• 검색 결과가 없습니다.

Quaternionic Lorentzian space

문서에서 사원수 선형 대수에 관한 연구 (페이지 33-41)

By the proof of Proposition 2.14, we immediately get the following corollary

Chapter 6. Quaternionic Lorentzian space

Quaternionic Lorentzian (n + 1)-space is the inner product space consisting of the right H-module Hn+1 together with the (n + 1)-dimensional Lorentzian inner product.

The Lorentzian inner product is defined by

hx, yi = ¯x1yn+1+ ¯xn+1y1+ ¯x2y2+ · · · + ¯xnyn.

where x = (x1, . . . , xn+1) and y = (y1, . . . , yn+1). We denote by Hn,1 the quaternionic Lorentzian (n + 1)-space.

Definition 6.1. A vector x is said to be spacelike (resp. timelike and lightlike) if hx, xi > 0 (resp. hx, xi < 0 and hx, xi = 0).

Example 6.2 (Spacelike, timelike and lightlike vectors). We give examples of spacelike, timelike and lightlike vectors. First, (0, 1, 0) is a spacelike vector with positive norm.

An example of timelike vector is (i√

2/2, 0, −i√

2/2) whose norm is −1 < 0. Lastly, (1, 0, 0) is lightlike vector whose norm is 0.

Let 0 denote the origin vector (0, 0, . . . , 0) in Hn+1. obvious. Thus we complete the proof.

Lemma 6.4. For x, y ∈ Hn,1, hx, yi = hy, xi.

If hx, yi ∈ R, then hx, yi = hx, yi and thus hy, xi = hx, yi by Lemma

6.4. We have

the following corollary.

Corollary 6.5. If hx, yi ∈ R, then hy, xi = hx, yi.

Let v be a nonzero vector in Hn,1. Denote by v the set of all vectors x ∈ Hn,1 with hv, xi = 0 i.e.,

v= {(x1, . . . , xn) ∈ Hn+1 | ¯v1xn+1+ ¯vn+1x1+ ¯v2x2+ · · · ¯vnxn= 0}.

By Lemma

4.6, v

is a right H-submodule of Hn+1 of dimension n. From now on, we will focus on the case of n = 2.

Lemma 6.6. Let x be a nonzero timelike vector in H2,1. Then x /∈ x and every nonzero vector of x is spacelike.

Proof. First of all, since x is timelike, it follows that hx, xi < 0. Thus obviously x /∈ x. Suppose that y = (y1, y2, y3) ∈ x. Then yq ∈ x for all quaternions q ∈ H.

This follows from hx, yqi = hx, yiq = 0. Furthermore by observing hyq, yqi = ¯qhy, yiq = |q|2hy, yi

in order to prove that y is spacelike, it is sufficient to prove that yq is spacelike for some non-zero quaternion q.

Let x = (x1, x2, x3). From the assumption that x is timelike, it follows that x16= 0.

By scaling x, we may assume that x1 = 1. The conditions of hx, xi < 0 and hx, yi = 0 give

2Re(x3) + |x2|2< 0, (6.1) y3+ ¯x2y2+ ¯x3y1= 0. (6.2) If y1 = 0, then hy, yi = |y2|2. If y1 = 0 and y2 = 0, (6.2) forces y3 = 0, contrary to y 6= (0, 0, 0). Hence if y1 = 0, then y2 6= 0, which implies that hy, yi = |y2|2 is positive.

In other words, y is spacelike.

From now on, we suppose that y1 6= 0. Then by scaling y, we can assume that y1= 1. Then (6.2) is written by y3+ ¯x2y2+ ¯x3 = 0 and

hy, yi = 2Re(y3) + |y2|2

= −2Re¯x2y2− 2Re¯x3+ |y2|2

> −2Re¯x2y2+ |x2|2+ |y2|2 = |x2− y2|2 ≥ 0.

This leads to a conclusion that every non-zero vector in x is spacelike.

Lemma 6.7. Let x be a nonzero spacelike vector in H2,1. Then x /∈ xand xcontains timelike, spacelike and lightlike vectors. Furthermore, there are linearly independent two lightlike vectors in x.

