Section-A
(Very Short Answer Type Questions)
Note:- Answer all questions. As per the nature of the question you delimit your answer in one word, one sentence or maximum up to 30 words. Each question carries 1 (one) mark.
(i). If H and K are two sub groups of G such that G is an internal direct product of H and K . Then find H nn KH \cap K.
(ii). Write class equation for the finite group.
(iii). Define normal extension of a field.
(iv). Write Schwartz inequality.
Section – B
(Short Answer Questions)
Note :- Answer any two questions . Each answer should be given in 200 words. Each question carries 4 marks.
Prove that every Euclidean ring is a principal ideal domain.
If W is any subspace of a finite dimensional inner product space V , the prove that (W^(_|_))^(_|_)=W\left(W^{\perp}\right)^{\perp}=W.
Prove that every additive abelian group is a module over the ring Z of integers.
If KK is a field and d_(1),d_(2),dots.d_(n)d_1, d_2, \ldots . d_n are distinct automorphisms of KK. Then prove that it is impossible to find elements b_(1),b_(2),dots.b_(n)b_1, b_2, \ldots . b_n not all zero in KK such that b_(1)phi_(1)(a)+b_(2)phi_(2)(a)+dots dots dots+b_(n)phi _(n)(a)=0AA a in Kb_1 \phi_1(a)+b_2 \phi_2(a)+\ldots \ldots \ldots+b_n \phi_n(a)=0 \forall a \in K.
Section – C
(Long Answer Questions)
Note :- Answer any one question. Each answer should be given in 800 words. Each question carries 08 marks.
Let M_(1)M_1 and M_(2)M_2 are submodule of an R-module M . Then prove that:
Let V and V^(‘)V^{\prime} be any two finite dimensional vector space over the same field F . Then prove that the vector space Hom(V,V^(‘))\operatorname{Hom}\left(\mathrm{V}, V^{\prime}\right) of all linear transformation of V to V^(‘)V^{\prime} is also finite dimensional and dim Hom(V,V^(‘))=dimVxx dimV^(‘)\operatorname{Hom}\left(\mathrm{V}, V^{\prime}\right)=\operatorname{dim} \mathrm{V} \times \operatorname{dim} V^{\prime}.
Answer:
Question:-01(a)
If H and K are two subgroups of G such that G is an internal direct product of H and K. Then find H nn KH \cap K.
Answer:
If GG is the internal direct product of the subgroups HH and KK, the following conditions must hold:
Every element of GG can be uniquely written as a product hkh k where h in Hh \in H and k in Kk \in K.
H nn K={e}H \cap K = \{ e \}, where ee is the identity element of GG.
Thus, H nn K={e}H \cap K = \{ e \}.
Question:-01(b)
Write class equation for the finite group.
Answer:
The class equation of a finite group GG is a fundamental result in group theory that relates the structure of GG to its conjugacy classes. It is expressed as:
Z(G)Z(G): The center of GG, i.e., the set of elements that commute with every element of GG,
|Z(G)||Z(G)|: The size (order) of the center Z(G)Z(G),
g_(i)g_i: A representative of each conjugacy class outside the center,
C_(G)(g_(i))C_G(g_i): The centralizer of g_(i)g_i in GG, i.e., the set of elements in GG that commute with g_(i)g_i,
[G:C_(G)(g_(i))][G : C_G(g_i)]: The index of the centralizer C_(G)(g_(i))C_G(g_i) in GG, which equals the size of the conjugacy class containing g_(i)g_i.
Interpretation:
The elements in Z(G)Z(G) form singleton conjugacy classes (since they commute with all elements).
The remaining conjugacy classes contribute terms of the form [G:C_(G)(g_(i))][G : C_G(g_i)], representing their sizes.
The class equation is a powerful tool to analyze the structure of GG, especially in applications like Sylow theorems and proving group properties.
Question:-01(c)
Define normal extension of a field.
Answer:
A normal extension of a field is a type of field extension that satisfies a specific algebraic condition related to the splitting of polynomials. Formally:
Let EE be a field extension of FF (F sube EF \subseteq E). The extension E//FE/F is called a normal extension if:
Every irreducible polynomial f(x)in F[x]f(x) \in F[x] that has at least one root in EEsplits completely into linear factors over EE, i.e., all the roots of f(x)f(x) are contained in EE.
Equivalent Definitions:
E//FE/F is normal if EE is the splitting field of some family of polynomials in F[x]F[x].
