǫ.

Evaluating these sums, one gets the following explicit formulas for I2, I3:

I =2S S 8S S +6S2 2 1111 2222 − 1112 1222 1122

I =6S S S 6S S2 6S2 S + 12S S S 6S3 3 1111 1122 2222 − 1111 1222 − 1112 2222 1112 1122 1222 − 1122

The algebraic discriminant D2|4, just like any other SL(2)-invariant function of S, is a function of I2, I3:

D = I3 6I2 =8S3 S3 96S2 S S S2 144S2 S2 S2 + 432S2 S S2 S 2|4 2 − 3 1111 2222 − 1111 1112 1222 2222 − 1111 1122 2222 1111 1122 1222 2222

216S2 S4 + 432S S2 S S2 48S S2 S2 S 1440S S S2 S S − 1111 1222 1111 1112 1122 2222 − 1111 1112 1222 2222 − 1111 1112 1122 1222 2222 4 3 3 2 2 2 + 648S1111S1122S2222 + 864S1111S1112S1122S1222 + 864S1112S1122S1222S2222 + 288S1112S1122S1222 432S S3 S2 216S4 S2 512S3 S3 432S2 S3 S (49) − 1111 1122 1222 − 1112 2222 − 1112 1222 − 1112 1122 2222 Our third example is a 5-form in 2 variables, which can be written as

5 4 3 2 2 3 4 5 S(x1, x2)= S11111x1 +5S11112x1x2 + 10S11122x1x2 + 10S11222x1x2 +5S12222x1x2 + S22222x2 By dimension counting, there are three elementary invariants in this case. They have degrees 4, 8 and 12, denoted as I4, I8 and I12 and given by diagrams at Fig. 5, Fig. 6 and Fig. 7.

Looking at the diagrams, it is straightforward to write an expression for I4

i1j1 i2j2 i3j3 i4k4 i5k5 j4l4 j5l5 k1l1 k2l2 k3l3 I4 = Si1i2i3i4i5 Sj1j2j3j4j5 Sk1k2k3k4k5 Sl1l2l3l4l5 ǫ ǫ ǫ ǫ ǫ ǫ ǫ ǫ ǫ ǫ

and equally straightforward to write expressions for I8, I12. Evaluating the contraction, one gets a formula

29 Figure 5: The degree 4 invariant I4 of a 5-form in 2 variables, represented as a diagram of tensor contraction. Black 5-valent vertices represent tensor S, white 2-valent vertices represent tensor ǫ.

Figure 6: The degree 8 invariant I8 of a 5-form in 2 variables, represented as a diagram of tensor contraction. Black 5-valent vertices represent tensor S, white 2-valent vertices represent tensor ǫ.

I = 2S2 S2 20S S S S +8S S S S + 4 11111 22222 − 11111 11112 12222 22222 11111 11122 11222 22222 32S S S2 24S S2 S + 32S2 S S + 18S2 S2 11111 11122 12222 − 11111 11222 12222 11112 11222 22222 11112 12222 − 24S S2 S 152S S S S + 96S S3 + 96S3 S 64S2 S2 11112 11122 22222 − 11112 11122 11222 12222 11112 11222 11122 12222 − 11122 11222

and similar formulas for I8, I12 – they are quite lengthy and we do not present them here. The algebraic discriminant D2|5, just like any other SL(2)-invariant function of S, is a function of I4, I8, I12:

D = I2 64I (50) 2|5 4 − 8

Our final example is a 3-form in 3 variables, which can be written as

S(x1, x2, x3)= 3 2 2 3 2 2 3 2 2 S111x1 +3S112x1x2 +3S113x1x3 +S222x2 +3S122x1x2 +3S223x2x3 +S333x3 +3S133x1x3 +3S233x2x3 +6S123x1x2x3

By dimension counting, there are two elementary invariants in this case. They have degrees 4 and 6, denoted as I4, I6 and given by diagrams at Fig. 8, Fig. 9.

30 Figure 7: The degree 12 invariant I12 of a 5-form in 2 variables, represented as a diagram of tensor contraction. Black 5-valent vertices represent tensor S, white 2-valent vertices represent tensor ǫ.

Looking at the diagrams, it is straightforward to write the algebraic expressions for I4 and I6:

i1j1k1 i2j2l2 i3k3l3 l1k2j3 I4 = Si1i2i3 Sj1j2j3 Sk1k2k3 Sl1l2l3 ǫ ǫ ǫ ǫ

i1k1l1 i2j2s2 j1k2m1 l2m2k3 m3s3j3 l3i3s1 I6 = Si1i2i3 Sj1j2j3 Sk1k2k3 Sl1l2l3 Sm1m2m3 Ss1s2s3 ǫ ǫ ǫ ǫ ǫ ǫ

Evaluating these sums, one gets the following explicit formulas for I4, I6:

I =6S4 12S S2 S +6S2 S2 +6S S S S 12S S2 S 6S S S S + 4 123 − 122 123 133 122 133 113 123 133 222 − 113 123 223 − 113 122 133 223 18S S S S 6S S2 S +6S2 S2 6S2 S S 6S S2 S + 18S S S S 113 122 123 233 − 113 122 333 113 223 − 113 222 233 − 112 133 222 112 123 133 223 − 12S S2 S 6S S S S +6S S S S 6S S S S +6S S S S + 112 123 233 − 112 122 133 233 112 122 123 333 − 112 113 223 233 112 113 222 333 6S2 S2 6S2 S S 6S S S2 +6S S S S +6S S S S 6S S S S 112 233 − 112 223 333 − 111 133 223 111 133 222 233 111 123 223 233 − 111 123 222 333 − 2 6S111S122S233 +6S111S122S223S333

