Saturday, November 29, 2008

Chateau Paradis Casseuil 2006

Château Paradis Casseuil gets its name from the combination of the registered name of the main parcel of vineyards called “Vines of Paradise” and Casseuil county. Taken under Domaines Barons de Rothschild (Lafite)’s wing in 1984, Château Paradis Casseuil then included 14 hectares of vines. In 1989, the estate grew by 9 hectares, and chais were included in the heart of the Sainte Foy la Longue vineyard. The Château Paradis Casseuil chais, located in the heart of the Sainte Foy la Longue vineyard are used for producing red wines. White wines are made at Château Rieussec, benefiting from the technical capacities of that great estate.

Lively and intense ruby color. Fine nose, red fruit and slight licorice aromas. The attack is supple (slight sensation of sweetness) with silky tannins.

2007
After a wet winter, the high temperatures in March and April helped give a good start to the vegetation. The following months were moderate until August. The fine weather settled early September encouraging the ripening of the grapes.

Beautiful crimson colour. A fresh nose with a touch of redcurrant and mint. The first impression on the palate is pleasant, frank with intense fruit. This wine can be appreciated now or kept for a few years when it will have reached its peak.

2006
The winter of 2005/2006 was dry and cold and the spring months were mild with little rain. August was disconcerting – cool and wet – and at the beginning of September ripeness levels were very low. However, the weather then became summery allowing the grapes to finish ripening well.

Beautiful straw yellow colour with hints of green. Discreet on the nose, but when swirled, the aromas are revealed with a fine lime bouquet. The attack is supple and the finish is fresh.

2005
The end of 2004 and the first few months of 2005 were dry. Moreover, maturation took place in perfect conditions.

Pale yellow colour. Notes of citrus fruit, mainly grapefruit, on the nose.

First impressions on the palate are full, lively and well rounded, leading to a silky finish marked by hints of hazelnuts.

2004
The year was marked by stormy weather until July, with no effects on the vines.
The beginning of the year was warmer than in 2003, but from March onwards the trend was reversed and an average drop of 2°C was noted. Rainfall was about the average for the past three vintages, with a dry June. July and August were very damp. Maturation was therefore slow but at harvest time the grapes for the dry whites were ripe.

Pale yellow colour. Very open and fresh on the nose: aromas of white flowers and violets.

Delicate first impressions on the palate. Notes of fresh fruit and Granny Smith apples.

Region : Sainte Foy la Longue, Médoc
Grape Varietals : Cabernet Sauvignon 50%, Merlot 45% and Cabernet franc 5%
Average wine production : 12 000 cases per year.


Thursday, November 27, 2008

Use of SU(2) & SO(3)

Every Lie group homomorphism gives rise to a Lie algebra homomorphism. In the case of a simply-connected matrix Lie group G, a Lie algebra homomorphism also gives rise to a Lie group homomorphism. In fact, for a simply-connected matrix Lie group, there is a natural one-to-one correspondence between the representations of G and the representations of the Lie algebra g. Each of the representations πm of su(2) was constructed from the corresponding representation Πm of the group SU(2).


SU(2) is simply connected but SO(3) is not (SU(2) can be thought of (topologically) as the three-dimensional sphere S3 sitting inside R4. It is well known that S3 is simply connected). There exists a Lie group homomorphism Φ which maps SU(2) onto SO(3) and which is two-to-one. Therefore, SU(2) and SO(3) are almost isomorphic.


Consider the space V of all 2x2 complex matrices which are self-adjoint (i.e., A* = A) and have trace zero. This is a three-dimensional real vector space with the following basis: A1 = (0,1: 1,0) ; A2 = (0,-i : i,0) ; A3 = (1,0 : 0,-1) . Define the inner product on V by <A, B> = 1/2 trace(AB). {A1, A2, A3} is an orthonormal basis for V. Next we are going to identify V with R3. Suppose U is an element of SU(2) and A is an element of V. Consider UAU-1, trace(UAU-1) = trace(A) = 0 and (UAU-1)* = UAU-1. So UAU-1is again in V. Because the map A --> UAU-1 is linear. Therefore, we can define a linear map ΦU of V to itself by ΦU = UAU-1. Given A, B V, <ΦU(A), ΦU(B)> = <A, B>. Thus, ΦU is an orthogonal transformation of V.