E//FE/F is normal if it is closed under conjugation by automorphisms, i.e., for any field automorphism sigma\sigma of a larger field containing EE, if sigma(e)in E\sigma(e) \in E for some e in Ee \in E, then sigma(e)\sigma(e) is also in EE.
Examples:
The extension Q(sqrt2)//Q\mathbb{Q}(\sqrt{2})/\mathbb{Q} is not normal, because the irreducible polynomial x^(2)-2x^2 – 2 over Q\mathbb{Q} has a root sqrt2\sqrt{2} in Q(sqrt2)\mathbb{Q}(\sqrt{2}), but its other root, -sqrt2-\sqrt{2}, is not in Q(sqrt2)\mathbb{Q}(\sqrt{2}).
The extension Q(sqrt2,root(3)(3))//Q\mathbb{Q}(\sqrt{2}, \sqrt[3]{3})/\mathbb{Q} is normal, because it is the splitting field of (x^(2)-2)(x^(3)-3)(x^2 – 2)(x^3 – 3) over Q\mathbb{Q}.
Normal extensions are fundamental in the study of Galois theory, where they play a key role in defining Galois extensions.
Question:-01(d)
Write Schwartz inequality.
Answer:
The Schwarz inequality (or Cauchy-Schwarz inequality) is a fundamental result in mathematics, particularly in linear algebra and analysis. It states:
For any two vectors u,v\mathbf{u}, \mathbf{v} in an inner product space, the following inequality holds:
(:u,v:)\langle \mathbf{u}, \mathbf{v} \rangle: The inner product of u\mathbf{u} and v\mathbf{v},
||u||=sqrt((:u,u:))\|\mathbf{u}\| = \sqrt{\langle \mathbf{u}, \mathbf{u} \rangle}: The norm (magnitude) of u\mathbf{u},
||v||=sqrt((:v,v:))\|\mathbf{v}\| = \sqrt{\langle \mathbf{v}, \mathbf{v} \rangle}: The norm (magnitude) of v\mathbf{v}.
Equality Condition:
Equality holds if and only if u\mathbf{u} and v\mathbf{v} are linearly dependent, i.e., there exists a scalar lambda\lambda such that u=lambdav\mathbf{u} = \lambda \mathbf{v}.
Special Cases:
In Euclidean space R^(n)\mathbb{R}^n, with the standard dot product:|u*v| <= ||u||||v||,|\mathbf{u} \cdot \mathbf{v}| \leq \|\mathbf{u}\| \|\mathbf{v}\|,where ||u||=sqrt(sum_(i=1)^(n)u_(i)^(2))\|\mathbf{u}\| = \sqrt{\sum_{i=1}^n u_i^2}.
For functions f,gf, g in an inner product space L^(2)[a,b]L^2[a, b]:|int_(a)^(b)f(x)g(x)dx| <= sqrt(int_(a)^(b)f(x)^(2)dx)*sqrt(int_(a)^(b)g(x)^(2)dx).\left| \int_a^b f(x)g(x) \, dx \right| \leq \sqrt{\int_a^b f(x)^2 \, dx} \cdot \sqrt{\int_a^b g(x)^2 \, dx}.
The Schwarz inequality is widely used in diverse areas of mathematics, physics, and engineering to establish bounds and relationships between quantities.
Question:-02
Prove that every Euclidean ring is a principal ideal domain.
Answer:
Statement:
Every Euclidean ring is a principal ideal domain (PID).
Proof:
Definitions:
Euclidean Ring: A commutative ring RR with unity is a Euclidean ring if there exists a function delta:R\\{0}rarrN\delta: R \setminus \{0\} \to \mathbb{N} (called a Euclidean norm) such that for all a,b in Ra, b \in R with b!=0b \neq 0:
There exist q,r in Rq, r \in R such that a=bq+ra = bq + r, where either r=0r = 0 or delta(r) < delta(b)\delta(r) < \delta(b).
Principal Ideal Domain (PID): A commutative ring RR with unity is a PID if every ideal I sube RI \subseteq R is a principal ideal, i.e., there exists some a in Ra \in R such that I=(a)={ra:r in R}I = (a) = \{ ra : r \in R \}.
Proof:
Let RR be a Euclidean ring, and let II be a nonzero ideal of RR. We need to show that II is a principal ideal.