I = 48S6 144S S4 S + 144S2 S2 S2 48S3 S3 + 72S S3 S S 144S S4 S 6 123 − 122 123 133 122 123 133 − 122 133 113 123 133 222 − 113 123 223 − 72S S S S2 S + 72S S S2 S S + 216S S S3 S + 72S S2 S2 S 113 122 123 133 222 113 122 123 133 223 113 122 123 233 113 122 133 223 − 216S S2 S S S 72S S2 S2 S + 72S S3 S S + 18S2 S2 S2 113 122 123 133 233 − 113 122 123 333 113 122 133 333 113 133 222 − 72S2 S S S S + 144S2 S2 S2 72S2 S2 S S + 72S2 S S S2 113 123 133 222 223 113 123 223 − 113 123 222 233 113 122 133 223 − 36S2 S S S S 216S2 S S S S + 144S2 S S S S + 162S2 S2 S2 113 122 133 222 233 − 113 122 123 223 233 113 122 123 222 333 113 122 233 − 144S2 S2 S S 48S3 S3 + 72S3 S S S 24S3 S2 S 72S S2 S2 S + 113 122 223 333 − 113 223 113 222 223 233 − 113 222 333 − 112 123 133 222 216S S3 S S 144S S4 S + 72S S S3 S 216S S S S2 S + 112 123 133 223 − 112 123 233 112 122 133 222 − 112 122 123 133 223 72S S S2 S S + 72S S S3 S + 72S S2 S2 S 72S S2 S S S 112 122 123 133 233 112 122 123 333 112 122 133 233 − 112 122 123 133 333 − 36S S S2 S S 216S S S S S2 +360S S S S S S +72S S S2 S S 112 113 133 222 223 − 112 113 123 133 223 112 113 123 133 222 233 112 113 123 223 233 −

31 Figure 8: The degree 4 invariant I4 of a 3-form in 3 variables, represented as a diagram of tensor contraction. Black 3-valent vertices represent tensor S, white 3-valent vertices represent tensor ǫ.

216S S S2 S S + 36S S S S S S 108S S S S S S 112 113 123 222 333 112 113 122 133 223 233 − 112 113 122 133 222 333 − 216S S S S S2 + 360S S S S S S 36S S S2 S S + 72S S2 S2 S 112 113 122 123 233 112 113 122 123 223 333 − 112 113 122 233 333 112 113 223 233 − 144S S2 S S2 + 72S S2 S S S + 162S2 S2 S2 144S2 S2 S S 112 113 222 233 112 113 222 223 333 112 133 223 − 112 133 222 233 − 216S2 S S S S + 144S2 S S S S + 144S2 S2 S2 72S2 S2 S S + 112 123 133 223 233 112 123 133 222 333 112 123 233 − 112 123 223 333 72S2 S S S2 36S2 S S S S 72S2 S S S S + 18S2 S2 S2 + 112 122 133 233 − 112 122 133 223 333 − 112 122 123 233 333 112 122 333 72S2 S S S2 144S2 S S2 S + 72S2 S S S S 48S3 S3 + 72S3 S S S 112 113 223 233 − 112 113 223 333 112 113 222 233 333 − 112 233 112 223 233 333 − 24S3 S S2 24S S3 S2 + 144S S S2 S S 72S S2 S S2 112 222 333 − 111 133 222 111 123 133 222 223 − 111 123 133 223 − 216S S2 S S S + 72S S3 S S + 120S S3 S S 144S S S2 S2 + 111 123 133 222 233 111 123 223 233 111 123 222 333 − 111 122 133 223 72S S S2 S S + 360S S S S S S 72S S S S S S 72S S S2 S2 111 122 133 222 233 111 122 123 133 223 233 − 111 122 123 133 222 333 − 111 122 123 233 − 216S S S2 S S 144S S2 S S2 + 72S S2 S S S + 144S S2 S S S 111 122 123 223 333 − 111 122 133 233 111 122 133 223 333 111 122 123 233 333 − 24S S3 S2 + 72S S S S3 108S S S S S S + 36S S S S2 S 111 122 333 111 113 133 223 − 111 113 133 222 223 233 111 113 133 222 333 − 72S S S S2 S + 144S S S S S2 72S S S S S S 36S S S S S2 + 111 113 123 223 233 111 113 123 222 233 − 111 113 123 222 223 333 − 111 113 122 223 233 72S S S S2 S 36S S S S S S 36S S S S2 S + 72S S S S S2 111 113 122 223 333 − 111 113 122 222 233 333 − 111 112 133 223 233 111 112 133 222 233 − 36S S S S S S 72S S S S S2 + 144S S S S2 S 111 112 133 222 223 333 − 111 112 123 223 233 111 112 123 223 333 − 72S S S S S S + 72S S S S3 108S S S S S S + 36S S S S S2 + 111 112 123 222 233 333 111 112 122 233 − 111 112 122 223 233 333 111 112 122 222 333 18S2 S2 S2 24S2 S3 S 24S2 S S3 + 36S2 S S S S 6S2 S2 S2 111 223 233 − 111 223 333 − 111 222 233 111 222 223 233 333 − 111 222 333

The algebraic discriminant D3|3, just like any other SL(3)-invariant function of S, is a function of I4 and I6:

3 2 D3|3 = 32I4 +3I6 (51)

When expanded, discriminant D3|3 contains 2040 monomials. See the appendix of the book version of [1], where it is written explicitly. Formula (51) is a remarkably concise expression of this disriminant through a pair of invariants, given by beautiful diagrams Fig.13 and Fig.14. It is an open problem to generalize this type of formulas – eqs. (48), (49), (50) and (51) – to higher n and r. For more information on , see [14].

32 Figure 9: The degree 6 invariant I6 of a 3-form in 3 variables, represented as a diagram of tensor contraction. Black 3-valent vertices represent tensor S, white 3-valent vertices represent tensor ǫ.

4.2 Discriminant through Newton polytopes

A homogeneous polynomial of degree r in n variables

a1 a2 an S(x1, x2,...,xn)= sa1,a2...an x1 x2 ...xn a +a +...+a =r 1 2 n can be visualized as a collection of points with coordinates (a1,...,an) in the n-dimensional . The convex hull of these points is a n-dimensional convex polytope, called Newton polytope of S. It is known, that different triangulations and cellular decompositions of this polytope are in correspondence with different monomials in the discriminant of S. We do not go into details on this topic, see [3] for more details.

