Advanced Algebra II

Total Page:16

File Type:pdf, Size:1020Kb

Advanced Algebra II Advanced Algebra II structure of rings Definition 0.1. For M 2 RM, we said that M is simple if RM 6= 0 and M has no proper submodule. A ring is R said to be simple if R2 6= 0 and R has no proper ideal. Example 0.2. A simple ring R maybe not a simple left RM. Let R := Mat2(k), the 2 £ 2 matrices over a field k. One can check that ideal of R is of the form Mat2(J) for some J C k. Hence R has no proper ideal, hence a simple ring. However, R has two left ideals Ii := f(ajk)jajk = 08k 6= ig, for i = 1; 2. Therefore, as an left RM, R has proper submodule Ii. Example 0.3. A division ring is a simple ring. Proposition 0.4. A simple RM is cyclic, i.e. M = Ra for some a 2 M. Proof. For each a 2 M, we consider Ra < M. Since M is simple, Ra = 0 or Ra = M. Since RM 6= 0, it follows that Ra 6= 0 hence Ra = M for some a 2 M. We can actually prove more. Let B = fa 2 MjRa = 0g. It’s easy to check that B < M. As we have seen that B 6= M. Therefore, B = 0. That is, M = Ra for all a 6= 0 2 M. ¤ » Proposition 0.5. Let M be a simple RM, then M = R=I for some maximal left ideal I. (Note that R=I is not necessarily a ring, we just consider it as a left RM.) Proof. Since M is cyclic, one has that the natural map ' : R ! M by '(r) = ra for a fixed a 6= 0 2 M is surjective. Consider all these as » left RM, then by the first isomorphism theorem, M = R=I for some submodule I := Ker' = A(a) ½ R. It’s clear that I is a left ideal. M is simple implies that I is maximal. ¤ Note that M = Ra, then a = ea for some e 2 R. e play the role as an ”idempotent”. It follows that ra = rea, i.e r ¡ re 2 A(a) = I for all r 2 R. A left ideal I with the property that there is e 2 R with r ¡ re 2 I for all r 2 R is called regular. We can actually prove the following: Theorem 0.6. A left R-module M is simple if and only if M »= R=I for some regular maximal left ideal I. Proof. We have seen one direction (only if part). Suppose now that M »= R=I for some regular maximal left ideal I. Since I is maximal, there is no proper submodule in M. We still need to check that RM »= R(R=I) 6= 0. Since there is e 2 R with r ¡ re 2 I for all r 2 R. We 1 2 have r(e + I) = r + I 6= 0 if r 62 I. Hence R(R=I) 6= 0. Thus R=I is simple. ¤ A similar and important notion is primitive. We need some more notions before we get into primitive. Proposition 0.7. Let A(M) := fr 2 RjrM = 0g be the annihilator of M. Then A(M) C R is an ideal. Moreover, M is a faithful R=A(M) module, i.e. AR=A(M) = 0 as R=A(M) module. Proof. For r 2 R; s 2 A(M), we have (rs)M = r(sM) = 0, hence rs 2 A(M). Also sr(M) = s(rM) < sM = 0, hence sr 2 A(M). The map R £ M ! M induces a natural map R=A(M) £ M ! M. If for some r 2 R such thatr ¯ 2 AR=A(M) ½ R=A(M).rM ¯ = rM = 0, r 2 A(M) and thusr ¯ = 0. ¤ For M 2 RM, we can consider End(M) the endomorphims of the abelian group M. It’s clear that End(M) is a ring. There is a natural map Φ : R ! End(M) define by Φ(r) = Tr such that Tr(x) := rx for x 2 M. Then KerΦ = A(M). By the isomorphism theorem, we have R=A(M) is isomorphic to a subring of End(M). Definition 0.8. A ring R is primitive if there exists a simple faithful left R-module. If R is primitive, let M be a simple faithful R-module, then R is isomorphic to a subring of End(M). Before we proceed to the study of the structure of primitive rings, we take a break by looking at some examples. Proposition 0.9. A simple ring with identity is primitive. Proof. Since R has identity, R has maximal left ideal, say I. Again, R with identity implies that every left ideal of R is regular. Thus R=I is a simple R-module. It remains to show that A(R=I) = 0. This is the case since 1 62 A(R=I) and A(R=I) C R which is simple. ¤ Proposition 0.10. A commutative ring R is primitive if and only if R is a field. Therefore, if R is a commutative ring with identity, then being sim- ple, primitive and being a