Once we identify V with R3 using the above orthonormal basis, we may think of ΦU as an element of O(3). Since ΦU1U2 = ΦU1ΦU2, we see that Φ (the map U --> ΦU) is a homomorphism of SU(2) into O(3). SU(2) is connected, Φ is continuous, and ΦI is equal to I, which has determinant one. It follows that Φ must map SU(2) into the identity component of O(3), namely SO(3). However, the map ΦU is not one-to-one, since for any U SU(2), ΦU = Φ-U. Actually ΦU is a two-to-one map of SU(2) onto SO(3) (recall that every element of O(3) has determinant ± 1).


If we have the basis E1 = 1/2(i,0 : 0,-i) ; E2 = 1/2(0,-1: 1,0) ; E3 = (0,i : i,0) for su(2) and the basis F1 = (0,0,0 : 0,0,1 : 0,-1,0) ; F2 = (0,0,-1 : 0,0,0 : 1,0,0) ; F3 = (0,1,0 : -1,0,0 : 0,0,0) for so(3). Then, we have [E1 , E2] = E3 , [E2 , E3] = E1 and [E3 , E1] = E2 , and similarly with the E's replaced by the F's. Thus the linear map Φ : su(2) --> so(3) which takes Ei to Fi will be a linear algebra isomorphism.


Let σm = πm º Φ-1 be the irreducible complex representations of the Lie algebra so(3) (m ≥ 0). If m is even, then there is a representation ∑m of the group SO(3) such that ∑m(exp X) = exp(σm(X)) for all X in so(3). If m is odd, then there is no such representation of SO(3).


Representation of su(2) so(3) in Physics

Representation of su(2) so(3) in Physics are labeled by the parameter l = m/2. In terms of this notation, a representation of so(3) comes from a representation of SO(3) if and only if l is an integer. The representations with l an integer are called "integer spin"; the others are called the "half-integer spin."Consider the path in SO(3) consisting of rotations by angle 2πt in the (x, y)-plane, which comes back to the identity when t = 1. However, this path is not homotopic to the constant path.


If one defines ∑m along the constant path, then one gets the value ∑m(I) = I, as expected. If m is odd and one defines ∑m along the path of rotations in the (x, y)-plane, then one gets the value ∑m(I) = -I. There is no way to define ∑m(m odd) as a "single-valued" representations of SO(3).


An electron is a "spin-1/2" particle, which means that it is described in quantum machines in a way that involves the representation σ1 of so(3). In the quantum machines, one finds statements to the effect that performing a 360º rotation on the wave function of the electron gives back the negative of the original wave function.This reflects that if one attempts to construct the nonexistent representation1 of SO(3), then when defining ∑1 along a path of rotations in some plane, one gets that ∑1(I) = -I.


A Unitary Representations of SO(3)

Consider the unit sphere S2 R3, with the usual surface measure Ω. Any R SO(3) maps S2 into S2 . For each R, we can define Π2(R) acting on L2(S2, dΩ) by [ Π2(R) f](x) = f(R-1x). Then, Π2 is a unitary representation of SO(3). Here, L2(S2, dΩ) has a very nice decomposition as the orthogonal direct sum of finite-dimensional invariant subspaces. This decomposition is the theory of "spherical harmonics" in physics.

Representations of SU(2) & su(2)

su(2) so(3) and the representation of so(3) is important in the computation of angular momentum. By studying the representation theory of su(2) we can know:

(1) how to commutation relations to determine the representations of a Lie algebra.

(2) how to determine the representations of semisimple Lie algebras, e.g., su(3).


Some Representations of SU(2)

By definition, an element of U of SU(2) is a linear transformation of C2. Let z denotes the (z1, z2) pair in C2. Then, we may define a linear transformation Πm(U) on the space Vm by the formula [Πm(U)f] (z) = f(U-1z). The inverse is necessary in order to make Πm a representation.  And Vm is the space of functions of the form f(z1, z2) = a0z1m + a1z1m-1z2  + a2z1m-2z22  + ... + amz2m .
 

Therefore, [Πm(U)f] (z1, z2) =  ∑ ak(U11-1z1 + U12-1z2)m-k (U21-1z1 + U22-1z2)k for k = 0 .. m.  Πm(U)f is a homogeneous polynomial of degree m. Thus, Πm(U) maps Vm into Vm. Moreover, Πm(U1) [Πm(U2)f] (z) = Πm(U1U2) f(z). So Πm is a (finite-dimensional complex) representation of SU(2).


