Ross 1.0
Ross 1.0
Ross 1.0
Here is our official definition of “field” (from Appendix C of Friedberg, Insel & Spence).
Definition. A field is a set F on which two operations + and · (called addition and multiplica-
tion) are defined (for all possible pairs in F, and always producing an element of F), and having
two privileged elements 0 and 1 with 0 ̸= 1, such that the following laws hold: for all a, b, c ∈ F,
(F 1) a+b=b+a
a·b=b·a
(F 2) a + (b + c) = (a + b) + c
a · (b · c) = (a · b) · c
(F 3) 0+a=a
1 · a = a.
(F 4) For every a ∈ F there exists x ∈ F with a + x = 0.
For every a ∈ F with a ̸= 0 there exists y ∈ F with a · y = 1.
(F 5) a · (b + c) = (a · b) + (a · c)
In any field F, we define −a to be the unique solution x to a + x = 0. Note that (F 4) guarantees
the existence of a solution, and uniqueness can be deduced from (F 1)–(F 3).
In any field F, if a ∈ F and a ̸= 0, then we define a−1 to be the unique solution x to ax = 1.
Note that (F 4) guarantees the existence of a solution, and uniqueness can again be deduced from
(F 1)–(F 3).
2
MATH 146 January 11 Lecture notes
Announcements:
P(F) with addition and scalar multiplication defined term-wise is a vector space over F.
Remark. A more standard notation for P(F) is F[x], but we will use P(F) in this course.
Next: some basic facts true of all vector spaces.
Theorem 1.1 (Cancellation Law). Let V be a vector space. If x, y, u ∈ V and x + u = y + u, then x = y.
Proof. Assume x + u = y + u. Let 0 be the zero vector of V (given by (VS 3). By (VS 4), there exists
z ∈ V with u + z = 0. Then
x = x+0 (VS 3)
= x + (u + z) by choice of z above
= (x + u) + z (VS 2)
= (y + u) + z assumption
= y + (u + z) (VS 2)
= y+0 choice of z above
= y (VS 3).
□
Corollary 1. Suppose V is a vector space. There is only one vector in V that can be the zero vector.
Proof. Suppose 01 , 02 ∈ V both satisfy (VS 3). Thus x + 01 = x + 02 = x for all x ∈ V . Use (VS 1) to flip
this to get 01 + x = 02 + x. Then 01 = 02 by the Cancellation Law. □
Corollary 2. Suppose V is a vector space and x ∈ V . There is only one vector u ∈ V satisfying x + u = 0.
Proof. Like the proof of Corollary 1; exercise. □
Definition. Let V be a vector space and x, y ∈ V .
(1) −x denotes the unique vector u ∈ V satisfying x + u = 0.
(2) x − y denotes x + (−y).
Theorem 1.2. Suppose V is a vector space over F, x ∈ V , and a ∈ F.
(1) 0x = 0.
(2) (−a)x = −(ax) = a(−x).
(3) a0 = 0.
Remark. Parts (1) and (2) are proved in the text. Parts (1) and (3) are also proved in Prof. Tatarko’s
notes.
Also note the overloaded notation in the statement. For example, in part (1), the first 0 is the scalar
0 ∈ F, while the second 0 is the zero vector in V . In part (2), the first − means minus in F, while the
second and third mean the minus in V defined above.
Here is a proof of (1), where I will use bold font for vectors, nonbold for scalars: In the field F we have
0 + 0 = 0; thus 0x = (0 + 0)x = 0x + 0x by (VS 8). On the other hand, we have 0x = 0x + 0 = 0 + 0x by
(VS 3) and (VS 1). Putting these together gives 0x + 0x = 0 + 0x. Now apply the Cancellation Law to
get 0x = 0.
2
MATH 146 January 13 Section 2
Assume V is a vector space over F and x1 , . . . , xn ∈ V . Because of (VS 2), we can (and do) write
expressions like x1 + x2 + · · · + xn without declaring where the brackets go. And we could (though we
won’t) prove true facts like
x1 + x2 + · · · + xn = xσ(1) + xσ(2) + · · · + xσ(n) for any permutation σ of {1, 2, . . . , n}
c(x1 + x2 + · · · + xn ) = cx1 + cx2 + · · · + cxn for any scalar c ∈ F
−(x1 + x2 + · · · + xn ) = −x1 − x2 − · · · − xn
Going forward, you can freely use these facts, except when told not to.