Proof. First of all, since hx, xi > 0, it immediately follows that x /∈ x. We claim that there is a timelike vector in x. Suppose, contrary to our claim, there is no timelike vector in x. Then every vector in x is either spacelike or lightlike. If there is a spacelike vector y ∈ x, then we can find an orthogonal basis {x, y, z} where z ∈ x∩ y. By hypothesis, hz, zi ≥ 0. Then every vector v ∈ H2,1 is written by v = xa + yb + zc for some a, b, c ∈ H and

hv, vi = |a|2hx, xi + |b|2hy, yi + |c|2hz, zi > 0 which contradicts the existence of a lightlike vector in H2,1.

Now we can assume that every vector in x is lightlike. Then there are linear independent vectors y and z in x. Since dimHx = 2, it follows that spanH{y, z} = x. From the assumption that every vector in x is lightlike, for all a, b ∈ H,

0 = hya + zb, ya + zbi = |a|2hy, yi + |b|2hz, zi + 2Re(¯ahy, zib) = 2Re(¯ahy, zib) which leads to a conclusion that hy, zi = 0. Then {x, y, z} is an orthogonal basis of H2,1. Since x is spacelike and, y and z are lightlike, it is derived that every vector has non-negative norm. This also contradicts the existence of a timelike vector in H2,1. Therefore, the claim holds. In other words, there is a timelike vector y ∈ x.

Choosing a vector z ∈ x∩y, we have an orthogonal basis {y, z} of x. By scaling y and z, we may assume that hy, yi = −1 and hz, zi = 1. Since x = spanH{y, z}, it can be easily seen that y + z and y − z are linearly independent lightlike vectors in x. Summarizing, there are a timelike vector y, a spacelike vector z and a lightlike vector y + z in x. Furthermore, there are linearly independent two lightlike vectors y + z and y − z in x.

Lemma 6.8. Let x be a nonzero lightlike vector in H2,1. Then x ∈ x and every vector of x is either spacelike or lightlike. Furthermore, y is a lightlike vector in x if and only if y = xq for some nonzero q ∈ H.

Proof. It follows from the assumption of hx, xi = 0 that x ∈ x. First observe that there are no timelike vectors in x. If there is a timelike vector perpendicular to x, then x must be spacelike, contrary to the hypothesis that x is lightlike. Thus every vector of x is either spacelike or lightlike. It remains to prove the second statement of the Lemma.

Suppose that there is a lightlike vector y such that {x, y} is linearly independent.

Then x= spanH{x, y}. For any a, b ∈ H,

hxa + yb, xa + ybi = |a|2hx, xi + |b|2hy, yi + 2Re(¯ahx, yib) = 2Re(¯ahx, yib).

If hx, yi = 0, then

2Re(¯x1x3) + |x2|2 = 0, (6.3) 2Re(¯y1y3) + |y2|2 = 0, (6.4)

¯

x1y3+ ¯x2y2+ ¯x3y1 = 0, (6.5) where x = (x1, x2, x3) and y = (y1, y2, y3). If x1 = 0, then x2 = 0 by (6.3) and hence x3 6= 0. Then y1 = 0 by (6.5) and y2= 0 by (6.4). This implies that x, y ∈ spanH{e3} and hence, x and y are linearly dependent, contrary to the hypothesis. Thus x1 6= 0 and y16= 0. By scaling x and y, we may assume that x1 = y1 = 1. Then

2Re(x3) + |x2|2 = 0, (6.6)

2Re(y3) + |y2|2 = 0, (6.7)

y3+ ¯x2y2+ ¯x3 = 0. (6.8) If we take the real part on both sides of (6.8), we have

0 = 2Re(y3) + 2Re(¯x2y2) + 2Re(¯x3) = −|y2|2+ 2Re(¯x2y2) − |x2|2= −|x2− y2|2. Therefore x2 = y2. By (6.6) and (6.7), we have Re(x3) = Re(y3). Taking the imaginary part on both sides of (6.8),

0 = Im(y3) + Im(¯x2y2) + Im(¯x3) = Im(y3) + Im(|x2|2) − Im(x3) = Im(y3) − Im(x3).