Since II is nonzero, there exists some b in Ib \in I with b!=0b \neq 0. Among all nonzero elements of II, choose an element d in Id \in I such that delta(d)\delta(d) is minimal (this is possible because delta\delta maps elements of R\\{0}R \setminus \{0\} to N\mathbb{N}, which is well-ordered).
We claim that I=(d)I = (d). To prove this, we need to show that every element a in Ia \in I can be written as a=rda = rd for some r in Rr \in R.
Since d in Id \in I, clearly (d)sube I(d) \subseteq I.
Let a in Ia \in I. By the division algorithm in RR, there exist q,r in Rq, r \in R such that:
a=qd+r,a = qd + r,
where either r=0r = 0 or delta(r) < delta(d)\delta(r) < \delta(d).
Since a,d in Ia, d \in I and II is an ideal, qd in Iqd \in I. Thus, r=a-qd in Ir = a – qd \in I.
If r!=0r \neq 0, then delta(r) < delta(d)\delta(r) < \delta(d), which contradicts the minimality of delta(d)\delta(d). Hence, r=0r = 0, and a=qda = qd.
Therefore, a in(d)a \in (d), and we conclude I sube(d)I \subseteq (d).
Combining (d)sube I(d) \subseteq I and I sube(d)I \subseteq (d), we get I=(d)I = (d), showing that II is a principal ideal.
Conclusion:
Since every ideal in RR is principal, RR is a principal ideal domain. Thus, every Euclidean ring is a PID. “Q.E.D.”\boxed{\text{Q.E.D.}}
Question:-03
If W is any subspace of a finite dimensional inner product space V, prove that (W^(_|_))^(_|_)=W\left(W^{\perp}\right)^{\perp}=W.
Answer:
Statement:
Let VV be a finite-dimensional inner product space, and let WW be a subspace of VV. Then:
(W^( _|_))^(_|_)=W.\left( W^\perp \right)^\perp = W.
Proof:
Definitions:
The orthogonal complementW^( _|_)W^\perp of a subspace W sube VW \subseteq V is defined as:
W^( _|_)={v in V:(:v,w:)=0″for all “w in W}.W^\perp = \{ v \in V : \langle v, w \rangle = 0 \ \text{for all } w \in W \}.
The space (W^( _|_))^(_|_)\left(W^\perp\right)^\perp is the orthogonal complement of W^( _|_)W^\perp, defined as:
(W^( _|_))^(_|_)={v in V:(:v,u:)=0″for all “u inW^( _|_)}.\left(W^\perp\right)^\perp = \{ v \in V : \langle v, u \rangle = 0 \ \text{for all } u \in W^\perp \}.
Key Observations:
Direct Sum Decomposition: In a finite-dimensional inner product space VV, any subspace W sube VW \subseteq V satisfies:
V=W o+W^( _|_),V = W \oplus W^\perp,
where:
W nnW^( _|_)={0}W \cap W^\perp = \{0\},
Every v in Vv \in V can be uniquely written as v=w+w^( _|_)v = w + w^\perp for w in Ww \in W and w^( _|_)inW^( _|_)w^\perp \in W^\perp.
Dimension Relation: The dimensions of WW, W^( _|_)W^\perp, and VV are related as:
Let v in(W^( _|_))^(_|_)v \in \left(W^\perp\right)^\perp. Then, vv satisfies (:v,u:)=0\langle v, u \rangle = 0 for all u inW^( _|_)u \in W^\perp.
By the direct sum decomposition V=W o+W^( _|_)V = W \oplus W^\perp, vv can be uniquely written as v=w+w^( _|_)v = w + w^\perp, where w in Ww \in W and w^( _|_)inW^( _|_)w^\perp \in W^\perp.
Since v in(W^( _|_))^(_|_)v \in \left(W^\perp\right)^\perp, we have:
because (:w,w^( _|_):)=0\langle w, w^\perp \rangle = 0 (orthogonality of WW and W^( _|_)W^\perp).
This implies ||w^( _|_)||^(2)=0\|w^\perp\|^2 = 0, so w^( _|_)=0w^\perp = 0. Hence, v=w in Wv = w \in W.
Therefore, (W^( _|_))^(_|_)sube W(W^\perp)^\perp \subseteq W.
Conclusion:
Combining W sube(W^( _|_))^(_|_)W \subseteq (W^\perp)^\perp and (W^( _|_))^(_|_)sube W(W^\perp)^\perp \subseteq W, we have:
(W^( _|_))^(_|_)=W.(W^\perp)^\perp = W.