The Lie algebra representation of Πm can be computed as πm(X) = d/dt Πm(etX) |t=0. So (πm(X)f) (z) = d/dt f(e-tXz) |t=0. Let z(t) be the curve e-tXz. We have z(0) = z and (dz/dt) |t=0 = -Xz. Since z(t) can also be written as z(t) = (z1(t), z2(t)), with zi(t) C. By chain rule, πm(X)f = ∂f / ∂z1(dz1/dt) |t=0 + ∂f / ∂z2(dz2/dt) |t=0 . We have

πm(X)f = -∂f / ∂z1(X11z1 + X12z2)  - ∂f / ∂z2(X21z1 + X22z2). 


Because every finite-dimensional complex representation of the Lie algebra su(2) extends uniquely to a complex-linear representation of the complexification of su(2). And the complexification of su(2) is sl(2;C). Therefore, the representation πm of su(2) given by above extends to a representation of sl(2;C). 


Consider H = (1,0 : 0,-1)

m(H)f) (z) = -(∂f / ∂z1)z1 + (∂f / ∂z2)z2 .

 πm(H) = -z1(∂f / ∂z) + z2(∂f / ∂z2). Apply this to a basis element z1kz2m-k , we have

πm(H) z1kz2m-k = -k z1kz2m-k  + (m-k)z1kz2m-k  = (m-2k) z1kz2m-k .

Thus, z1kz2m-k is an eigenvector for πm(H) with eigenvalue (m-2k). In particular, πm(H) is diagonalizable. 


Let X and Y be the elements, X = (0,0 : 1,0) , Y = (0,1 : 0,0) in sl(2;C). We have πm(X) = -z2(∂f / ∂z1) and πm(X) = -z1(∂f / ∂z2). Apply these to the basis element 

πm(X) z1kz2m-k = -k z1k-1z2m-k+1  

πm(Y) z1kz2m-k = -(m-k) z1k+1z2m-k-1  

It suffices to show that every nonzero invariant subspace of Vm is equal to Vm. Let W be such a space. Since W is assumed nonzero, there is at least one nonzero element w in W. Then w can be written uniquely in the form 

w =  a0z1m + a1z1m-1z2  + a2z1m-2z22  + ... + amz2m

with at least one of the ak's nonzero. Let k0 be the smallest value of k for which  ak ≠ 0 and consider πm(X)m-k0 w. Since πm(X) lowers the power of z1 by 1, it will kill all the terms in w except ak0 z1m-k0 z2k0 . So we have  πm(X)m-k0 z1m-k0 z2k0 = (-1)m-k0(m-k0)! z2m . Since W is assumed invariant, W must contain z2m . Furthermore, πm(Y)k z2m is a nonzero multiple of z1kz2m-k for all 0 ≤ k ≤ m. Because these elements form a basis for Vm . In fact W = Vm. Therefore, representation πm is an irreducible representation of sl(2;C). 


Irreducible Representations of su(2)

Every finite-dimensional complex representation π of su(2) extends to a complex-linear representation of the complexification of su(2), namely sl(2;C). Studying the irreducible representations of su(2) is equivalent to studying the irreducible representation of sl(2;C). Passing to the complexified Lie algebra makes computations easier, in that there is a nice basis for sl(2;C) that has no counterpart among the bases of su(2).


We can use commutation relations to determine the representation of a Lie algebra. Consider the following basis for sl(2;C) and commutation relations:

H = (1,0 : 0,-1);  X = (0,0 : 1,0);  Y = (0,1 : 0,0)

[H, X] = 2X,  [H, Y] = -2Y,  [X, Y] = H.

If V is a (finite-dimensional complex) vector space and A, B, and C are operators on V satisfying

[A, B] = 2B,  [A, C] = -2C,  [B, C] = A, then

because of the skew symmetry and bilinearity of brackets, the linear map π : sl(2;C) --> gl(V) satisfying π(H) = A, π(X) = B, π(Y) = C will be a representation of sl(2;C).


We call π(X) the "raising operator", because it has the effect of raising the eigenvalue of π(H) by 2, and call π(Y) the "lowering operator". Since [π(H), π(X)] = π([H, X]) = 2π(X). Let u be an eigenvector of π(H) with eigenvalue α C. Thus,

π(H)π(X)u = π(X)π(H)u + 2π(X)u

= π(X)(αu) + 2π(X)u = (α + 2)π(X)u.