field are equivalent. However, if R is com- mutative without identity, then it’s possible to be simple but not a field. Proof. A field is simple with identity, hence is primitive. Conversely, let M be a faithful simple R-module. Then M »= R=I for some regular maximal left ideal I. R is commutative, so I is in fact an ideal. M is faithful, thus A(M) = A(R=I) = 0. It’s clear that I ½ A(R=I) = 0. Moreover, I is regular, that is, there exist e 2 R such 3 that r ¡ re = 0. It follows that e is the identity. We have that R is a commutative ring with identity such that 0 is a maximal ideal. Thus R is a field. ¤ The following example also exhibit the difference between simple and primitive. Example 0.11. Let V be a infinite dimensional vector space over a di- vision ring D. Let R := EndD(V ) be the ring of linear transformations. Id clear that V is a left R-module. Claim 1. V has no proper submodule. And thus V is a simple R-module. Claim 2. A(V ) = 0, thus R is primitive. Claim 3. R is not simple. This is because that we have an ideal I = f½ 2 Rjdim(Im½) < 1g. Remark 0.12. If V is finite dimensional, then R is simple and prim- itive. One notice hat End(M) plays an important role here. Inside End(M), there is yet another interesting object Definition 0.13. The commuting ring of R on M is C(M) = fà 2 End(M)jTrà = ÃTr; 8r 2 Rg: One notice that if à 2 C(M), then Ã(rx) = ÃTr(x) = TrÃ(x) = rÃ(x); for all x 2 M. In other words, à is an RM-homomorphism. So C(M) is actually EndR(M). Theorem 0.14 (Schur’s Lemma). If M is a simple module over R, then EndR(M) is a division ring. Proof. Suppose that à 6= 0 2 EndR(M). Let W := ÃM. One sees that W 6= 0 < M. By the simplicity of M, one has W = M. Thus à is surjective. Similarly, one checks that Kerà M. Thus Kerà = 0, à is injective. ¡1 ¡1 Hence à 2 End(M). It’s direct to check that à 2 EndR(M). ¤ Example 0.15. Let k be a field and V be a n-dimensional vector space ¯ over k. Let R = Matn(k) = Endk(V ). If S ½ R, let S be the subalgebra generated by S in R. Then we can view V as a faithful S¯-module. We say that S is irreducible if V is an simple S¯-module. Then by ¯ Schur’s Lemma, C(S) is a divisionµ ring.¶ 0 ¡1 1. Let n = 2, and S = . One can check that S is 1 0 irreducible and C(S¯) »= C. 4 80 1 0 19 > 0 ¡1 0 0 0 0 ¡1 0 > <B C B C= B 1 0 0 0 C B 0 0 0 1 C 2. Let n = 4, and S = @ A ; @ A . > 0 0 0 ¡1 1 0 0 0 > : 0 0 1 0 0 ¡1 0 0 ; One can check that S is irreducible and C(S¯) »= H the real quaternions..
Recommended publications
  • Quaternion Algebras and Modular Forms
    QUATERNION ALGEBRAS AND MODULAR FORMS JIM STANKEWICZ We wish to know about generalizations of modular forms using quater- nion algebras. We begin with preliminaries as follows. 1. A few preliminaries on quaternion algebras As we must to make a correct statement in full generality, we begin with a profoundly unhelpful definition. Definition 1. A quaternion algebra over a field k is a 4 dimensional vector space over k with a multiplication action which turns it into a central simple algebra Four dimensional vector spaces should be somewhat familiar, but what of the rest? Let's start with the basics. Definition 2. An algebra B over a ring R is an R-module with an associative multiplication law(hence a ring). The most commonly used examples of such rings in arithmetic geom- etry are affine polynomial rings R[x1; : : : ; xn]=I where R is a commu- tative ring and I an ideal. We can have many more examples though. Example 1. If R is a ring (possibly non-commutative), n 2 Z≥1 then the ring of n by n matrices over R(henceforth, Mn(R)) form an R- algebra. Definition 3. A simple ring is a ring whose only 2-sided ideals are itself and (0) Equivalently, a ring B is simple if for any ring R and any nonzero ring homomorphism φ : B ! R is injective. We show here that if R = k and B = Mn(k) then B is simple. Suppose I is a 2-sided ideal of B. In particular, it is a right ideal, so BI = I.