(VS 4): given x ∈ W , recall that −1 ∈ F so (−1)x ∈ W . Applying Theorem 1.2(2) and (VS 5) to V , we
get (−1)x = −(1x) = −x (the additive inverse of x in V ), which proves −x ∈ W . We now show that −x
is also an additive inverse of x in W :
x + (−x) = x +V (−x) = 0V by definition of −x. □
Incidentally, the proof shows that if W is a subspace of V , then:
• W always contains the zero vector of V (which is also the zero vector of W );
• W is closed under additive inverses (calculated in V ). Moreover, if x ∈ W then the additive inverse
of x calculated in V is the same as the additive inverse of x calculated in the vector space W .
Example.
(1) V = R3 as a vector space over R (usual operations). Let W be the following plane in R3 :
W = {(x, y, z) ∈ R3 : x + y + z = 0}.
It’s easy to check that W is a subspace of R3 : i.e., it is a subset of R3 , it is closed under addition
and scalar multiplication, and it is nonempty.
(2) Same V . Let W1 = {(x, y, z) ∈ R3 : x + y + z = 1}. W1 is not a subspace of R3 , for many reasons:
(i) it doesn’t contain the zero element of R3 , so it can’t be a subspace by the observation following
the proof of the Lemma; (ii) W isn’t closed under scalar multiplication; and (iii) W isn’t closed
under vector addition.
(3) Let W2 be the paraboloid
W2 = {(x, y, z) ∈ R3 : z = x2 + y 2 }.
W2 is not a subspace of R3 , as it is not closed under scalar multiplication. (It’s also not closed
under vector addition.)
(4) Given any vector space V :
• V is a subspace of itself.
• {0} is a subspace.
(5) The subspaces of R3 are R3 , all planes through 0, all lines through 0, and {0}.
(6) Let V be C considered as a vector space over R. (So scalar multiplication is R × C → C.)
(a) R is a subspace of V . (Proof: R ̸= ∅, R is closed under (complex) addition, and R is closed
under scaling by real numbers.)
(b) Can you find another subspace of V (other than C and {0})?
(7) Is R2 a subspace of R3 ?
(8) For each field F and n ≥ 0, Pn (F) is a subspace of Pn+1 (F) and is also a subspace of P(F).
2
MATH 146 January 16 Section 2
Theorem 1.4. Let V be a vector space over F. Let W1 , W1 be subspaces of V . Then W1 ∩ W2 is also a
subspace of V .
Proof. We check the definition of being a subspace.
(1) Let x, y ∈ W1 ∩ W2 . Then x, y ∈ W1 so x + y ∈ W1 . Similarly, x, y ∈ W2 so x + y ∈ W2 . Thus
x + y ∈ W1 ∩ W2 , proving W1 ∩ W2 is closed under addition.
(2) Let x ∈ W1 ∩ W2 and c ∈ F. A similar proof shows cx ∈ W1 ∩ W2 .
(3) We have 0 ∈ W1 and 0 ∈ W2 by the earlier Lemma, so 0 ∈ W1 ∩ W2 , proving W1 ∩ W2 ̸= ∅. □
Note: in general, it is not true that the union of two subspaces is a subspace. See Assignment 1.
§1.4
2
MATH 146 January 18 Section 2
Step 5: Neither occurrence of d can be eliminated (without re-introducing a, b or c). This signals that we
are done with the elimination process. Since we didn’t find a contradiction (such as 0 = 1), this system
has solutions. Since solutions exist, this means that the answer to our original question is YES:
2 1 1 0 1 1 0 1 1 2
is in span , , , .
−1 1 −1 0 0 −1 1 1 1 0
To actually see how to express the first matrix as a linear combination of the other four, we first describe
all solutions to the system (∗). Introducing a new symbol, or parameter, for d, say d = t, we can express
2
a, b, c, d in terms of t using the equations in (∗∗), obtaining
a = 2
b = − t
t ∈ R.
c = 1 − t
d = t
As t ranges over R, the equations above describe all solutions to (∗). For example, when t = 0 we get the
solution (a, b, c, d) = (2, 0, 1, 0). Thus one way to describe the original matrix as a sum of the other four is
“2 times the first matrix plus the third matrix,” i.e.,
2 1 1 0 1 1 0 1 1 2
=2· +0· +1· +0· .
−1 1 −1 0 0 −1 1 1 1 0
More information about the method of elimination will be placed on Learn.
3
MATH 146 January 20 Section 2
§1.5
Jargon. Let V be a vector space over F. Given a linear combination
a1 v1 + · · · + an vn
of vectors from V , we say call it trivial if a1 = a2 = · · · = an = 0, and nontrivial otherwise.