“Q.E.D.”\boxed{\text{Q.E.D.}}
Question:-04
Prove that every additive abelian group is a module over the ring Z\mathbb{Z} of integers.
Answer:
Statement:
Every additive abelian group GG is a module over the ring Z\mathbb{Z} of integers.
Proof:
Definitions:
Additive Abelian Group: A set GG with a binary operation ++ is called an additive abelian group if:
GG is closed under addition,
Addition is associative: (a+b)+c=a+(b+c)(a + b) + c = a + (b + c) for all a,b,c in Ga, b, c \in G,
There exists an identity element 0in G0 \in G such that a+0=aa + 0 = a for all a in Ga \in G,
Every element a in Ga \in G has an additive inverse -a in G-a \in G such that a+(-a)=0a + (-a) = 0,
Addition is commutative: a+b=b+aa + b = b + a for all a,b in Ga, b \in G.
Module over Z\mathbb{Z}: A module over the ring Z\mathbb{Z} is a set GG with:
An additive abelian group structure,
A scalar multiplication *:Zxx G rarr G\cdot : \mathbb{Z} \times G \to G such that:
(m+n)*g=m*g+n*g(m + n) \cdot g = m \cdot g + n \cdot g for all m,n inZm, n \in \mathbb{Z}, g in Gg \in G,
m*(g+h)=m*g+m*hm \cdot (g + h) = m \cdot g + m \cdot h for all m inZm \in \mathbb{Z}, g,h in Gg, h \in G,
(m*n)*g=m*(n*g)(m \cdot n) \cdot g = m \cdot (n \cdot g) for all m,n inZm, n \in \mathbb{Z}, g in Gg \in G,
1*g=g1 \cdot g = g for all g in Gg \in G.
Proof Steps:
Define Scalar Multiplication:
Define scalar multiplication *:Zxx G rarr G\cdot : \mathbb{Z} \times G \to G as:
m*g=ubrace(g+g+cdots+gubrace)_(m” times”)quad”if “m > 0,m \cdot g = \underbrace{g + g + \dots + g}_{m \text{ times}} \quad \text{if } m > 0,
m*g=0quad”if “m=0,m \cdot g = 0 \quad \text{if } m = 0,
m*g=-(ubrace(g+g+cdots+gubrace)_(|m|” times”))quad”if “m < 0.m \cdot g = -(\underbrace{g + g + \dots + g}_{|m| \text{ times}}) \quad \text{if } m < 0.
This scalar multiplication is well-defined because GG is an additive abelian group, and we can repeatedly add or subtract elements.
Check Module Axioms:
We need to verify that the scalar multiplication satisfies the module axioms.
(1) Distributivity over Z\mathbb{Z}-addition:
(m+n)*g=ubrace(g+g+cdots+gubrace)_((m+n)” times”)=ubrace(g+g+cdots+gubrace)_(m” times”)+ubrace(g+g+cdots+gubrace)_(n” times”)=m*g+n*g.(m + n) \cdot g = \underbrace{g + g + \dots + g}_{(m + n) \text{ times}}
= \underbrace{g + g + \dots + g}_{m \text{ times}} + \underbrace{g + g + \dots + g}_{n \text{ times}}
= m \cdot g + n \cdot g.
(2) Distributivity over group addition:
m*(g+h)=ubrace((g+h)+(g+h)+cdots+(g+h)ubrace)_(m” times”)=ubrace(g+g+cdots+gubrace)_(m” times”)+ubrace(h+h+cdots+hubrace)_(m” times”)=m*g+m*h.m \cdot (g + h) = \underbrace{(g + h) + (g + h) + \dots + (g + h)}_{m \text{ times}}
= \underbrace{g + g + \dots + g}_{m \text{ times}} + \underbrace{h + h + \dots + h}_{m \text{ times}}
= m \cdot g + m \cdot h.
(3) Compatibility with multiplication in Z\mathbb{Z}:
(m*n)*g=ubrace(g+g+cdots+gubrace)_((m*n)” times”)=m*(ubrace(g+g+cdots+gubrace)_(n” times”))=m*(n*g).(m \cdot n) \cdot g = \underbrace{g + g + \dots + g}_{(m \cdot n) \text{ times}}
= m \cdot (\underbrace{g + g + \dots + g}_{n \text{ times}})
= m \cdot (n \cdot g).
(4) Action of 1:
1*g=ubrace(gubrace)_(1″ time”)=g.1 \cdot g = \underbrace{g}_{1 \text{ time}} = g.