Either π(X)u = 0 or π(X)u is an eigenvector for π(H) with eigenvalue α+2. More general, π(H)π(X)nu = (α + 2n)π(X)nu. Similarly, for [π(H), π(Y)] = -2π(Y), we have π(H)π(Y)u = (α - 2)π(Y)u.


An operator on a finite-dimensional space can have only finitely many distinct eigenvalues. Therefore, there is some N ≥ 0 such that π(X)N+1u = 0.

Define u0 = π(X)Nu and λ = α + 2N. Then,

π(H)u0 = λu0, π(X)u0 = 0. 


Define uk = π(Y)ku0 , for k ≥ 0. Thus, we have π(H)uk = (λ - 2k)uk . Since π(H) can have only finitely many eigenvalues, the uk's cannot be all be nonzero.

For k = 1,

π(H)u1 = π(H)π(Y)u0 = (α - 2)π(Y)u0 = (α - 2)u1.

π(X)u1 = π(X)π(Y)u0 = (π(Y)π(X) + π(H))u0 = π(H)u0 = λu0 (as π(X)u0 = 0)

π(Y)u1 = π(Y)π(Y)u0  = π(Y)2u0 = u2  

If π(X)uk = [kλ - k(k-1)]uk-1. By induction, π(X)uk+1= π(X)π(Y)uk

= (π(Y)π(X) + π(H))uk = π(Y)π(X)uk + (λ - 2k)uk

= π(Y)[kλ - k(k-1)]uk-1 + (λ - 2k)uk  = [kλ - k(k-1) + (λ - 2k)]uk

= [(k+1)λ - (k+1)k]uk  


Because π(H) can have only finitely many eigenvalues, the uk's cannot all be nonzero. For all k ≤ m, um+1 = π(Y)m+1u0  = 0. If um+1 = 0. Then π(X)um+1  = (m+1)(λ - m)um  = 0. Since m ≠ 0 and m + 1 ≠ 0. So we have λ = m, where m is a non-negative integer. 


In summary, given a finite-dimensional irreducible representation π of sl(2;C) acting on a space V and putting λ = m, there exists an integer m ≥ 0 and nonzero vectors u0, ..., um such that

π(H)uk = (m - 2k)uk, 

π(Y)uk = uk+1 (k < m),

π(Y)um = 0, 

π(X)uk = [km - k(k-1)]uk-1 (k > 0),

π(X)u0 = 0.

The vectors u0, ..., um must be linearly independent, since they are eigenvectors of π(H) with distinct eigenvalues. Moreover, the (m+1)-dimensional span u0, ..., um is explicitly invariant under π(H), π(X), and π(Y). Hence under π(Z) for all Z sl(2;C). Since π is irreducible, this space must be all of V. 


The (m+1)-dimensional representation Πm described above must be equivalent to π. This can be seen explicitly by introducing the following basis for Vm :

uk = [πm(Y)]k (z2)m = (z2)m (m! / (m-k)!) z1kz2m-k  (k ≤ m).

In other words, πm have a basis of the form π(H), π(X), and π(Y). π's have the right commutation relations to forma representation of sl(2;C) and that this representation is irreducible. 

Wednesday, November 19, 2008

Complexification

Complexification of a Real Lie Algebra

The complexification of a finite-dimensional real vector space V, as denoted by Vc, is the space of formal linear combinations v1+ iv2 , with v1, v2 V. 


Let g be a finite-dimensional real Lie algebra and gc its complexification (as a real vector space). Then, the bracket operation on g has a unique extension to gc which makes gc into a complex Lie algebra. The complex Lie algebra gc is called the complexification of the real Lie algebra g


Isomorphisms of Complex Lie Algebra

The Lie algebra gl(n;C), sl(n;C), so(n;C), and sp(n;C) are complex Lie algebras. In addition, there are also following isomorphisms of complex Lie algebras:

gl(n;R)c  gl(n;C)

u(n)c  gl(n;C)

su(n)c  sl(n;C)

sl(n;R)c  sl(n;C)

so(n)c  so(n;C)

sp(n;R)c  sp(n;C)

sp(n)c  sp(n;C)

(u(n) is the space of all nxn complex skew-self-adjoint matrices. If X is any nxn complex matrix, then X = (X-X*)/2 + i(X+X*)/2i. Thus, X can be written as a skew matrix plus i times a skew matrix. Every X in gl(n;C) can be written uniquely as X1 + iX2, with X1 and X2 in u(n). It follows that u(n)c  gl(n;C). If X has trace zero, then so do X1 and X2 , which has su(n)c  sl(n;C))


Note that u(n)c  gl(n;R)c  gl(n;C). However, u(n)c is not isomorphic to gl(n;R), except when n = 1. The real algebra u(n) and gl(n;R) are called real forms of the complex Lie algebra gl(n;C). 