Definition. Let V be a vector space over a field F. Let S ⊆ V .
(1) Say S is linearly dependent if there exists a nontrivial linear combination of distinct vectors from
S which equals 0; i.e., if it is possible to write
(∗) a1 v1 + a2 v2 + · · · + an vn = 0
for some distinct vectors v1 , . . . , vn ∈ S , and such that at least one ai is ̸= 0 .
(2) Say S is linearly independent if S is not linearly dependent: i.e., if
v1 , . . . , vn ∈ S
v1 , . . . , vn distinct =⇒ a1 = a2 = · · · = an = 0.
a1 v1 + · · · + an vn = 0
Example. (1) In R2 , the set {(1, −1), (−2, 2), (3, 4)} is linearly dependent because we can write
1 −2 3 0
2 +1 +0 = .
−1 2 4 0
(2) In F3 , is the set {(1,
1, 0), (1,
0,1), (0,1,
1)} linearly
dependent or independent?
1 1 0 0
Solution: solve a 1 + b 0 + c 1 = 0. We get the system
0 1 1 0
a + b = 0
a + c = 0
b + c = 0.
If F = R (or C or Q), we can solve to get a = b = c = 0, meaning the set is linearly independent.
But if F = Z2 , we get the solution a = b = c = t (t ∈ Z2 ). In fact, the sum of the three vectors in
(Z2 )3 is the zero vector, so in (Z2 )3 these vectors are linearly dependent.
(3) In any vector space, {0} is linearly dependent. Any set containing 0 is linearly dependent.
(4) {v} is linearly dependent iff v = 0. (Use A1 Problem 3 for ⇒.)
(5) What about ∅? Is it linearly dependent, or independent? (Answer: independent)
(6) Suppose S ⊆ T ⊆ V . Which of the following implications are correct?
?
S is linearly dependent =⇒ T is linearly dependent. (True)
?
T is linearly dependent =⇒ S is linearly dependent. (False)
?
S is linearly independent =⇒ T is linearly independent. (False)
?
T is linearly independent =⇒ S is linearly independent. (True)
Proposition. Let V be a vector space over F. Let S ⊆ V . TFAE:
(1) S is linearly dependent.
(2) S = {0} or ∃v ∈ S such that v can be expressed as a linear combination of other vectors in S.
Proof. (⇐) We have already explained why {0} is linearly dependent. Suppose the vector v ∈ S can be
written as a linear combination of other vectors u1 , . . . , un ∈ S, say
v = c1 u1 + · · · + cn un , c1 , . . . , cn ∈ F.
We can assume that u1 , . . . , un are distinct. Since v ̸∈ {u1 , . . . , un } by assumption, it follows that
u1 , . . . , un , v are distinct. Now note that (−1)v = −v by Theorem 1.2(3) and so
c1 u1 + · · · + cn un + (−1)v = 0.
As u1 , . . . , un , v are distinct vectors in S and at least one of the coefficients (−1) in the above linear
combination is not 0, we get that S is linearly dependent.
(⇒) Assume S is linearly dependent. By assumption there exist distinct u1 , . . . , un ∈ S and a1 , . . . , an ∈
F, not all 0, such that
(∗) a1 u1 + · · · + an un = 0.
By “weeding out” terms where ai = 0, we can assume that ai ̸= 0 for all i = 1, . . . , n.
Case 1: n = 1. Then a1 u1 = 0 with a1 ̸= 0. Thus u1 = 0 by A1 Problem 3, so {0} ⊆ S. One option is
S = {0}; the other is that there exists v ∈ S with v =
̸ 0. In the second option, we can write 0 ∈ S as a
linear combination of v ∈ S, namely, 0 = 0v.
Case 2: n > 1.
In this case we can show that every ui can be written as a linear combination of the other uj ’s. For
example, here is the proof that u1 can be written as a linear combination of u2 , . . . , un : rewite (∗) as
a1 u1 = (−a2 )u2 + · · · + (−an )un .
Recall that a1 ̸= 0, so a−1
1 exists (in F). Multiply both sides of the equation to get
u1 = (−a2 /a1 )u2 + · · · + (−an /a1 )un
proving u1 ∈ S can be written as a linear combination of u2 , . . . , un ∈ S. □
Comment. Another equivalent condition to “S is linearly dependent” is
∃v ∈ S with v ∈ span(S \ {v}).
Indeed, if v ∈ S and v = a1 u1 + · · · + an un with u1 , . . . , un ∈ S and ui ̸= v for all i, then v ∈ span(S \ {v}).