Since Re(x3) = Re(y3) and Im(x3) = Im(y3), we conclude that x3 = y3. Therefore x = y, contrary to the assumption that x and y are linearly independent. Finally, there are no lightlike vectors in x which are linearly independent with x.

Lemma 6.9. Let V be a 2-dimensional right H-submodule of H3. Then there is a nonzero vector x ∈ H3 such that V = x.

Proof. Since V is a 2 dimensional right H submodule of H3, there exists a basis {u, v}.

V = {uq1+ vq2 | q1, q2∈ H} = spanH{u, v}.

Find an element x ∈ H2,1 such that x∈ V.

Since hu, xi = hv, xi = 0, we have huq1 + vq2, xi = q1hu, xi + q2hv, xi = 0 for any q1, q2∈ H. Let u = (u1, u2, u3), v = (v1, v2, v3), then

u3x1+ u2x2+ u1x3 = 0.

v3x1+ v2x2+ v1x3 = 0.

We can find a soluton of these equations. Because the number of xi is larger than the number of equations.

Corollary 6.10. Let V be a 2-dimensional right H-submodule of H3. Then

(1) if V has linearly independent two lightlike vectors, then V = x where x is space-like,

(2) if V has only one lightlike vectors up to scaling by a nonzero quaternion, then V = x where x is lightlike.

(3) if V has no lightlike vectors, then V = x where x is timelike.

Definition 6.11. A basis {v1, v2, v3} of H2,1 is said to be orthonormal if |hvi, vii| = 1 for i = 1, 2, 3 and hvi, vji = 0 for all distinct i, j.

Example 6.12. An example of an orthonormal basis of H2,1 is

ne1√+e3 con-sists of two spacelike vectors and one timelike vector.

Proof. If everty vector vi is spacelike, for any v = v1a1+ v2a2+ v3a3, hv, vi = |a1|2hv1, v1i + |a2|2hv2, v2i + |a3|2hv3, v3i ≥ 0

which means that every vector in H2,1 is spacelike, contrary to the fact that there is a timelike vector in H2,1. If every vector vi is timelike, every vector in H2,1is timelike in a similar way, contrary to the fact that there is a spacelike vector. Thus any orthonormal basis of H2,1 has at least one spacelike and one timelike vector.

An example of orthonormal basis of H2,1 is {e1+e3

2 , e2,e1−e3

2 }. Here e1+e3

2 and e2 are spacelike vectors and e1−e3

2 is a timelike vector. Let v1 be a spacelike vector and v2, v3 timelike vectors. Then spanH{e1+e3 of H2,1 has two spacelike vectors and one timelike vector.

Let f : H2,1→ H2,1 be a right H-module homomorphism preserving the Lorentzian inner product i.e.

hf (x), f (y)i = hx, yi.

Then the quaternion 3 × 3 matrix A associated to f satisfies AJ A = J where

J =

Let Sp(2, 1) denote the set of all quaternion 3 × 3 matrices A with AJ A = J . Lemma 6.14. Every element of Sp(2, 1) is invertible.

Proof. For any A ∈ Sp(2, 1), AJ A = J and hence J · AJ A = J · J = I. Then it follows that A−1 = J AJ . In other words every element of Sp(2, 1) is invertible.

Proposition 6.15. The set Sp(2, 1) is a group.

Proof. Let A, B and C be matrices in Sp(2, 1). The associativity of quaternions gives A · (B · C) = (A · B) · C. This implies that Sp(2, 1) satisfies the associative law.

Obviously, the identity matrix I is in Sp(2, 1) and thus it is the identity of Sp(2, 1).

By Lemma

6.14, each matrix A in Sp(2, 1) has an inverse A

−1 = J AJ . Moreover, (J AJ )· J · (J AJ ) = J A(J AJ ) = J AA−1= J.

Thus A−1∈ Sp(2, 1). Therefore Sp(2, 1) is a group.