    [Show full text]
  • Ring (Mathematics) 1 Ring (Mathematics)
    Ring (mathematics) 1 Ring (mathematics) In mathematics, a ring is an algebraic structure consisting of a set together with two binary operations usually called addition and multiplication, where the set is an abelian group under addition (called the additive group of the ring) and a monoid under multiplication such that multiplication distributes over addition.a[›] In other words the ring axioms require that addition is commutative, addition and multiplication are associative, multiplication distributes over addition, each element in the set has an additive inverse, and there exists an additive identity. One of the most common examples of a ring is the set of integers endowed with its natural operations of addition and multiplication. Certain variations of the definition of a ring are sometimes employed, and these are outlined later in the article. Polynomials, represented here by curves, form a ring under addition The branch of mathematics that studies rings is known and multiplication. as ring theory. Ring theorists study properties common to both familiar mathematical structures such as integers and polynomials, and to the many less well-known mathematical structures that also satisfy the axioms of ring theory. The ubiquity of rings makes them a central organizing principle of contemporary mathematics.[1] Ring theory may be used to understand fundamental physical laws, such as those underlying special relativity and symmetry phenomena in molecular chemistry. The concept of a ring first arose from attempts to prove Fermat's last theorem, starting with Richard Dedekind in the 1880s. After contributions from other fields, mainly number theory, the ring notion was generalized and firmly established during the 1920s by Emmy Noether and Wolfgang Krull.[2] Modern ring theory—a very active mathematical discipline—studies rings in their own right.
    [Show full text]
  • Lectures on Non-Commutative Rings
    Lectures on Non-Commutative Rings by Frank W. Anderson Mathematics 681 University of Oregon Fall, 2002 This material is free. However, we retain the copyright. You may not charge to redistribute this material, in whole or part, without written permission from the author. Preface. This document is a somewhat extended record of the material covered in the Fall 2002 seminar Math 681 on non-commutative ring theory. This does not include material from the informal discussion of the representation theory of algebras that we had during the last couple of lectures. On the other hand this does include expanded versions of some items that were not covered explicitly in the lectures. The latter mostly deals with material that is prerequisite for the later topics and may very well have been covered in earlier courses. For the most part this is simply a cleaned up version of the notes that were prepared for the class during the term. In this we have attempted to correct all of the many mathematical errors, typos, and sloppy writing that we could nd or that have been pointed out to us. Experience has convinced us, though, that we have almost certainly not come close to catching all of the goofs. So we welcome any feedback from the readers on how this can be cleaned up even more. One aspect of these notes that you should understand is that a lot of the substantive material, particularly some of the technical stu, will be presented as exercises. Thus, to get the most from this you should probably read the statements of the exercises and at least think through what they are trying to address.
    [Show full text]
  • RING THEORY 1. Ring Theory a Ring Is a Set a with Two Binary Operations
    CHAPTER IV RING THEORY 1. Ring Theory A ring is a set A with two binary operations satisfying the rules given below. Usually one binary operation is denoted `+' and called \addition," and the other is denoted by juxtaposition and is called \multiplication." The rules required of these operations are: 1) A is an abelian group under the operation + (identity denoted 0 and inverse of x denoted x); 2) A is a monoid under the operation of multiplication (i.e., multiplication is associative and there− is a two-sided identity usually denoted 1); 3) the distributive laws (x + y)z = xy + xz x(y + z)=xy + xz hold for all x, y,andz A. Sometimes one does∈ not require that a ring have a multiplicative identity. The word ring may also be used for a system satisfying just conditions (1) and (3) (i.e., where the associative law for multiplication may fail and for which there is no multiplicative identity.) Lie rings are examples of non-associative rings without identities. Almost all interesting associative rings do have identities. If 1 = 0, then the ring consists of one element 0; otherwise 1 = 0. In many theorems, it is necessary to specify that rings under consideration are not trivial, i.e. that 1 6= 0, but often that hypothesis will not be stated explicitly. 6 If the multiplicative operation is commutative, we call the ring commutative. Commutative Algebra is the study of commutative rings and related structures. It is closely related to algebraic number theory and algebraic geometry. If A is a ring, an element x A is called a unit if it has a two-sided inverse y, i.e.