Conclusion:
Since GG satisfies the axioms of a module over Z\mathbb{Z}, every additive abelian group is indeed a module over Z\mathbb{Z}.
“Q.E.D.”\boxed{\text{Q.E.D.}}
Question:-05
If KK is a field and d_(1),d_(2),dots,d_(n)d_1, d_2, \ldots, d_n are distinct automorphisms of KK. Then prove that it is impossible to find elements b_(1),b_(2),dots,b_(n)b_1, b_2, \ldots, b_n not all zero in KK such that
b_(1)phi_(1)(a)+b_(2)phi_(2)(a)+dots+b_(n)phi _(n)(a)=0quad AA a in K.b_1 \phi_1(a) + b_2 \phi_2(a) + \ldots + b_n \phi_n(a) = 0 \quad \forall a \in K.
Answer:
Statement:
Let KK be a field, and let phi_(1),phi_(2),dots,phi _(n)\phi_1, \phi_2, \ldots, \phi_n be distinct automorphisms of KK. Then there are no elements b_(1),b_(2),dots,b_(n)in Kb_1, b_2, \ldots, b_n \in K, not all zero, such that:
b_(1)phi_(1)(a)+b_(2)phi_(2)(a)+dots+b_(n)phi _(n)(a)=0quad”for all “a in K.b_1 \phi_1(a) + b_2 \phi_2(a) + \ldots + b_n \phi_n(a) = 0 \quad \text{for all } a \in K.
Proof:
1. Assume the contrary:
Suppose there exist elements b_(1),b_(2),dots,b_(n)in Kb_1, b_2, \ldots, b_n \in K, not all zero, such that:
b_(1)phi_(1)(a)+b_(2)phi_(2)(a)+dots+b_(n)phi _(n)(a)=0quad AA a in K.b_1 \phi_1(a) + b_2 \phi_2(a) + \ldots + b_n \phi_n(a) = 0 \quad \forall a \in K.
3. Evaluating L(a)=0L(a) = 0 for all a in Ka \in K:
Since L(a)=0L(a) = 0 for all a in Ka \in K, the coefficients b_(1),b_(2),dots,b_(n)b_1, b_2, \ldots, b_n must satisfy a specific linear independence property.
4. Distinct automorphisms imply independence:
The automorphisms phi_(1),phi_(2),dots,phi _(n)\phi_1, \phi_2, \ldots, \phi_n are distinct. If b_(1),b_(2),dots,b_(n)b_1, b_2, \ldots, b_n are not all zero, consider the set of functions phi_(1),phi_(2),dots,phi _(n)\phi_1, \phi_2, \ldots, \phi_n acting on KK. These functions are linearly independent over KK, meaning no non-trivial linear combination of these functions can yield the zero map.
In particular, if:
b_(1)phi_(1)(a)+b_(2)phi_(2)(a)+dots+b_(n)phi _(n)(a)=0quad AA a in K,b_1 \phi_1(a) + b_2 \phi_2(a) + \ldots + b_n \phi_n(a) = 0 \quad \forall a \in K,
then b_(1)=b_(2)=dots=b_(n)=0b_1 = b_2 = \ldots = b_n = 0, contradicting the assumption that b_(1),b_(2),dots,b_(n)b_1, b_2, \ldots, b_n are not all zero.
5. Conclusion:
Thus, it is impossible to find b_(1),b_(2),dots,b_(n)in Kb_1, b_2, \ldots, b_n \in K, not all zero, such that:
b_(1)phi_(1)(a)+b_(2)phi_(2)(a)+dots+b_(n)phi _(n)(a)=0quad AA a in K.b_1 \phi_1(a) + b_2 \phi_2(a) + \ldots + b_n \phi_n(a) = 0 \quad \forall a \in K.
“Q.E.D.”\boxed{\text{Q.E.D.}}
Question:-06
Let M_(1)M_1 and M_(2)M_2 be submodules of an RR-module MM. Then prove that:
varphi(m_(1))\varphi(m_1) represents the coset m_(1)+M_(2)m_1 + M_2, which is a member of the quotient module (M_(1)+M_(2))/(M_(2))\frac{M_1 + M_2}{M_2}.
Step 2: Verify varphi\varphi is Well-Defined
Since m_(1)inM_(1)subeM_(1)+M_(2)m_1 \in M_1 \subseteq M_1 + M_2, the coset m_(1)+M_(2)m_1 + M_2 is a valid element of (M_(1)+M_(2))/(M_(2))\frac{M_1 + M_2}{M_2}. Thus, varphi\varphi is well-defined.