In physics, we do not always clearly distinguish a matrix Lie algebra and its Lie algebra, or between a real Lie algebra and its complexification. For example, some references in the literature to SU(2) actually refer to the complexified Lie algebra sl(2;C). 


Representation and Complexification

Let g be a real Lie algebra and gc its complexification. Then, every finite-dimensional complex representation π of g has a unique extension to a complex-linear representations of gc , also denoted as π and given by π(X+iY) = π(X) + iπ(Y) for all X, Y g. Furthermore, π is irreducible as a representation of gc  if and only if it is irreducible as a representation of g.


If π is a complex representation of the real Lie algebra g, acting on the complex vector space V. Then, saying that π is irreducible means that there is no nontrivial invariant complex subspace W V. Even though g is a real Lie algebra, when considering complex representations of g, we are interested only in complex invariant subspaces.


Credit Default Swaps

Nov 18 (Reuters) - A report on Bloomberg said the cost of protecting Berkshire's debt against default using credit default swaps (CDS) has risen from 140 (two months ago) to 388 basis points, or $388,000 a year to protect $10 million for five years. BRK-A shares fall 4% to $91,700 after the report.

Sunday, November 16, 2008

Chateau Sainte Barbe 2004 Merlot

Blind Test
After a blind tasting of 350 Bordeaux wines by the Belgium Wine tasting committee of the Revue Vino magazine, Merlot Sainte Barbe 2004, Chateau Sainte Barbe 2003 & 2004 were selected amongst the finalists and Chateau Sainte Barbe was awarded 2 Bacchus.

The wine's taste is fruity with harmonious aromas of ripe grapes, plums and white flowers. On the palate (roof of the mouth), it combines a well-balanced structure with a velvety (closely woven fabric of silk), fine and voluptuous flavour. Evolution: 3 to 5 years

Region: Gironde, right bank of the Garonne, Bordeaux
Grape Varieties: 100% Merlot



Reference
"In 2004, observing how some parcels were reacting to our husbandry, we decided to produce a still higher quality wine, called 'Cuvée VSP' lowering the yield to an extreme 4 bunches of grapes per vine, using picking boxes for hand harvesting and making malolactic fermentation directly in 100% new oak barrels.

The results have exceeded even our expectations."



Complex wine with an aroma of blackberries, blackcurrants, chocolate, and expresso. Medium to full bodied with rich ripe tannins.

Thursday, November 13, 2008

Generating Representations

One way of generating representations is to take some representations one knows and combine them in some fashion.There are three standard methods of obtaining new representations from old, namely, direct sum of representations, tensor products of representations, and dual representations.


Direct Sum of Representations

Let G be a matrix Lie group and let Π1, Π2, ..., Πm be representations of G acting on vector spaces V1, V2, ..., Vm. Then, the direct sum of  Π1, Π2, ..., Πm is a representation Π1 Π2 ... Πm of G acting on the space V1 V2 ... Vm, defined by [Π1 Π2 ... Πm (A)] (v1, v2, ..., vm) = (Π1(A)v1, Π2(A)v2, ..., Πm(A)vm) for all A G. 


Similarly, if g is a Lie algebra, and π1, π2, ..., πm are representations of g acting on V1, V2, ..., Vm, then we define the direct sum of  π1, π2, ..., πm , acting on  V1 V2 ... Vm by [π1 π2 ... πm (X)] (v1, v2, ..., vm) = (π1(X)v1, π2(X)v2, ..., πm(X)vm) for all X g.


Tensor Products of Representations

Consider an element u of U and v of V, the "product" of these two are denoted by uv. The space UV is then the space of linear combination of such products. The space of elements is in the form of a1u1v1 + a2u2v2 + ... + anunvn . The product is not necessary commutative (vu is in different space) but is bilinear ((u1+ au2)v = u1 v + au2 v, u (v1+ av2) = u v1 + au v2). 