    [Show full text]
  • Determination of the Differentiably Simple Rings with a Minimal Ideal Author(S): Richard E
    Annals of Mathematics Determination of the Differentiably Simple Rings with a Minimal Ideal Author(s): Richard E. Block Source: The Annals of Mathematics, Second Series, Vol. 90, No. 3 (Nov., 1969), pp. 433-459 Published by: Annals of Mathematics Stable URL: http://www.jstor.org/stable/1970745 Accessed: 07/11/2010 19:03 Your use of the JSTOR archive indicates your acceptance of JSTOR's Terms and Conditions of Use, available at http://www.jstor.org/page/info/about/policies/terms.jsp. JSTOR's Terms and Conditions of Use provides, in part, that unless you have obtained prior permission, you may not download an entire issue of a journal or multiple copies of articles, and you may use content in the JSTOR archive only for your personal, non-commercial use. Please contact the publisher regarding any further use of this work. Publisher contact information may be obtained at http://www.jstor.org/action/showPublisher?publisherCode=annals. Each copy of any part of a JSTOR transmission must contain the same copyright notice that appears on the screen or printed page of such transmission. JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms of scholarship. For more information about JSTOR, please contact [email protected]. Annals of Mathematics is collaborating with JSTOR to digitize, preserve and extend access to The Annals of Mathematics. http://www.jstor.org Determinationof the differentiablysimple ringswith a minimalideal* By RICHARD E.
    [Show full text]
  • Do Simple Rings Have Unity Elements?
    JOURNAL OF ALGEBRA 7, 140-143 (1967) Do Simple Rings Have Unity Elements? J. C. ROBSON Department of Mathematics, The University, Leeds, England Communicated by A. W. Goldie Received October 20, 1966 1. It is well known that a simple artinian ring has a unity element, and that a commutative simple ring, being a field, has a unity element. However, it can easily be seenthat not every simple ring hasa unity element, this being shown in the first proposition of this note. Nevertheless, it could be asked whether a simple right noetherian ring necessarilyhas a unity element and the answer (which comprisesthe main result of this note) is a qualified “yes”. 2. There are simple integral domains which are not division rings-see, for example, [2]. Thus the following result demonstrates that there exist simple integral domains without unity elements. PROPOSITION. Let R be a simple integral domain which is not a division ring and let I be a proper nonzero right ideal of R. Then IR is a simpleintegral domain without a unity element. Proof. Let T be a nonzero ideal of IR. Then IR = I(RTIR) = (IR) T(IR) C T, and so T = IR and IR is a simple integral domain. Say IR has a unity element e. Then e is an idempotent. Thus ear = er for all r E R, and hence er = r since R is an integral domain. Therefore R=eRCIR*R=IRCI, which contradicts the assumptionthat I was a proper right ideal of R. So IR does not have a unity element. 3.
    [Show full text]
  • Some Aspects of Semirings
    Appendix A Some Aspects of Semirings Semirings considered as a common generalization of associative rings and dis- tributive lattices provide important tools in different branches of computer science. Hence structural results on semirings are interesting and are a basic concept. Semi- rings appear in different mathematical areas, such as ideals of a ring, as positive cones of partially ordered rings and fields, vector bundles, in the context of topolog- ical considerations and in the foundation of arithmetic etc. In this appendix some algebraic concepts are introduced in order to generalize the corresponding concepts of semirings N of non-negative integers and their algebraic theory is discussed. A.1 Introductory Concepts H.S. Vandiver gave the first formal definition of a semiring and developed the the- ory of a special class of semirings in 1934. A semiring S is defined as an algebra (S, +, ·) such that (S, +) and (S, ·) are semigroups connected by a(b+c) = ab+ac and (b+c)a = ba+ca for all a,b,c ∈ S.ThesetN of all non-negative integers with usual addition and multiplication of integers is an example of a semiring, called the semiring of non-negative integers. A semiring S may have an additive zero ◦ defined by ◦+a = a +◦=a for all a ∈ S or a multiplicative zero 0 defined by 0a = a0 = 0 for all a ∈ S. S may contain both ◦ and 0 but they may not coincide. Consider the semiring (N, +, ·), where N is the set of all non-negative integers; a + b ={lcm of a and b, when a = 0,b= 0}; = 0, otherwise; and a · b = usual product of a and b.