If S = {0} then 0 ∈ span(∅) = span(S \ {0}).
Taking the negation, we get
Corollary. A set S in a vector space is linearly independent iff ∀v ∈ S we have v ̸∈ span(S \ {v}).
2
MATH 146 January 23 Section 2
Definition. Let V be a vector space over F. A subset S ⊆ V is a basis for V if S is linearly independent
and spans V .
Example. (1) V = Rn : a basis is En = {(1, 0, 0, . . .), (0, 1, 0, . . .), . . . , (0, . . . , 0, 1)}.
| {z } | {z } | {z }
e1 e2 en
(2) More generally, En is a basis for Fn (for any F).
(3) V = Mm×n (F): a basis is Em,n = {Eij : 1 ≤ i ≤ m, 1 ≤ j ≤ n} where Eij is the matrix of all 0’s
except one 1 in the (i, j) spot.
(4) V = P(F): a basis is B = {1, x, x2 , . . .}.
Proof sketch of (4). Show span(B) = P(F): ⊆ is obvious. For ⊇, let f ∈ P(F), say
f = a0 + a1 x + a2 x2 + · · · + an xn
= a0 ·1 + a1 x + · · · + an xn .
So f is a linear combination of 1, x, x2 , . . . , xn , all in B; so f ∈ span(B). This proves P(F) ⊆ span(B).
Prove B is linearly independent: assume xi1 , . . . , xik are distinct vectors in B, ai1 , . . . , aik ∈ F, and
ai1 xi1 + · · · + aik xik = 0.
By definition, a polynomial is the zero polynomial iff all of its coefficients are 0, so
ai 1 = · · · = ai k = 0
proving B is linearly independent. So B is a basis. □
(5) Here is another basis for P(R), invented by Legendre: {P0 , P1 , . . . , Pn , . . .} where
n k
X n n+k x−1
Pn = .
k=0
k k 2
For example, P0 = 1, P1 = x, P2 = 12 (3x2 − 1), . . . This basis for P(R) is “better” than B in some
applications.
Theorem 1.8. Suppose B = {v1 , . . . , vn } ⊆ V is finite. TFAE:
(1) B is a basis for V .
(2) Every v ∈ V admits a unique representation v = a1 v1 + · · · + an vn as a linear combination of B.
Proof sketch. (1) ⇒ (2). Since B spans V , every vector in V admits some representation as a linear
combination of B. To prove uniquencess, assume that v ∈ V admits two representations
v = a1 v1 + · · · + an vn
and v = b1 v1 + · · · + bn vn .
Subtract the 2nd from the first to get
0 = (a1 − b1 )v1 + · · · + (an − bn )vn .
Use linear independence of B to get
a1 − b1 = · · · = an − bn = 0, i.e., ai = bi for all i = 1, . . . , n.
(2) ⇒ (1). By assumption, every v ∈ V admits a representation as a linear combination of B, so
span(B) = V . To prove B is linear independent, note that one representation of the zero vector (as a
linear combination of B) is
0 = 0v1 + · · · + 0vn .
By the uniqueness assumption, this is the only representation of the zero vector. That is,
a1 v1 + · · · + an vn = 0 =⇒ a1 = · · · = an = 0,
proving B is linear independent. □
Corollary. Let V be a vector space/F. If V has a finite basis B = {v1 , . . . , vn }, then V is “naturally” in
1-1 correspondence with Fn via
a1 v1 + · · · + an vn ↭ (a1 , . . . , an )
| {z } | {z }
∈V ∈Fn
Existence of bases
Theorem 1.9. Suppose the vector space V is spanned by a finite set S. Then S can be “shrunk” to a
basis B of V (i.e., ∃ basis B ⊆ S).
Proof sketch. We have span(S) = V . If S is linearly independent, then S itself is a basis. Otherwise,
∃v ∈ S such that v ∈ span(S \ {v}), and hence span(S \ {v}) = span(S) = V . Choose such v and let
S ′ = S \ {v}. It still spans V but is smaller. Repeat until you can’t anymore; the final S (k) must be linearly
independent and a basis. □
Question: is the theorem true for infinite spanning sets? More care would be needed in the proof. For
example, it is possible to construct a vector space V ̸= {0} and an infinite spanning set {v0 , v1 , . . .} with
the property that for every k, vk ∈ span({vk+1 , vk+2 , . . .}). So you might first throw away v0 , then v1 , then
v2 , etc. and at the end you will have reduced S to ∅ (which doesn’t span V ).1
Question. Suppose V has a finite linearly independent set S. Can S always be “grown” (i.e., extended)
to a basis for V ?