4.3 The symmetric case

An important special case is the case of symmetric polynomials S(x1,...,xn), i.e, those which are invariant under of variables x1,...,xn. It turns out that exactly for such polynomials discriminant greatly simplifies and factorizes into many smaller parts. A few examples for low degrees of S can be given to illustrate this phenomenon. The simplest example is the case of degree two, because it is completely treatable by methods of linear algebra. A homogeneous of degree two should have a form

2 S(x1,...,xn)= C2p2 + C11p1 (52)

k where C2, C11 are two arbitrary parameters and pk = i xi . To calculate the discriminant, we take derivatives:

33 ∂S =2C2xi +2C11p1 =0 (53) ∂xi

The simplest option is just to write a matrix and calculate its determinant by usual rules:

C2 + C11 C11 ... C11   C11 C2 + C11 ... C11 n−1 n|2 C2, C11 = det = C C2 + nC11 (54) D n×n   2    . . .     C11 C11 ... C2 + C11      As one can see, the discriminant is highly reducible: factorises into simple elementary constitutients. In fact, this is a general property of symmetric polynomials. The next-to-simplest case is a homogeneous symmetric polynomial of degree 3 in n variables, which has a form

3 S(x1,...,xn)= C3p3 + C21p2p1 + C111p1 (55)

and contains three parameters C3, C21, C111. As shown in [15], the discriminant of this polynomial equals

(n 1)! n−1 2 − n−1 (n−3)2 n 2k 2 4k(n k) 3 k!(n 1 k)! C3,C21,C111 = B3 − B1B + − B − − (56) Dn|3 9n 3 27n2 2 k=0 „ « ! ` ´ ` ´ Y where

2 B1 = n C111 + nC21 + C3, 8 > B2 = nC21 + 3C3, (57) > <> B3 = C3, > > :> This formula is remarkably concise and, most importantly, it contains n just as a parameter and is valid in any dimension n. This property makes it possible to study the large n asymptotics and various continous limits. Moreover, the above examples for r = 2 and r = 3 can be generalized to arbitrary degree. Generally, a symmetric polynomial homogeneous of degree r in n variables has a form

S(x1,...,xn)= CY pY (58) |YX|=r

where the sum goes over Young diagrams (partitions) Y : k1 k2 . . . 0 of fixed sum Y = k1 + k2 + . . . = r, ≥ ≥ | |

34 and pY = i pYi . The number of free parameters is equal to the number of partitions of degree r, which we denote as PQ(r). For example, P (2) = 2 corresponds to partitions (2) and (1, 1). Similarly P (3) = 3 corresponds to partitions (3), (2, 1) and (1, 1, 1). Several first values of P (r) are

r 123456 7 8 910 (59) P (r) 1 2 3 5 7 11 15 22 30 42

The generating function for P (r) is given by

∞ ∞ 1 P (r)qr = (60) 1 qn r=0 n=1 X Y −

A symmetric homogeneous polynomial of degree r has P (r) independent coefficients CY . Its discriminant is a function of these coefficients, which we denote (C). According to [15], this discriminant is equal to Dn|r

#M ! (M1 + . . . + Mr−1)! β n|r (r 1)! M1! ...Mr−1! (C) = const Cr dM (C) (61) Dn|r − M1+...+YMr−1=n “ ”

where the product is taken over all decompositions of n into r 1 nonnegative parts and #M is the number of − zeroes among M1,...,Mr−1. The degree βn|r is fixed by the total degree of the discriminant

#M ! deg dM (M1 + . . . + Mr−1)! βn|r = deg n|r (62) D − (r 1)! M1! ...Mr−1! M1+...X+Mr−1=n − and

d (C)= R P (i), P (ij), P (ijk), ... (63) M 8 M M M 9 i i

(i) 1 yi PM

(i) 0 (j) 1 1 P det 1 yj P M 3×3 M det B C 2×2 0 1 B C (j) B (k) C 1 P B 1 yk P C P (ij) = B M C , P (ijk) = B M C , ... (64) M @ A M @ 2 A 1 yi 1 yi yi det 0 1 2×2 0 2 1 1 yj det 1 yj yj B C 3×3 B C @ A B C B 2 C B 1 yk y C B k C @ A

35 and so on, generally

k−2 (i1) 1 yi1 ... yi1 PM 0 1 k−2 (i2) 1 yi2 ... y P B i2 M C B C det B C k×k B . . . C B C B C B C B k−2 (ik ) C B 1 yik ... yi PM C B k C B C P (i1...ik ) = B C (65) M @ k−2 k−1 A 1 yi1 ... yi1 yi1

0 k−2 k−1 1 1 yi2 ... yi2 yi2 B C B C det B C k×k B . . . C B C B C B k−2 k−1 C B 1 yi ... y y C B k ik ik C B C B C @ A

(i1...ik ) Note, that each of quantities P is a polynomial of degree r k in variables yi. They are defined only for M − pairwise distinct upper indices i1,...,ik and are symmetric in them. For k r they are undefined: one can ≥ not choose r distinct items out of (r 1). As one can see, each particular factor dM (S) can be computed as a − resultant in no more than in (r 1) variables. This allows to calculate discriminants of symmetric polynomials − on the practical level even for n>>r.

5 Integral discriminants and applications 5.1 Non-Gaussian integrals and Ward identities

Resultants and discriminants from the previous sections are purely algebraic (or combinatorial) objects – i.e, they are polynomials, possibly lengthy and complicated, but still finite linear combinations of monomials. The next level of complexity in non-linear algebra is occupied by partition functions, which are usually not polynomials and not even elementary functions. Still, they posess a number of remarkable properties and are related closely to resultants and discriminants. The simplest partition functions of non-linear algebra are integral discriminants

−S(x1,...,xn) Jn|r S = dx1 ...dxn e (66) Z ` ´

For quadratic forms, diagonalisation and other linear-algebra methods can be used to obtain a simple answer −1/2 Jn|2 S = det S , but for higher degrees r such methods do not work. Instead, to calculate integral discriminants` ´ ` one´ needs to use a powerful technique of Ward identities, which are linear differential equations

∂ ∂ ∂ ∂ Jn|r(S) = 0, ai + bi = ci + di (67) ∂sa ,...,a ∂sb ,...,b − ∂sc ,...,c ∂sd ,...,d „ 1 n 1 n 1 n 1 n « satisfied by integral discriminants. Validity of these identities follows simply from the fact, that differential

36 operator in the left hand side of (67) annihilates the integrand in (66). Making use of the Ward identities, in [5] integral discriminants were found in the simplest low-dimensional cases:

n r Invariants Discriminant Dn|r First branch of the integral discriminant Jn|r

−1/2 2 2 I2 I2 I2

−1/6 2 3 I4 I2 I4

∞ 2 i 3 2 −1/4 1 (1/12)i(5/12)i 6I3 2 4 I2,I3 I2 6I3 I2 3 − i! (1/2)i I i=0 „ 2 « P