If U and V are finite-dimensional real or complex vector spaces, then tensor product (W, ϕ) of U and V is a vector space W, together with a bilinear map ϕ : U x V --> W with the following property: If ψ is any bilinear map of UxV into a vector space X, then there exists a unique linear map ψ of W into X. Bilinear maps on UxV turn into linear maps on W. Suppose e1, e2 , en is a basis for U and f1, f2 , fm  is a basis for V. Then {ϕ(ei, fj) | 1 ≤ i ≤ n, 1≤ j ≤ m} is a basis for W. In this case, {eifj | 1 ≤ i ≤ n, 1≤ j ≤ m} is a basis for UV and dim (UV) = (dim U)(dim V). 


The tensor product (W, ϕ)  is unique up to canonical isomorphism. That is, if (W1, ϕ1) and (W2, ϕ2) are two tensor products, then there exists a unique vector space isomorphism Φ : W1 --> W2 . 


The defining property of UV is called the universal property of tensor products. Suppose that ψ(u,v) is some bilinear expression in (u,v). Then , the universal property says precisely that there is a unique linear map T(= ψ ) such that T(uv) = ψ(u,v). Let A : U --> U and B : V --> V be linear operators. Then, there exists a unique linear operator from UV to UV, denoted by AB, such that (AB)(uv) = (Au)(Bv) for all u U and v V. Moreover, (A1B1)(A2B2) = (A1A2)⊗(B1B2). 


There are two approaches to define tensor of representations. 

(A) Starts with a representation of a group G acting on a space V and a representation of another group H acting on space U and produces a representation of the product group GxH acting on space UV.


(B) Starts with two different representations of the same group G, acting on spaces U and V, and produces a representation of G acting on UV.


(A) Let G and H be matrix Lie groups. Let Π1 be a representation of G acting on a space U and let Π2 be a representation of G acting on a space V. Then, the tensor product of Π1 and Π2 is a representation Π1Π2 of GxH acting on UV defined by  Π1Π2 (A,B) =  Π1(A)Π2(B) for all A G and B H. Let  π1 π2 denote the associated representation of the Lie algebra of GxH, namely gh. Then, for all X g and Y h, π1 π2(X,Y) = π1(X) I + I π2(Y).  


(B) Let G be a matrix Lie groups and let Π1 and Π2 be representation of G acting on a space V1 and V2. Then, the tensor product of Π1 and Π2 is a representation of G acting on V1 V2 defined by Π1Π2 (A) =  Π1(A)Π2(A) for all A G. The associated Lie algebra g satisfies π1 π2(X) = π1(X) I  + I π2(X) all X g. 


Suppose Π1 and Π2 are irreducible representations of a group G. If regard Π1Π2 as representation of G, it may no longer be irreducible. If it is not irreducible, one can attempt to decompose it as a direct sum of irreducible representations. This process is called the Clebsch-Gordan theory. In the physics, the problem of analyzing tensor products of representations of SU(2) is called "addition of angular momentum."


Dual Representations

A linear functional on a vector space V is a linear map of V into C. If v1, v2, ..., vn is a basis for V, then for each set of constants a1, a2, ..., an, there is a unique linear functional ϕ such that ϕ(vk) = ak. If V is a finite-dimensional complex vector space, then the dual space to V, denoted by V*, is the set of all linear functionals on V. This is also a vector space and its dimension is the same as that of V.  If A is a linear operator on V, let Atr denote the dual or transpose operator on V*, (Atrϕ)(v) = ϕ(Av) for all ϕ V*, v  V. Note that the matrix of Atr is the transpose of the matrix A and not the conjugate transpose. If A and B are linear operators on V, then (AB)tr = Btr Atr . 


Suppose G is a matrix Lie group and Π is a representation of G acting on a finite-dimensional vector space V. Then, the dual representation Π* to Π is the representation of G acting on V* given by Π*(g) = [Π(g-1)]tr . Similarly, if π is a representation of a Lie algebra g acting on a finite-dimensional vector space V, then π* is the representation of g acting on V* given by  π*(g) = -π(X)tr. The dual representation is also called contragredient representation. Note that the transpose is an order-reversing operation, we cannot simply define Π*(g) = Π(g)tr. This would not be a representation. 

Sunday, November 9, 2008

Los Vascos 2006 Cabernet Sauvignon

Under the direct control of Domaines Barons de Rothschild (Lafite), Los Vascos, one of the oldest wine estates in Chile, is located in the Colchagua valley. The pre-phylloxra Bordeaux rootstock, Cabernet Sauvignon is the grape that made the estate famous.