Naive algorithm proving “YES.”
Assume S ⊆ V is linearly independent. If span(S) = V , then S is already a basis. Otherwise, there
exists x ∈ V with x 6∈ span(S). Let S 0 = S ∪ {x}. I want to say that S 0 is also linearly independent. (I’ll
justify this later.)
Repeat: if span(S 0 ) = V then S 0 is a basis. Otherwise, there exists x0 ∈ V with x0 6∈ span(S 0 ). Let
S 00 = S 0 ∪ {x0 }. Then S 00 is still linearly independent.
Continue: we get S ⊂ S 0 ⊂ S 00 ⊂ · · · ⊂ S (n) ⊂ · · · with each set S (k) linearly independent. Now there is
the question of whether this process terminates. It turns out that we can prove termination, provided we
assume that V has a finite spanning set.
To fully justify the claims above, we need the following two theorems.
Theorem 1.7. Suppose V is a vector space, S ⊆ V , and x ∈ V with x 6∈ S. Then
S ∪ {x} is linearly independent ⇐⇒ (S is linearly independent and x 6∈ span(S)).
Proof sketch. (⇒) By known facts.
(⇐) Assume S is linearly independent and x 6∈ span(S). Suppose S ∪ {x} is linearly dependent. So
there exist v1 , . . . , vn ∈ S ∪ {x} (distinct) and scalars a1 , . . . , an ∈ F, not all 0, with
(∗) a1 v1 + · · · + an vn = 0.
By weeding out those ai vi for which ai = 0, we can assume that we have ai 6= 0 for all i.
We must have x ∈ {v1 , . . . , vn } (why?) WLOG assume that vn = x. Since an 6= 0, we can “solve for x”
in (∗) to get x ∈ span(v1 , . . . , vn−1 ) ⊆ span(S), contradiction.
Theorem 1.10 (Baby Replacement Theorem1). Suppose V is spanned by some finite set of size n. Then
every linearly independent set S ⊆ V satisfies |S| ≤ n.
(Proof deferred until Friday).
Now consider again the Question and the naive algorithm for solving it. Assume V is spanned by a
finite set of size n, and S is linearly independent. Recall the sequence of growing linearly independent sets
S ⊂ S 0 ⊂ S 00 ⊂ · · ·
By the Baby Replacement Theorem, all of these sets must have size ≤ n, so the sequence cannot go on
forever. Therefore at some step, say S (k) , the algorithm terminates, which must be because S (k) is a basis.
This proves:
Corollary 2 (c). Suppose V is spanned by a finite set. Then every linearly independent set S ⊆ V can be
extended to a basis for V .
Here is another application of the Baby Replacement Theorem.
Corollary 1. Assume V is spanned by a finite set.
(1) V has a finite basis B.
(2) Let n = |B|. Every basis of V has size n.
1The full Replacement Theorem says that, if T is the spanning set of size n, then not only do we have |S| ≤ n, but also
there exists a subset H ⊆ T of size |H| = n − |S|, such that S ∪ H spans V .
Proof. (1) Let S be a finite spanning set. S can be shrunk to a basis B, by Theorem 1.9 (Jan 23). Obviously
B is also finite.
(2) Let C be another basis. Apply the Baby Replacement Theorem 1.10 using B as the finite spanning
set and S := C as the linearly independent set to get |C| ≤ |B|. So C is also a finite basis. So we can
reverse the roles of B and C in the above argument to get |B| ≤ |C|. So |C| = |B| = n.
2
MATH 146 January 27 Section 2
Recap: suppose V has a finite spanning set.
• Theorem 1.9: finite spanning sets shrink to bases.
• (Baby) Replacement Theorem: finite spanning set B, S lin. indep. =⇒ |S| ≤ |B|.
• Corollary 2(c): lin. indep. sets extend to bases.
• Corollary 1(1): V has a basis.
• Corollary 1(2): all bases have same (finite) size.
Still need to prove the (Baby) Replacement Theorem. First we prove:
Exchange Lemma∗ . Suppose V is a vector space, span(x1 , . . . , xk , v1 , . . . , v` ) = V (possibly with k = 0),
and x ∈ V with x 6= span(x1 , . . . , xk ). Then there exists i ∈ {1, . . . , `} such that
span(x1 , . . . , xk , x, v1 , . . . , vi−1 , vi+1 , . . . , v` ) = V.