    [Show full text]
  • Azumaya Algebras
    Azumaya Algebras Dennis Presotto November 4, 2015 1 Introduction: Central Simple Algebras Azumaya algebras are introduced as generalized or global versions of central simple algebras. So the first part of this seminar will be about central simple algebras. Definition 1.1. A ring R is called simple if 0 and R are the only two-sided ideals. Simple rings are only interesting if they are noncommutative because we have the following: Proposition 1.2. If R is a commutative, simple ring. Then R is a field. Proof. Take x a nonzero element in R, then Rx is a nonzero twosided ideal and hence is equal to R. In particular 1 2 Rx and thus x is invertible. Definition 1.3. Let k be a field and A a finite dimensional associative k-algebra. Then A is called a central simple algebra (CSA) over k if A is a simple ring and Z(A) = k Note that the inclusion of k in the center of A is automatic as A is a k algebra. Example 1.4. Let n be some natural number, then the matrix ring Mn(k) is a CSA over k. It obviously has dimension n2 over k so we only need to check that it is central and simple. To see this, let eij denote the matrix with a 1 at position (i; j) and zeroes at all other positions, i.e. 20 0 ··· 03 6 .. 7 60 . 7 6 7 6 . 7 eij = 6 . 1 7 6 7 6 .. 7 4 . 5 0 0 Then for a matrix eiiM = Meii for all i implies that M is diagonal and eijM = Meij for all i and j implies that all entries on the diagonal must be the same.
    [Show full text]
  • 7 Rings with Semisimple Generators
    7 Rings with Semisimple Generators. It is now quite easy to use Morita to obtain the classical Wedderburn and Artin-Wedderburn charac- terizations of simple Artinian and semisimple rings. We begin by reminding you of a few elementary facts about semisimple modules. Recall that a module RM is simple in case it is a non-zero module with no non-trivial submodules. 7.1. Lemma. [Schur’s Lemma] If M is a simple module and N is a module, then 1. Every non-zero homomorphism M → N is a monomorphism; 2. Every non-zero homomorphism N → M is an epimorphism; 3. If N is simple, then every non-zero homomorphism M → N is an isomorphism; In particular, End(RM) is a division ring. A module RM is semisimple if it is the sum of its simple submodules. Then an easy, but important, characterization of semisimple modules is given in [1] (see Theorem 9.6): 7.2. Proposition. For left R-module RM the following are equivalent: (a) M is semisimple; (b) M is a direct sum of simple submodules; (c) Every submodule of M is a direct summand. Let’s begin with a particularly nice example. Indeed, let D be a division ring. Then the regular module DD is a simple module, and hence, a simple progenerator. In fact, every nitely generated module over D is just a nite dimensional D-vector space, and hence, free. That is, the progenerators for D are simply the non-zero nite dimensional D-vector spaces, and so the rings equivalent to D are nothing more than the n n matrix rings over D.
    [Show full text]
  • Central Simple Algebras and the Brauer Group
    Central Simple Algebras and the Brauer group XVIII Latin American Algebra Colloquium Eduardo Tengan (ICMC-USP) Copyright c 2009 E. Tengan Permission is granted to make and distribute verbatim copies of this document provided the copyright notice is preserved on all copies. The author was supported by FAPESP grant 2008/57214-4. Chapter 1 CentralSimpleAlgebrasandthe Brauergroup 1 Some conventions Let A be a ring. We denote by Mn(A) the ring of n n matrices with entries in A, and by GLn(A) its group of units. We also write Z(A) for the centre of A×, and Aop for the opposite ring, which is the ring df with the same underlying set and addition as A, but with the opposite multiplication: a Aop b = b A a op ≈ op × × for a,b A . For instance, we have an isomorphism Mn(R) Mn(R) for any commutative ring R, given by∈M M T where M T denotes the transpose of M. → 7→ A ring A is simple if it has no non-trivial two-sided ideal (that is, an ideal different from (0) or A). A left A-module M is simple or irreducible if has no non-trivial left submodules. For instance, for any n field K, Mn(K) is a simple ring and K (with the usual action) is an irreducible left Mn(K)-module. If K is a field, a K-algebra D is called a division algebra over K if D is a skew field (that is, a “field” with a possibly non-commutative multiplication) which is finite dimensional over K and such that Z(D)= K.