Cabernet Sauvignon 2006 has a very red fruit nose and chocolate and bay leaf touches. It is fresh in the mouth, light, very fruity and balanced with good persistent tannins well blended into the wine. Highly concentrated with strawberry & cherries fruit notes and marked spices. International Wine Cellar 2006 Cebernet Sauvignon 88 point rating. Other years: 2004 (better), 2005.

Region: VALLE CENTRAL, Chile
Sub-Region: RAPEL
Grape Varieties: CABERNET SAUVIGNON

Friday, November 7, 2008

Representation Theory

Let G be a matrix Lie group. Then, a finite-dimensional complex representation of G is a Lie group homomorphism: Π : G --> GL(V), where V is a finite-dimensional complex vector space (with dim(V) ≥ 1). We may think of a representation as a linear action of a group on a vector space, i.e. to every g G, there is an operator Π(g) acts on the vector space. 


If g is a real or complex Lie algebra, then a finite-dimensional complex representation of g is a Lie algebra homomorphism π of g into gl(V). It is a unique representation π of g acting on the same space V such that Π(eX) = eπ(X) for all X g. The representation π can be computed as π(X) = d/dt  Π(etX) |t=0 and satisfies π(AXA-1) =  Π(A) π(X) Π(A)-1 for all X g  and all A G.  

If Π or π is a one-to-one homomorphism, then the representation is called faithful.


Let Π be a finite-dimensional real or complex representation of a matrix Lie group G, acting on a space V. A subspace W of V is called invariant if  Π(A)w W for all w W and all A G. An invariant subspace W is called nontrivial if W {0} and W V. A representation with no nontrivial invariant subspaces is called irreducible.


Let G be a connected matrix Lie group with Lie algebra g. Let Π be a representation of G and π the associated representation of g. Then, Π is irreducible if and only if π is irreducible.


Let G be a matrix Lie group, let H be a Hilbert space, and let U(H) denote the group of unitary operators on H. Then, a homomorphism Π : G --> U(H) is called a unitary representation of G if Π satisfies the following continuity condition (strong) condition: If An, A G and A --> An, then Π(An)v --> Π(A)v for all v H. A unitary representation with no nontrivial closed invariant subspace is called irreducible.  


The terms invariant, nontrivial, and irreducible are defined analogously for representations of Lie algebra.


Equivalence

Let G be a matrix Lie group, let  Π be a representation of G acting on the space V, and let ∑ be a representation of G acting on the space W. A linear map ϕ : V --> W is called an intertwining map of representation if ϕ(Π(A)v) = ∑(A)ϕ(v) for all A G and v V. The analogous property defines intertwining maps of representations of a Lie algebra. 

If ϕ is an intertwining map of representations and, in addition, ϕ is invertible, then ϕ is said to be an equivalence of representations. If there exists an isomorphism between V and W, then the representations are said to be equivalent. 


Let G be a connected matrix Lie group, let Π1 and Π2 be represenation of G, and let π1and π2 be the associated Lie algebra representations. Then, π1and π2 is equivalent if and only if Π1 and Π2 are equivalent.


Schur's Lemma

1. Let V and W be irreducible real or complex representations of a group or Lie algebra and let  ϕ : V --> W be an intertwining map. Then either ϕ = 0 or ϕ is an isomorphism.

2. Let V be an irreducible complex representation of a group or Lie algebra and let ϕ : V --> V be an intertwining map of V with itself. Then ϕ = λI, for some λ C.

3. Let V and W be irreducible complex representations of a group or Lie algebra and let ϕ1 , ϕ2  : V --> W , be nonzero intertwining maps. Then, ϕ1 = λϕ2, for some λ C. 


Applications of Representation Theory

Studying the representations of a group G (or of a Lie algebra) can give information about the group (or Lie algebra) itself. For example, if G is a finite group, then associated to G is something called the group algebra. The structure of this group algebra can be described very nicely in terms of the irreducible representations of G. 

One of the chief applications of representation theory is to exploit symmetry. If a system has symmetry, then the set of symmetries will form a group, and understanding the representations of the symmetry group allows one to use symmetry to simplify the problem. For example, if the equation has rotational symmetry, then the space of solutions will be invariant under rotations. Thus, the space of solutions will constitute a representation of the rotation group SO(3). If one knows what all of the representation of SO(3) are, this can help in narrowing down what the space of solutions can be. 