Repeating, we have x2 6∈ span(x1 ) (why?). Again by the Exchange Lemma, there exists j ∈ {2, . . . , n}
such that (B 0 \ {vj }) ∪ {x2 } still spans V . For concreteness, assume it is j = 2, so
span(x1 , x2 , v3 , . . . , vn ) = V.
Keep repeating until we run out of v’s. Then we will have
span(x1 , x2 , . . . , xn ) = V.
In particular, xn+1 ∈ span(x1 , . . . , xn ), contradicting that S is linearly independent.
Corollary 1 justifies the next definition.
Definition. Let V be a vector space.
(1) Suppose V is spanned by a finite set. The dimension of V , written dim V , is the unique size n of
every basis of V . We also say that V is finite-dimensional.
(2) If V has no finite spanning set, then we say V is infinite-dimensional.
Example.
(1) Fn . It has the finite basis En of size n. So dim Fn = n.
(2) Mm×n (F). It has the finite basis Em,n of size mn. So dim Mm×n (F) = mn.
(3) P(F). It has the infinite basis {1, x, x2 , . . .} (which is linearly independent). By the Replacement
Theorem, P(F) has no finite spanning set, i.e., is infinite-dimensional.
(4) F(R, R). It has an infinite linearly independent subset (A2). So by the Replacement Theorem, it
has no finite spanning set. So it is infinite-dimensional.
Here is an intuitively obvious but important consequence of our results.
Theorem 1.11. Let V be a finite-dimensional vector space. Let W be a subspace. Then:
(1) W is finite-dimensional and dim W ≤ dim V .
(2) If W is a proper subspace (i.e., W ⊂ V ), then dim W < dim V .
Proof. Let n = dim V . Fix a subspace W . By the Replacement Theorem or Corollary 2(c), every lin.
indep. set in V has size ≤ n. In particular, every lin. indep. set in W has size ≤ n. This means that if we
start with some lin. indep. subset S ⊆ W (say S = ∅), the Naive Algorithm for extending S to a basis
for W will terminate, say at C. Then |C| = dim W (by definition). As C is linearly independent (in W ,
and hence in V ), we have |C| ≤ n (see the comments at the start of this proof). Thus dim W ≤ n, proving
(1).
(2) Let C be the basis for W we found in (1). C is linearly independent in V , so we can extend C to a
basis B of V (by Corollary 2(c)). We have C ⊆ B, |C| = dim W ≤ n, and |B| = dim V = n. If C = B, then
we would have W = span(C) = span(B) = V , contradiction. So C ⊂ B, so |C| < |B|, proving (2).
Here is another cute result that is sometimes useful.
Corollary 2 (continued). Assume V is a finite-dimensional vector space with dim V = n. Suppose S ⊆ V
is a subset with |S| = n. Then
(1)
S is linearly independent ⇐⇒ S spans V ⇐⇒ S is a basis for V .
(1)
Proof. (⇒) By Corollary 2(c), S can be extended to a basis C. Then |C| = n. But S ⊆ C and |S| = n. So
S = C and S is already a basis for V .
(1)
(⇐) By Theorem 1.9 (Jan 23), we can shrink S to a basis C. Then |C| = n. But C ⊆ S and |S| = n. So
S = C and S is already a basis for V .
2
MATH 146 January 30 Section 2
The material in this week’s lectures is not in our textbooks (only in the exercises of Friedberg-Insel-Spence).
Example. V = M2×2 (R),
a b 0 b
W1 = | a, b, c ∈ R , W2 = | b, d ∈ R .
c 0 b d
Warm-up: what is dim W1 ? dim W2 ? W1 ∩ W2 ? dim(W1 ∩ W2 )?
Definition. Suppose V is a vector space and W1 , W2 are subspaces. Their sum is the set
def
W1 + W2 = {w1 + w2 : wi ∈ Wi for i = 1, 2} ⊆ V.
Example. V = M2×2 (R), W1 and W2 as in the previous lecture. W1 + W2 = V .
Example. V = R3 , W1 = {the x-axis} = span(e1 ) and W2 = {the y-axis} = span(e2 ). Then
W1 + W2 = {(a, 0, 0) + (0, b, 0) : a, b ∈ R} = {(a, b, 0) : a, b ∈ R} = {the x, y-plane} = span(e1 , e2 ).
Proposition 1. If W1 , W2 are subspaces of V , then W1 + W2 is a subspace of V . Moreover, W1 + W2
contains both W1 and W2 .
Proof. 0 = 0 + 0 ∈ W + W 0 so W + W 0 6= ∅.