    [Show full text]
  • On the Essential Dimension of Central Simple Algebras
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Electronic Thesis and Dissertation Archive - Università di Pisa Dipartimento di Matematica Tesi di laurea magistrale On the Essential Dimension of Central Simple Algebras Relatore: Prof. Angelo Vistoli Candidato: Justin Lacini Controrelatore: Prof. Marco Franciosi Anno Accademico 2014/2015 2 Contents 1 Noncommutative rings 9 1.1 General properties . 9 1.2 Central simple algebras . 12 1.3 The Brauer Group . 16 1.4 Azumaya algebras . 19 2 Descent theory 23 2.1 Projective linear group scheme . 23 2.2 Equivalence . 24 2.3 Cohomology . 27 3 Essential dimension 31 3.1 General properties . 31 3.2 Essential p-dimension . 32 3.3 Essential dimension of algebraic groups . 33 3.4 Versal pairs . 35 4 Computations 41 4.1 Upper bounds . 41 4.1.1 Essential dimension of crossed products . 41 4.1.2 Brauer factor sets . 44 4.1.3 Universal algebras . 46 4.1.4 Essential p-dimension . 47 4.2 Lower bounds . 50 4.2.1 Preliminaries . 50 4.2.2 Brauer group and algebraic tori . 52 4.2.3 Essential dimension of algebraic tori . 54 4.2.4 Degeneration . 56 4.2.5 Multiple degeneration . 58 3 4 CONTENTS Introduction In this thesis we present the notion of essential dimension and give estimates for the essential dimension of the algebraic group P GLn. Essential dimension was introduced for finite groups by J.Buhler and Z.Reichstein in [5] and was extended to the class of algebraic groups over algebraically closed fields by Z.Reichstein in [25].
    [Show full text]
  • 1844, Hamilton Discovered the Quaternions and Kelly, Graves Independently Discovered the Octonions
    COMPLEX RINGS, QUATERNION RINGS AND OCTONION RINGS ISAO KIKUMASA, KAZUTOSHI KOIKE AND KIYOICHI OSHIRO Abstract. In [2], for any ring R, complex rings C(R), quaternion rings H(R) and octonion rings O(R) are studied. For the real numbers R, H(R) is the Hamilton's quaternions and O(R) is the Kelly-Graves's( ) octonions. In view of progress of quater- a;b nions, generalized quaternion algebras F are introduced for commutative fields F and nonzero elements a; b 2 F , and these quaternions have been extensively( ) studied as alge- a;b braic number theory. In this paper, we use H(F ; a; b) instead of F . For a division ring D and nonzero elements a; b 2 Z(D), the center of D, we introduce generalized complex rings C(D; a) and generalized quaternion rings H(D; a; b), and study the structure of these rings. We show that these rings are simple rings if the characteristic of D is not 2, that is, 2 =6 0, and study the structure of these simple rings. When 2 = 0, these rings are local quasi-Frobenius rings. 1. Introduction In 1843 - 1844, Hamilton discovered the quaternions and Kelly, Graves independently discovered the octonions. These numbers are defined over the real numbers and contain the complex numbers. Through Frobenius, Wedderburn and Noether, these numbers have been studied by many mathematicians. We may say that one of the roots of our ring and representation theory began with these numbers. In order to define these numbers for any ring R, we consider free right R-modules: C(R) = e0R ⊕ e1R; H(R) = e0R ⊕ e1R ⊕ e2R ⊕ e3R; O(R) = e0R ⊕ e1R ⊕ · · · ⊕ e7R: We define rei = eir for any r 2 R and any i (1 ≤ i ≤ 7), and multiplications for feigi are defined by the following Cayley-Graves multiplication table: × e0 e1 e2 e3 e4 e5 e6 e7 e0 e0 e1 e2 e3 e4 e5 e6 e7 e1 e1 -e0 e3 -e2 e5 -e4 -e7 e6 e2 e2 -e3 -e0 e1 e6 e7 -e4 -e5 e3 e3 e2 -e1 -e0 e7 -e6 e5 -e4 e4 e4 -e5 -e6 -e7 -e0 e1 e2 e3 e5 e5 e4 -e7 e6 -e1 -e0 -e3 e2 e6 e6 e7 e4 -e5 -e2 e3 -e0 -e1 e7 e7 -e6 e5 e4 -e3 -e2 e1 -e0 The detailed version of this paper will be submitted for publication elsewhere.
    [Show full text]