Step 3: Check Linearity of varphi\varphi
Let m_(1),m_(1)^(‘)inM_(1)m_1, m_1′ \in M_1 and r in Rr \in R. Then:
By definition of M_(1)+M_(2)M_1 + M_2, every element of M_(1)+M_(2)M_1 + M_2 can be written as m_(1)+m_(2)m_1 + m_2 for some m_(1)inM_(1)m_1 \in M_1 and m_(2)inM_(2)m_2 \in M_2. Hence, the coset m_(1)+M_(2)m_1 + M_2 represents all elements in (M_(1)+M_(2))/(M_(2))\frac{M_1 + M_2}{M_2}. Therefore:
Let VV and V^(‘)V^{\prime} be any two finite dimensional vector spaces over the same field FF. Then prove that the vector space Hom(V,V^(‘))\operatorname{Hom}(V, V^{\prime}) of all linear transformations from VV to V^(‘)V^{\prime} is also finite dimensional and
dim Hom(V,V^(‘))=dim V xx dim V^(‘).\dim \operatorname{Hom}(V, V^{\prime}) = \dim V \times \dim V^{\prime}.
Answer:
Statement:
Let VV and V^(‘)V’ be finite-dimensional vector spaces over the same field FF, with dim V=n\dim V = n and dim V^(‘)=m\dim V’ = m. Then:
The vector space Hom(V,V^(‘))\operatorname{Hom}(V, V’), consisting of all linear transformations from VV to V^(‘)V’, is finite-dimensional.
Its dimension is given by:dim Hom(V,V^(‘))=dim V xx dim V^(‘)=n xx m.\dim \operatorname{Hom}(V, V’) = \dim V \times \dim V’ = n \times m.
The set Hom(V,V^(‘))\operatorname{Hom}(V, V’) is the collection of all linear transformations T:V rarrV^(‘)T: V \to V’.
For any T in Hom(V,V^(‘))T \in \operatorname{Hom}(V, V’), the action of TT is completely determined by how it maps basis elements of VV. This is because TT is linear, and any vector v in Vv \in V can be written as a linear combination of basis elements.
Step 2: Basis of VV and V^(‘)V’:
Let {v_(1),v_(2),dots,v_(n)}\{v_1, v_2, \ldots, v_n\} be a basis of VV, so every v in Vv \in V can be written uniquely as:
Each column of AA corresponds to the image of a basis element v_(i)v_i of VV under TT.
Step 5: Vector Space Structure of Hom(V,V^(‘))\operatorname{Hom}(V, V’):
The set Hom(V,V^(‘))\operatorname{Hom}(V, V’) has a natural structure as a vector space:
Addition of linear transformations: For T_(1),T_(2)in Hom(V,V^(‘))T_1, T_2 \in \operatorname{Hom}(V, V’) and v in Vv \in V:(T_(1)+T_(2))(v)=T_(1)(v)+T_(2)(v).(T_1 + T_2)(v) = T_1(v) + T_2(v).
Scalar multiplication: For c in Fc \in F and T in Hom(V,V^(‘))T \in \operatorname{Hom}(V, V’):(cT)(v)=c*T(v).(cT)(v) = c \cdot T(v).
The space Hom(V,V^(‘))\operatorname{Hom}(V, V’) is isomorphic to the space of m xx nm \times n matrices over FF.
Step 6: Dimension of Hom(V,V^(‘))\operatorname{Hom}(V, V’):
The space of m xx nm \times n matrices over FF has dimension m*nm \cdot n, as there are m xx nm \times n independent entries in the matrix.
Each entry of the matrix corresponds to a coefficient a_(ij)in Fa_{ij} \in F, and choosing these coefficients independently determines a unique linear transformation TT.
Thus:
dim Hom(V,V^(‘))=m*n=dim V^(‘)*dim V.\dim \operatorname{Hom}(V, V’) = m \cdot n = \dim V’ \cdot \dim V.
Conclusion:
The vector space Hom(V,V^(‘))\operatorname{Hom}(V, V’) is finite-dimensional, and its dimension is:
dim Hom(V,V^(‘))=dim V xx dim V^(‘)=n xx m.\boxed{\dim \operatorname{Hom}(V, V’) = \dim V \times \dim V’ = n \times m.}