Suppose x, y ∈ W1 + W2 (must show x + y ∈ W1 + W2 ). By definition this means x = w1 + w2 and
y = u1 + u2 for some w1 , u1 ∈ W1 and w2 , u2 ∈ W2 . So
x + y = (w1 + w2 ) + (u1 + u2 )
= (w1 + u1 ) + (w2 + u2 ) (by commutativity, associativity).
| {z } | {z }
∈W1 ∈W2
So x + y ∈ W1 + W2 . A similar proof, using one of the distributive axioms, shows W1 + W2 is closed under
scalar multiplication. So W1 + W2 is a subspace of V .
Finally, w1 ∈ W1 =⇒
w1 = w1 + 0 ∈ W1 + W2
so W1 ⊆ W1 + W2 . Similarly, w2 = 0 + w2 ∈ W1 + W2 shows W2 ⊆ W1 + W2 .
Remark. W1 + W2 is the smallest subspace containing both W1 and W2 ; that is, every subspace that
contains W1 and W2 also contains W1 + W2 .
Proposition 2. span(S ∪ T ) = span(S) + span(T ).
Proof. (⊆) S ⊆ span(S) ⊆ span(S) + span(T ). Similarly, T ⊆ span(T ) ⊆ span(S) + span(T ). So
S ∪ T ⊆ span(S) + span(T ).
Since the RHS is a subspace (Prop. 1), we get
span(S ∪ T ) ⊆ span(S) + span(T )
by Theorem 1.6.
(⊇) Clearly span(S ∪ T ) is a subspace containing both span(S) and span(T ). Thus span(S ∪ T ) contains
the smallest subspace containing span(S) and span(T ), which by a Remark is span(S) + span(T ).
Remark. Can define W1 + W2 + · · · + Wk similarly. It is always a subspace, and is the smallest subspace
containing all W1 , . . . , Wk . span(S1 ∪ · · · ∪ Sk ) = span(S1 ) + · · · + span(Sk ).
Definition. If W1 , W2 are subspace of V , then we say that V is the direct sum of W1 and W2 , and we
write V = W1 ⊕ W2 , provided:
(1) W1 + W2 = V .
(2) Every v ∈ V can be written uniquely as v = w1 + w2 with w1 ∈ W1 and w2 ∈ W2 . This means if
v ∈ V and v = w1 + w2 = w10 + w20 with wi , wi0 ∈ Wi for i = 1, 2, then wi = wi0 for i = 1, 2.
Remark. We view this as “decomposing” or “factoring” the space V . This is especially useful in analyzing
the structure of linear operators on vector spaces (MATH 245).
0 b
Example. Let V = M2×2 (R) and W1 be as before. Let W3 = | b ∈ R . Then V = W1 ⊕ W3 ,
0 b
a b
because every ∈ V can be written
c d
a b a b−d 0 d
= +
c d c 0 0 d
| {z } | {z }
∈W1 ∈W3
v+W W
Note that v + W = v ′ + W
v′
2
MATH 146 February 3 Section 2
Lemma 1. Let W be a subspace of V . Let u, v ∈ V . Then
u + W = v + W ⇐⇒ u − v ∈ W.
Proof. (⇒) Assume u + W = v + W . Observe that u ∈ U + W . So u ∈ v + W , meaning u = v + w for
some w ∈ W . Then u − v ∈ W .
(⇐) Assume u − v =: w ∈ W . Then
u + W = (v + w) + W = v + (w + W ) = v + W (as w + W = W ; exercise.) □
Lemma 2. Let W be a subspace of V . The translations of W partition V ; that is,
(1) Their union is V , and
(2) If u + W ̸= v + W then (u + W ) ∩ (v + W ) = ∅.
Proof. (1) is obvious (v ∈ v + W for each v ∈ V ). To prove (2), we prove its contrapositive. Assume
(u + W ) ∩ (v + W ) ̸= ∅. Choose x ∈ (u + W ) ∩ (v + W ). This means
x=u+w for some w ∈ W
′
=v+w for some w′ ∈ W .
Then u + w = v + w′ , so u − v = w′ − w ∈ W . So u + W = v + W by Lemma 1. □
Note: the translations of W are the equivalence classes of an equivalence relation we can define on V ,
called “congruence mod W .” The definition is
u≡v (mod W ) ⇐⇒ u − v ∈ W.
Recall that for a fixed subspace W , V /W denotes the set of all translations of W . Thus
elements of V /W ↭ (certain) subsets of V
We use vectors in V to “name” elements of V /W (v “names” v + W ), but the vector names are not the
translations. In particular, elements of V /W can be “named” in more than one way. For example, if
V = R2 , W = span({(1, 1)}), and S = (−1, 0) + W , then also S = (0, 1) + W , since
(−1, 0) − (0, 1) = (−1, −1) ∈ W, so (−1, 0) + W = (0, 1) + W (Lemma 1).