∞ i j 2 −1/10 1 (3/10)i+j (1/10)2i+3j (1/10)j 16I8 128I12 2 5 I4,I8,I12 I4 64I8 I4 2 3 − i!j! (2/5)i+2j (3/5)i+2j I 3I i,j=0 „ 4 « „ 4 « P

−1/2 3 2 I3 I3 I3

∞ 2 i 3 2 −1/4 1 (1/12)i(5/12)i 3I6 3 3 I4,I6 32I4 + 3I6 I4 3 i! (1/2)i −32I i=0 „ 4 « P

...... where

Γ(a + k) (a)k = a(a + 1) (a + k 1) ≡ Γ(a) −

Already by definition the integral (66) is SL(n)-invariant and, consequently, depends only on the elementary

SL(n)-invariants Ik, which were described in section 4.1. above. It is necessary to note, that Jn|r is a multi-valued function: for one and the same n r there can be several branches, which satisfy one and the same equations | (67), but have different asymptotics. For example, for n r = 2 4 integral discriminant has exactly two branches: | |

2 ∞ 2 i (1) −1/4 1/12, 5/12 6I3 −1/4 1 (1/12)i(5/12)i 6I3 J2|4 S = I2 2F1 3 = I2 3 (68) 1/2 ˛ I2 ! i! (1/2)i I2 ˛ i=0 „ « ` ´ ˛ X ˛ ˛

2 ∞ 2 i (2) −7/4 7/12, 11/12 6I3 −7/4 1 (7/12)i(11/12)i 6I3 J2|4 S = I3I2 2F1 3 = I3I2 3 (69) 3/2 ˛ I2 ! i! (3/2)i I2 ˛ i=0 „ « ` ´ ˛ X ˛ ˛ For the sake of brevity, only one of branches is presented in the table above – the one with normalisation

37 factor depending only on the invariant of lowest degree Imin. Other branches are omitted. As one can see, non- Gaussian integrals posess a nice structure: they are all hypergeometric, i.e, coefficients in their series expansions are ratios of Γ-functions. As shown in [5], in all these cases integral discriminant Jn|r (S) has singularities in the zero locus of the ordinary algebraic discriminant Dn|r(S). This interesting relation between purely algebraic objects (discriminants) and non-Gaussian integrals deserves to be further investigated.

5.2 Action-independence

Actually, an even stronger statement is valid: not only

∂ ∂ ∂ ∂ −S(x1,...,xn) dx1 ...dxn e = 0, ai + bi = ci + di (70) ∂sa ,...,a ∂s − ∂sc ,...,c ∂s „ 1 n b1,...,bn 1 n d1,...,dn « Z but also

∂ ∂ ∂ ∂ dx1 ...dxn f S(x1,...,xn) = 0, ai + bi = ci + di (71) ∂sa1,...,an ∂sb1,...,bn − ∂sc1,...,cn ∂sd1,...,dn „ « Z “ ” with (almost) arbitrary choice of function f(S). Again, the validity of these differential equations obviously follows from the fact, that differential operator in the left hand side annihilates the integrand. As one can see, two different integrals – with integrands e−S and f(S) – satisfy one and the same Ward identities. By itself, this fact does not guarantee equality of these two integrals. However, both integrals are SL(n) invariant and have one and the same degree of homogeneity: for any f(S) we have

−n/r dx1 ...dxn f λS(x1,...,xn) = λ dx1 ...dxn f S(x1,...,xn) (72) Z “ ” Z “ ”

Together, these three properties – common Ward identities, common degree of homogeneity and SL(n) invariance – can be considered as a proof (or, at least, as a strong evidence in favor) of equivalence of these two differently looking integrals: for arbitrary good function f(S), we have

−S(x1,...,xn) f S(x1,...,xn) dx1 ...dxn = const(f) e dx1 ...dxn (73) Z “ ” Z

The constant here depends on the choice of f, but does not depend on S. Property (73) is sometimes called ”action-independence” of integral discriminants. Even before specifying the class of good functions f(S) and 2 proving the action independence, let us consider a simple illustration with f(S)= e−S :

2 2 1 2 2 const e− Sij xixj dnx = e− λ1x1+...+λnxn dnx = e− x1+...+xn dnx = √λ1 ...λn √det S Z ` ´ Z ` ´ Z ` ´

2 2 2 ′ 2 2 1 2 2 const e− Sij xixj dnx = e− λ1x1+...+λnxn dnx = e− x1+...+xn dnx = √λ1 ...λn √det S Z ` ´ Z ` ´ Z ` ´

38 As one can see, these two differently-looking integrals are indeed essentially one and the same (they are pro- portional with S-independent constant of proportionality). To specify the class of good functions and to prove (73), let us make a change of integration variables

x1 = ρ, x2 = ρz2, x3 = ρz3, ...,xn = ρzn

i.e. pass from xi to non-homogeneous coordinates zi = xi/x1. Then

r −S(x1,x2,...,xn) n−1 −ρ S(1,z2,...,zn) e dx1 ...dxn = ρ dρ dz2 ...dzn e = Z Z Z

n−1 −ρr dz2 ...dzn = ρ e dρ n/r S(1,z2,...,zn) „Z « Z

For the right hand side of (73) we have

n−1 r f S(x1,x2,...,xn) dx1 ...dxn = ρ dρ dz2 ...dzn f ρ S(1,z2,...,zn) = Z “ ” Z Z “ ”

n−1 r dz2 ...dzn = ρ f ρ dρ n/r S(1,z2,...,zn) „Z “ ” « Z

Both integrals over ρ are just S-independent constants. As one can see, (73) is valid, iff these integrals

r ρn−1e−ρ dρ and ρn−1f ρr dρ Z Z “ ” are finite over one and the same contour. This condition specifies the class of good functions f(S). For example, all functions f(S) = exp Sk for k > 0 fall into this class. − ` ´ 5.3 Evaluation of areas

Action independence of integral discriminants allows to make different choices for f(S). This explains why integral discriminants appear in a variety of differently looking problems. For example, we can take

1, S 1 ≤ f(S)= θ(1 S)= (74) − 8 <> 0, S> 1

:> which makes sence only in the real setting, i.e, when the contour of integration is real (and the form S is positive definite). This choice of f(S) provides an important geometric interpretation of integral discriminants:

−S(x1,x2,...,xn) (60)−(61) e dx1 ...dxn = dx1 ...dxn = volume, bounded by the S = 1 (75) Z S

39 This relation to bounded volume is a nice practical application of integral discriminant theory. As an example, let us consider the case of J2|4. Take a particular algebraic

S(x,y)=(x2 + y2)2 + ǫx2y2 = 1 (76)

The , bounded by this curve, by definition can be calculated in polar coordinates as

2π dφ 4 ǫ V (ǫ)= = K 2 1+ ǫ cos2 φ sin2 φ √ǫ + 4 ǫ + 4 Z0 „ « p where K is the complete elliptic integral of the first kind [17] and can be represented as

4 ǫ 2π 1/2, 1/2 ǫ V (ǫ)= K = 2F1 (77) √ǫ + 4 ǫ + 4 √ǫ + 4 1 ˛ ǫ + 4 ! „ « ˛ ˛ ˛ ˛ From the integral discriminant side, the homogeneous quartic form S(x,y) has two invariants

8 2 1 2 16 2 1 2 1 3 I2 = + ǫ + ǫ ,I3 = + ǫ ǫ ǫ (78) 3 3 6 9 3 − 6 − 36 and the two branches of the integral discriminant are

−1/4 2 2 2 (1) 8 2 1 2 1/12, 5/12 (ǫ + 2) (ǫ + 8) (ǫ 4) J2|4 S = + ǫ + ǫ 2F1 2 −3 (79) 3 3 6 1/2 ˛ (ǫ + 4ǫ + 16) ! „ « ˛ ` ´ ˛ ˛ ˛ −7/4 2 2 2 (2) 16 2 1 2 1 3 8 2 1 2 7/12, 11/12 (ǫ + 2) (ǫ + 8) (ǫ 4) J2|4 S = + ǫ ǫ ǫ + ǫ + ǫ 2F1 2 −3 (80) 9 3 − 6 − 36 3 3 6 3/2 ˛ (ǫ + 4ǫ + 16) ! „ «„ « ˛ ` ´ ˛ ˛ ˛ As noticed in [5], only one of these two branches stays regular at ǫ 0: namely, →

Γ(3/2) Γ(1/2) J (reg)(ǫ)= 6−1/4 J (1)(ǫ) 6+1/4 J (2)(ǫ) (81) 2|4 Γ(7/12)Γ(11/12) 2|4 − Γ(1/12)Γ(5/12) 2|4 „ « „ «

Therefore, it is this branch that should be identified with the bounded area (regularity of the bounded area in the limit ǫ 0 is fairly obvious). Comparing the bounded area with the regular branch of the integral → discriminant, we find a hypergeometric identity

V (ǫ) = const J (reg)(ǫ) (82) 2|4

40 Eq. (82) is indeed a valid hypergeometric identity, which can be seen by expanding both sides in powers of ǫ:

1 9 25 1225 V (ǫ)= π πǫ + πǫ2 πǫ3 + πǫ4 + . . . = (83) − 16 1024 − 16384 4194304

1 1 9 25 1225 J (reg)(ǫ)= ǫ + ǫ2 ǫ3 + ǫ4 + . . . (84) 2|4 4 − 64 4096 − 65536 16777216 what fixes const = 4π in (82). As one can see, the bounded area and corresponding integral discriminant indeed coincide, up to a constant factor. The above calculation with invariants is certainly not the most practical way to evaluate the integral dis- criminant: it is rather a conceptual demonstration of equivalence between these quantities, emphasizing the importance of correct choice of the regular branch. Usually, a much more convenient way to do is to perform a series expansion directly in the integral (66). Doing so automatically guarantees that a correct branch is chosen, and often allows to calculate the bounded volume in a very convenient way. As an illustration, consider the algebraic hypersurface

2 2 2 2 2 2 2 2 4 a(x1 + x2 + x3) + b(x1 + x2 + x3)x4 + cx4 = ~ (85)

in four-dimensional space. Using the relation with integral discriminants, we find the bounded volume

+∞ +∞ +∞ +∞ a 2 2 2 2 b 2 2 2 2 c 4 V (a,b,c)= dx1 dx2 dx3 dx4 exp (x + x + x ) (x + x + x )x x (86) − ~ 1 2 3 − ~ 1 2 3 4 − ~ 4 −∞Z −∞Z −∞Z −∞Z „ «

Passing to new integration variables

x1 = ρ cos(θ1) 8 > x2 = ρ sin(θ1) cos(φ) > > > <> x3 = ρ sin(θ1) sin(φ)

> > x4 = t > > :> with

0 θ π, 0 φ 2π, 0 ρ , t (87) ≤ ≤ ≤ ≤ ≤ ≤∞ −∞ ≤ ≤∞ and Jacobian ρ2 sin(θ), we obtain

41 2π π ∞ ∞ a b c V (a,b,c)= dφ sin(θ)dθ ρ2dρ dt exp ρ4 ρ2t2 t4 (88) − ~ − ~ − ~ Z0 Z0 Z0 −∞Z „ «

The angular integration is easily done, and we find

∞ ∞ a b c V (a,b,c) = 4π ρ2dρ dt exp ρ4 ρ2t2 t4 (89) − ~ − ~ − ~ Z0 −∞Z „ «

Expanding into series

∞ ∞ ∞ bk a c V (a,b,c) = 4π ρ2+2k exp ρ4 dρ t2k exp t4 dt (90) k!~k − ~ − ~ Xk=0 Z0 “ ” −∞Z “ ” and evaluating both integrals, we find

∞ π bk k 1 k 3 a −3/4−k/2 c −1/4−k/2 V (a,b,c)= Γ + Γ + (91) 2 k!~k 2 4 2 4 ~ ~ Xk=0 „ « „ « “ ” “ ”

The sums are easily evaluated in elementary functions, and we finally obtain

π2 ~ V (a,b,c)= (92) √a b + 2√ac p In this example, the use of integral discriminant allows to obtain a simple formula for the volume in radicals. Thus integral discriminants are useful in application to the calculation of bounded volume.