Let V0 be the set of lightlike vectors of H2,1. Let V+ (resp. V−) be the space of spacelike (resp. timelike) vectors whose norms are 1 (resp. −1).

There is a natural Sp(2, 1)-action the space of bases of H2,1 as follows: Let A ∈ Sp(2, 1) and {v1, v2, v3} be an orthonormal basis of H2,1. Since A preserves the Lorentzian inner product, it immediately follows that {Av1, Av2, Av3} is also an orthonormal basis of H2,1.

Definition 6.16 (transitivity). A group action G × X → X is transitive if it possesses only a single group orbit, i.e., for every pair of elements x and y, there is a group element g such that gx = y. In this case, X is isomorphic to the left cosets of the isotropy group, X ∼ G/Gx. The space X, which has a transitive group action, is called a homogeneous space when the group is a Lie group.

If, for every two pairs of points x1, x2 and y1, y2, there is a group element g such that gxi = yi, then the group action is called doubly transitive. Similarly, a group action can be triply transitive and, in general, a group action is k-transitive if every set {x1, ..., yk} of 2k distinct elements has a group element g such that gxi = yi.

Theorem 6.17. The Sp(2, 1)-action on the set of orthonormal bases of H2,1 is tran-sitive.

Proof. Let u1 = (e1 + e3)/√

2, u2 = e2 and u3 = (e1− e3)/√

2. Then we have shown that {u1, u2, u3} is an orthonormal basis of H2,1. To prove the theorem, it suffices to show that for a given orthonormal basis {v1, v2, v3} ∈ F1, there exists an element

A ∈ Sp(2, 1) such that A · {u1, u2, u3} = {v1, v2, v3}. Let |v1| = |v2| = 1 and |v3| = −1.

. Then by a direct computation,

Au1 = A e1+ e3

0. By scaling y, we may assume that hx, yi = 1. According to corollary

6.10, there is a

spacelike vector z such that hz, zi = 1 and z= spanH{x, y}. Define A = (x, z, y). Theorem

6.17, there exists an element A ∈ Sp(2, 1) such that Ae

2 = y. Hence the Sp(2, 1)-action on V+ is transitive. Similarly, one can prove that the Sp(2, 1)-action on V− is transitive.

Proposition 6.19. The Sp(2, 1)-action on V0×V0\∆ is transitive where ∆ = {(x, x) | x ∈ V0}.

Proof. In the proof of Lemma

6.18, we have shown that for any linearly independent

vectors x and y in V0, there exists an element A ∈ Sp(2, 1) such that Ae1 = x and Ae3= y, which implies the transitivity of the Sp(2, 1)-action on V0× V0\ ∆.

Bibliography

[1] N. Jacobson, Basic Algebra I, W. H. Freeman. (1974).

[2] S. Eilenberg and I. Niven, The fundamental theorem of algebra for quaternions, Bull. Amer. Math. Soc. 50, 246–248 (1944).

[3] R. E. Johnson, On the equation χα = γχ + β over an algebraic division ring, Bull.

Amer. Math. Soc. 50, 202–207 (1944).

[4] P. G. Tait An Elementary Treatise on Quaternions, Cambridge University Press (1890).

[5] K. Shoemake, Animating Rotation with Quaternion Curves, Computer Graphics.

19(3), 245–254 (1985).

[6] J. J. Shu & L. S. Ouw, Pairwise alignment of the DNA sequence using hypercomplex number representation, Bulletin of Mathematical Biology. 66(5), 1423–1438 (2004).

[7] J. J. Shu & Y. Li, Hypercomplex cross-correlation of DNA sequences, Journal of Biological Systems. 18(4), 711–725 (2010).

[8] J. L. Brenner, Matrices of quaternions, Pacific J. Math. 1, 329–335 (1951).

[9] Y. H. Au-Yeung, On the convexity of numerical range in quaternionic Hilbert spaces, Linear and Multilinear Algebra. 16, 93–100 (1984).

문서에서 사원수 선형 대수에 관한 연구 (페이지 33-41)

관련 문서