Wednesday, November 5, 2008

Universal Cover

Connectedness

A matrix Lie group G is said to be path-connected (in topology) if given any two matrices A and B in G, there exists a continuous path A(t), a ≤ t ≤ b, lying in G with A(a) = A and A(b) = B. A matrix Lie group is connected if and only if it is path-connected. 


A matrix Lie group G which is not connected can be decomposed (uniquely) as a union of several pieces, called components, such that two elements of the same component can be joined by a continuous path, but two elements of different components cannot.

Examples:
Group      Connected   Components
GL(n;C)      yes             1
SL(n;C)      yes             1
GL(n;R)      no              2
SL(n;R)      yes             1
O(n)           no               2
SO(n)         yes             1
U(n)           yes             1
SU(n)         yes             1
O(n;1)        no               4
SO(n;1)      no               2
Heisenberg yes            1
E(n)           no               2
P(n;1)        no               4


Simple Connectedness

A matrix Lie group G is said to be simply connected if it is connected and, in addition, every loop in G can be shrunk continuously to a point in G. More precisely, assume that G is connected. Then, G is simply connected if given any continuous path A(t), 0 ≤ t ≤ 1, lying in G with A(0) = A(1), there exists a continuous function A(s,t), 0 ≤ s,t ≤ 1, taking values in G and having the following properties:

(1) A(s,0) = A(s,1) for all s, is a loop condition;
(2) A(0,t) = A(t), when s=0 the loop is the specified loop A(t);
(3) A(1,t) = A(1,0) for all t, when s=1 the loop is a point.


Group versus Lie Algebra Homomorphisms

(A) Lie algebra homomorphism

Every Lie group homomorphism give rise to a Lie algebra homomorphism. Let G and H be matrix Lie groups, with Lie algebras g and h, respectively. Suppose that Φ : G --> H is a Lie group homomorphism. Then, there exists a unique real linear map ϕ :   g --> h such that  for all X g. The map ϕ has the following additional properties:

(1) ϕ (AXA-1) = Φ(A) ϕ(X) Φ(A)-1, for all X g, A G.

(2) ϕ([X,Y]) = [ ϕ(X),  ϕ(Y)], for all  X, Y g (Lie algebra homomorphism)

(3) ϕ(X) = d/dt Φ(etX) |t=0, for all X g


(B) Lie group homomorphism

Let G and H be matrix Lie groups with Lie algebras  g and h. Let ϕ :   g --> h be a Lie algebra homomorphism. If G is simply connected , then there exists a unique Lie group homomorphism Φ : G --> H such that  Φ(eX) = eϕ(X) for all X g. However, if G is not simply connected, this will not be true. We should look for Ğ that has the same Lie algebra as G but such that Ğ is simply connected. 


Baker-Campbell-Hausdorff formula says that if X and Y are sufficiently small, then 

log(eXeY) = X + Y + 1/2[X,Y] + 1/12[X,[X,Y]] - 1/12[Y,[X,Y]] + ...

Because ϕ is a Lie algebra homomorphism, 

ϕ(log(eXeY)) = ϕ(X) + ϕ(Y) + 1/2[ϕ(X),ϕ(Y)] + 1/12[ϕ(X),[ϕ(X),ϕ(Y)]] - 1/12[ϕ(Y),[ϕ(X),ϕ(Y)]] + ... = log(eϕ(X)eϕ(Y))

For group homomorphism, Φ(eXeY) = exp(ϕ(eXeY) = exp(ϕ(exp(log(eXeY) ))). From above, Φ(eXeY) = exp(log(eϕ(X)eϕ(Y))) = eϕ(X)eϕ(Y) = Φ(eX)Φ(eY). Therefore, by Baker-Campbell-Hausdorff formula, if X is small, Φ is a group homomorphism. 


Covering Groups

Let G be a connected Lie group. Then, a universal covering group (or universal cover) of G is a simply-connected Lie group H together with with a Lie group homomorphism Φ : H --> G such that the associated Lie algebra homomorphism ϕ :   h --> g is a Lie algebra homomorphism. The homomorphism Φ is called the covering homomorphism (or projection map). 


For any connected Lie group, a universal cover exists. If G is a connected Lie group and (H1, Φ1) and (H2, Φ2) are universal covers of G, then there exists a Lie group isomorphism Ψ : H1 --> H2 such that Φ2 o Ψ =  Φ1. 

The universal cover of a matrix Lie group may not be a matrix Lie group.