5.4 Algebraic numbers as hypergeometric series

A characteristic property of integral discriminants, which is evident from definition, is SL(n) invariance: integral

−S(x1,...,xn) Jn|r S = dx1 ...dxn e Z ` ´ does not depend on the choice of basis in the n-dimensional linear space and, consequently, depends only on

SL(n) invariants Ik(S) (this dependence is made explicit in the above summary table). This property is not generic: many other quantities, which belong to the class of partition functions, are not SL(n) invariant. An important example is provided by roots of polynomials, also known as algebraic numbers:

r f0 + f1λ + . . . + frλ = 0

42 This equation defines λ as a function of f0,...,fr:

λ = λ(f0,...,fr)

and this function is naturally multi-valued – there are exactly r branches of λ, since of degree r has exactly r roots in the complex plane. For low r this function can be expressed in radicals. For r = 2 this is merely the solution of :

2 1 f1 f1 4f2f0 λ(f0, f1, f2)= ± − − 2 p f2

Similarly, for r = 3 one of the roots is given by

2 3 3 2 2 2 2 3 1/3 1 f2 (36f1f2f3 108f0f3 8f2 + 12f3 12f1 f3 3f1 f2 54f1f2f3f0 + 81f0 f3 + 12f0f2 ) λ(f0, f1, f2, f3) = + − − − − − 3 f3 p 6f3 −

2 6f1f3 2f − 2 2 3 3 2 2 2 2 3 1/3 − 3f3(36f1f2f3 108f0f 8f + 12f3 12f f3 3f f 54f1f2f3f0 + 81f f + 12f0f ) − 3 − 2 1 − 1 2 − 0 3 2 p and the others are analogously expressed through cubic roots. For r = 4, radicals of order 1/4 are already necessary, and for generic r expression in terms of radicals is impossible (as rigorously shown by Abel and Galois). Still, despite it is impossible to express them through elementary functions, for any r the algebraic numbers satisfy the same differential equations

∂ ∂ ∂ ∂ λ(f0,...,fr) = 0 (93) ∂fa ∂f − ∂fc ∂f „ b d « as integral discriminants do. For example, it is possible to check the Ward identities directly for r = 2:

2 ∂ ∂ ∂ ∂ f1 f 4f2f0 ± 1 − = 0 ∂f0 ∂f2 − ∂f1 ∂f1 f2 „ « p ! similarly for r = 3

2 3 3 2 2 2 2 3 1/3 ∂ ∂ ∂ ∂ 1 f (36f1f2f3 108f0f 8f + 12f3 12f f3 3f f 54f1f2f3f0 + 81f f + 12f0f ) 2 + − 3 − 2 1 − 1 2 − 0 3 2 ∂f0 ∂f2 − ∂f1 ∂f1 − 3 f3 6f3 − „ « p

2 6f1f3 2f − 2 = 0 2 3 3 2 2 2 2 3 1/3 − 3f3(36f1f2f3 108f0 f 8f + 12f3 12f f3 3f f 54f1f2f3f0 + 81f f + 12f0f ) ! − 3 − 2 1 − 1 2 − 0 3 2 p 43 2 3 3 2 2 2 2 3 1/3 ∂ ∂ ∂ ∂ 1 f (36f1f2f3 108f0f 8f + 12f3 12f f3 3f f 54f1f2f3f0 + 81f f + 12f0f ) 2 + − 3 − 2 1 − 1 2 − 0 3 2 ∂f0 ∂f3 − ∂f1 ∂f2 − 3 f3 6f3 − „ « p

2 6f1f3 2f − 2 = 0 2 3 3 2 2 2 2 3 1/3 − 3f3(36f1f2f3 108f0 f 8f + 12f3 12f f3 3f f 54f1f2f3f0 + 81f f + 12f0f ) ! − 3 − 2 1 − 1 2 − 0 3 2 p and also for r = 4. For arbitrary r (when no expressions through radicals are available) validity of Ward identities (93) becomes obvious, if one recalls the conventional Cauchy integral representation for λ:

r r λ(f0,...,fr)= z d log f0 + f1z + . . . + frz = log f0 + f1z + . . . + frz dz (94) − I “ ” I “ ”

All this – multivaluedness, Ward identities and integral representations – make algebraic numbers quite similar to integral discriminants. Of course, since roots are not SL-invariant, they are not literally equal to inte- gral discriminants. Still, as a consequence of Ward identities, solutions of algebraic equations are given by hypergeometric functions, similarly to integral discriminants. The basic example here is the case of r = 2:

2 f0 + f1λ + f2λ = 0 (95)

This equation has two independent solutions, given by explicit formulas

2 2 1 f1 + f1 4f2f0 1 f1 f1 4f2f0 λ+(f0, f1, f2)= − , λ−(f0, f1, f2)= − − (96) − 2 p f2 − 2 p f2

If viewed as a series, these functions take a form

2 3 2 4 3 5 4 6 5 f0 f0 f2 2f0 f2 5f0 f2 14f0 f2 42f0 f2 x1 = 3 5 7 9 11 . . . (97) − f1 − f1 − f1 − f1 − f1 − f1 − and

2 3 2 4 3 5 4 6 5 f1 f0 f0 f2 2f0 f2 5f0 f2 14f0 f2 42f0 f2 x2 = + + 3 + 5 + 7 + 9 + 11 + . . . (98) − f2 f1 f1 f1 f1 f1 f1

Given these series, one can also verify the validity of Ward identities term-by-term. Note again the similarity with integral discriminants: there are two branches, which satisfy one and the same differential equation and differ only by asymptotics (at large f1, one branch grows asymptotically as f1/f2, while the other tends to − zero). The coefficients here are famous Catalan numbers,

44 (2k)! 1, 2, 5, 14, 42,... = (99) k!(k + 1)!

They are and have a celebrated combinatorial meaning of counting the number of tree diagrams, which appear in the Bogolubov-like iteration procedure [16]. However, we are interested not so much in diagrammatical and/or combinatorial meaning of these numbers, but rather in their hypergeometric nature: the solutions are given by explicit hypergeometric series

∞ k+1 k (2k)! f0 f2 f0 1 4f2f0 x1 = 2k+1 = 2F1 , 1 , [2] , 2 (100) − k!(k + 1)! f1 − f1 2 f1 Xk=0 „» – « and

∞ k+1 k f1 (2k)! f0 f2 f1 f0 1 4f2f0 x1 = + 2k+1 = + 2F1 , 1 , [2] , 2 (101) − f2 k!(k + 1)! f1 −f2 f1 2 f1 Xk=0 „» – « which just occasionally reduce to square roots. Note again the similarity with integral discriminants: singularity 2 of these series is situated at the unit value of argument, when f1 = 4f2f0, i.e, when the algebraic discriminant vanishes. In variance with expressions through radicals, expressions through hypergeometric functions can be written for arbitrary r: say, an equation x6 + x + c = 0 which is not solvable in radicals, can be solved in series

x = c c6 6c11 51c16 506c21 5481c26 62832c31 749398c36 + . . . (102) − − − − − − − −

Looking at these numbers, it is easy to find

(6k)! 1, 6, 51, 506, 5481, 62832, 749398,... = (103) k!(5k + 1)!

These numbers are again integer and enumerate sextic (6-ary) trees (rooted, ordered, incomplete) with k vertices (including the root). Moreover, their generating function is again hypergeometric:

∞ (6k)! k+1 1 2 3 4 5 2 3 4 46656c x = c = c 5F4 , , , , , , , , (104) − k!(5k + 1)! − 6 6 6 6 6 5 5 5 3125 Xk=0 „» – » – «

Note, that this function represents only one of the roots: the other roots are represented by other hypergeo- metric functions, which are completely analogous and which we do not include here. For more details about hypergeometricity of solutions of algebraic equations, see [18]. Note also, that integerness of the coefficients is not a generic property, it is due to particular form of the above examples. Say, algebraic equation

(x2 + a)3 +(x + b)4 = 0

45 has a solution

3I 21I x = I + a + 2b a2 6ab + 3Ib2 + . . . 2 − 8 − with already non-integer coefficients and exact formula

3 1 3 ∞ Γ 3s + m Γ 2s + m 1 1 2 − 2 2 − as bm x = I + ( 1)s+1(2I)m+1 „ « „ « 2 − 1 s! m! s,m=0 Γ 2s + m Γ (2s + m) X − 2 „ «

Let us make one more short note about SL(n) invariance. As already mentioned, the roots theirselves are not SL(2) invariant: under transformation G SL(2) they transform as ∈

G11λ + G12 λ , G11G22 G12G21 = 1 (105) → G21λ + G22 −

Thus they are not expressible through the elementary invariants. However, the double ratios

(λi λj )(λk λl) Λijkl = − − (106) (λi λk)(λj λl) − − are invariant under SL(n) transformations (this is easy to check directly). For example, for r = 4

2 3 4 f0 + f1λ + f2λ + f3λ + f4λ = 0 (107)

there are two elementary invariants

1 1 2 I2 = 2f0f4 f1f3 + f (108) − 2 6 2

3 2 3 2 1 1 3 I3 = f0f2f4 f0f f f4 + f1f2f3 f (109) − 8 3 − 8 1 4 − 36 2

and a single double ratio Λ1234, which is SL(2) invariant and thus expressible through I2 and I3. To express

Λ1234 through the elementary invariants, let us consider a particular case

4 f0 = A , f1 = ǫ, f2 = 0, f3 = 0, f4 = 1 (110) −

46 4 2 For these values of parameters I2 = 2A and I3 = 3ǫ /8. It is easy to obtain, that − − 1 ǫ 1 ǫ2 7 ǫ4 1 ǫ5 λ1 = A + + + . . . − 4 A2 − 32 A5 2048 A11 512 A14 8 > 2 4 5 > 1 ǫ 1 ǫ 7 ǫ 1 ǫ > λ2 = AI + + I + I + . . . > 4 A2 32 A5 2048 A11 − 512 A14 > > > 2 4 5 <> 1 ǫ 1 ǫ 7 ǫ 1 ǫ λ3 = A + + + . . . − − 4 A2 32 A5 − 2048 A11 512 A14 > > > 1 ǫ 1 ǫ2 7 ǫ4 1 ǫ5 > λ4 = IA + I I + . . . > − 4 A2 − 32 A5 − 2048 A11 − 512 A14 > > :> These four branches correspond to ǫ-deformations of the four roots of the simple equation λ4 = A4. Accordingly, the double ratio can be calculated as a series in ǫ and equals

2 6 10 (λ1 λ2)(λ3 λ4) 1 3I ǫ 33I ǫ 1665I ǫ Λ1234 = − − = 6 + 18 30 + . . . (111) (λ1 λ3)(λ2 λ4) 2 − 32 A 16384 A − 16777216 A − −

Expressed through invariants, it takes form

3 5 1 √2 I3 11√2 I3 185√2 I3 Λ1234 = . . . (112) 2 − 2 3/2 − 18 9/2 − 108 15/2 − I2 I2 I2

This is the simplest way to express the double ratio through I2 and I3. Because of its SL(2) invariance, this equality will remain the same for arbitrary f0,...,f4, not just for particular values (110). Let us finish this section by noting, that all algebraic numbers – in other words, the set of numbers which are solutions of some algebraic equations – form a field. This in particular means that the set of all algebraic numbers is closed under addition and multiplication. For example, suppose that x1 and x2 are algebraic numbers. I.e, suppose that x1 is a root of f(z) and x2 is a root of g(z). Then polynomials f(z) and g(t z) have a common − root for t = x1 + x2, so that x1 + x2 is a root of resultant of f(z) and g(t z). Since resultant is a polynomial, − x1 + x2 is an . Similarly one can prove that x1 x2, x1x2 and x1/x2 for x2 = 0 are algebraic − numbers. Thus, algebraic numbers actually form a field. By analogy, one could in principle consider 2-algebraic numbers (x,y): solutions of systems of equations

f(x,y) = 0, 8 <> g(x,y) = 0

:> where f and g are some non-homogeneous polynomials, and their straightforward generalisations for arbitrary number n of independent variables. However, multiplication and division of such multi-component objects seems to be unnatural. A more natural direction of generalisation makes use of the similarity between algebraic numbers and integral discriminants: it is interesting to understand, in what exact sence the set of integral discriminants of homogeneous forms is a ring or even a field. This brings us directly into the field of motivic theories (see [19] for a recent review ”for physicists”) which is, however, beyond the scope of this paper.

47 6 Conclusion

We have briefly described (hopefully in explicit and elementary way) several research lines in resultant theory. These include Sylvester-Bezout (determinantal) and Schur (analytic) formulas for the resultant, invariant formu- las and symmetry reductions for the discriminant, Ward identities and hypergeometric series representations for partition functions (integral discriminants and roots of algebraic equations). Unfortunately, several important directions – such as Macaulay matrices [10], Newton polytopes [3], Mandelbrot sets [20] and others – remain untouched in this review. They will be discussed elsewhere. Let us emphasise that the objects and relations, which we consider in present paper, are somewhat under- investigated. Many interesting properties of resultants, discriminants and especially of partition functions have attracted attention only recently. Moreover, new unexpected properties and relations continue to appear. All this indicates, that resultant theory and the entire non-linear algebra is only in its beginning.

Acknowledgements

Our work is partly supported by Russian Federal Nuclear Energy Agency, Russian Federal Agency for Science and Innovations under contract 02.740.11.5029 and the Russian President’s Grant of Support for the Scientific Schools NSh-3035.2008.2, by RFBR grant 07-02-00645, by the joint grants 09-02-90493-Ukr, 09-01-92440-CE, 09-02-91005-ANF and 09-02-93105-CNRS. The work of Sh.Shakirov is also supported in part by the Moebius Contest Foundation for Young Scientists.

References

[1] V. Dolotin and A. Morozov, Introduction to Non-Linear Algebra, World Scientific, 2007; arXiv:hep-th/0609022

[2] E. B´ezout, Th´eorie g´en´erale des Equations Alg´ebriques, 1779, Paris; J.J. Sylvester , On a general method of determining by mere inspection the derivations from two equations of any degree, Philosophical Magazine 16 (1840) 132 - 135; A. Cayley, On the theory of elimination, Cambridge and Dublin Mathematical J. 3 (1848) 116 - 120; F.S. Macaulay , On some Formulae in Elimination, Proceedings of The London Mathematical Society, Vol. XXXV (1903) 3 - 27; A.L. Dixon, The eliminant of three quantics in two independent variables, Proceedings of The London Mathe- matical Society, 6 (1908) 468 - 478

[3] I. Gelfand, M. Kapranov, and A. Zelevinsky, Discriminants, Resultants and Multidimensional Determinants, Birkhauser, 1994; Discriminants of polynomials in several variables and triangulations of Newton polyhedra, Leningrad Mathematical Journal, 2:3 (1991) 499505

[4] A.Morozov, String Theory, What is it?, Sov. Phys. Usp. 35 (1992) 671-714

[5] A. Morozov and Sh. Shakirov, Introduction to Integral Discriminants, arXiv:0903.2595

[6] P. Aluffi and F. Cukierman, Multiplicities of discriminants, Manuscripta Mathematica 78 (1993) 245-458; M.Chardin, Multivariate subresultants. J. Pure Appl. Algebra 101-2 (1995) 129138; Ducos, L. Optimizations of the Subresultant Algorithm, J. Pure Appl. Algebra 145, 149-163, 2000; L.Buse and C.D’Andrea, On the irreducibility of multivariate subresultants, Ct.Rd.Math. 338-4 (2004) 287-290; C.D’Andrea, T.Krick and A.Szanto, Multivariate Subresultants in Roots, arXiv:math.AG/0501281

48 [7] A.Vlasov, Complanart of polynomial equations, arXiv:0907.4249

[8] A.Anokhina, A. Morozov and Sh. Shakirov, Resultant as Determinant of Koszul Complex, arXiv:0812.5013

[9] L.I. Nicolaescu, Notes on the Reidemeister Torsion, http://www.nd.edu/ lnicolae/Torsion.pdf; V.Turaev, Reidemeister torsion in knot theory, Russian Mathematical Surveys, 41:1 (1986) 119182

[10] K. Kalorkoti, On Macaulay Form of the Resultant http://homepages.inf.ed.ac.uk/kk/Reports/Macaulay.pdf

[11] A.Khetan, Exact matrix formula for the unmixed resultant in three variables, arxiv.org:math/0310478

[12] B. Gustafsson and V.Tkachev, The resultant on compact Riemann surfaces, arXiv:math/0710.2326, Comm. Math. Phys. 286 (2009) 313-358; A.Morozov and Sh.Shakirov, Analogue of the identity LogDet = T raceLog for resultants, arXiv:0804.4632; Resultants and Contour Integrals, arXiv:0807.4539

[13] D. Eisenbud and F.O. Schreyer, Resultants and Chow forms via exterior syzygies, J. Amer. Math. Soc. 16 (2003) 537-579, arXiv:math/0111040

[14] N. Beklemishev, Invariants of cubic forms of four variables, Moscow Univ.Bull. 37 (1982) 54-62; D. Hilbert, B. Sturmfels and R. Laubenbacher, Theory of algebraic invariants, Cambridge University Press (1993); H. Derksen and G. Kemper, Computational invariant theory. Invariant Theory and Algebraic Transformation Groups, Springer-Verlag, Berlin, encyclopaedia of Mathematical Sciences 130 (2002); B.Sturmfels, Algorithms in invariant theory, Springer-Verlag, New York (2008)

[15] N.Perminov and Sh.Shakirov, Discriminants of Symmetric Polynomials, arXiv:0910.5757; N. Perminov, The configuratrix and resultant, Hypercomplex Numbers in and Physics, Ed. ”Mozet”, Russia, 1, 6 (2009), 69-72

[16] A. Morozov and M. Serbyn, Non-Linear Algebra and Bogolubov’s Recursion, Theor.Math.Phys. 154 (2008) 270-293, hep-th/0703258

[17] M. Abramowitz and I. Stegun, Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 9th printing. Dover, New York, (1972); I. Gradstein and I. Ryzhik, Tables of Integrals, Series and Products, New York: Academic (1980)

[18] B.Sturmfels, Solving algebraic equations in terms of A-hypergeometric series Disc. Math. 15 13 (2000) 171–181; M. Glasser, The made hard: A less radical approach to solving equations, arXiv:math/9411224

[19] A. Rej and M. Marcolli, Motives: an introductory survey for physicists, arXiv:0907.4046

[20] V. Dolotin and A. Morozov, The Universal Mandelbrot Set. Beginning of the Story, arXiv:hep-th/0501235; Int.J.Mod.Phys. A23 (2008) 3613-3684, arXiv:hep-th/0701234; An. Morozov, Universal Mandelbrot Set as a Model of Phase Transition Theory, JETP Lett. 86 (2007) 745-748, arXiv:nlin/0710.2315

49