We want to turn the set V /W into a vector space. So we need to define addition and scalar multiplication
on the elements of V /W . The idea is illustrated in the following example.
Example. V = R2 , W = span({(1, 1)}), S = (−1, 0) + W , T = (2, −1) + W .
W
S S+T T
u = (−1, 0)
Remark. Similarly, if F is the field of scalars for V , then we can define scalar multiplication on V /W by
def
c · (v + W ) = (cv) + W
and there is no problem (it is well-defined: if v + W = v ′ + W then (cv) + W = (cv ′ ) + W ).
Proposition. V /W with the operations defined as above is a vector space over F.
Proof. Ugh. There are 8 axioms that we must verify. Some are easy: e.g., commutativity (VS 1). Suppose
S, T ∈ V /W , say S = x + W and T = y + W . Then
df ∗ df
S + T = (u + W ) + (v + W ) = (u + v) + W = (v + u) + W = (v + W ) + (u + W ) = T + S.
∗
(= marks an application of (VS 1) in V .) Or left distributivity (VS 7):
df
a(S + T ) = a((x + W ) + (y + W )) = a((x + y) + W )
df ∗
= (a(x + y) + W ) = ((ax + ay) + W )
df df
= (ax + W ) + (ay + W ) = a(x + W ) + a(y + W ) = aS + aT.
Existence of a zero vector is more interesting. What is the zero vector in V /W ? (It is W .)
And the additive inverse of v+W is (−v)+W , because obviously (v+W )+((−v)+W ) = 0+W = W . □
Now that we know that V /W is a vector space, we can talk about linear combinations of “vectors” (i.e.,
translations of W ), linear independence, span, etc. The next theorem tells us what is dim(V /W ) when
dim V is finite.
Theorem. Suppose If V is finite-dimensional, say dim V = n. Let W be a subspace, say of dimension k.
Then dim(V /W ) = n − k = dim V − dim W .
Proof. Let B = {v1 , . . . , vk } be a basis for W . We can view B as a linearly independent subset of V . So
we can extend B to a basis C = {v1 , . . . , vk , vk+1 , . . . , vn } for V . Let D := {vk+1 + W, . . . , vn + W }. We
will show that D is a basis for V /W . Since |D| = n − k, this will prove the theorem.
We first show that D is linearly independent (in V /W ). Suppose ak+1 , . . . , an ∈ F and
(∗) ak+1 (vk+1 + W ) + · · · + an (vn + W ) = W.
Using the definitions of the operations in V /W , we can simplify this to
(ak+1 vk+1 + · · · + an vn ) + W = W.
2
By Lemma 1, this implies
ak+1 vk+1 + · · · + an vn ∈ W,
so
ak+1 vk+1 + · · · + an vn = b1 v1 + · · · bk vk for some b1 , . . . , bk ∈ F,
since {v1 , . . . , v)k } spans W . Then
(−b1 )v1 + · · · + (−bk )vk + ak+1 vk+1 + · · · + an vn = 0.
This is a linear combination of C. Since C is linearly independent, all the coefficients are 0. In particular,
ak+1 = · · · = an = 0. This proves that the only solution to (∗) is the trivial one, so D is linearly
independent.
Next we will show that D spans V /W . Let v + W ∈ V /W be given. Because C spans V , we can write
v = a1 v1 + · · · + ak vk + ak+1 vk+1 + · · · + an vn for some a1 , . . . , an ∈ F.
Claim. v + W = ak+1 (vk+1 + W ) + · · · + an (vn + W ) (so v + W ∈ span(D)).
Proof of Claim. The RHS of the Claim simplifies to u + W where u = ak+1 vk+1 + · · · + an vn . We want to
prove v + W = u + W . We can use Lemma 1: it is enough to prove v − u ∈ W . Well,
v − u = (a1 v1 + · · · + ak vk + ak+1 vk+1 + · · · + an vn ) − (ak+1 vk+1 + · · · + an vn )
= a1 v1 + · · · + ak vk ,
which is in W because v1 , . . . , vk ∈ W and W is a subspace. □
3
MATH 146 February 6 Section 2
Every element of V can be written as on the LHS, and in just one way (Thm. 1.8, Jan 23), so there is no
issue of being well-defined.
Clearly T vi = wi for each i = 1, . . . , n since
df
T vi = T (0·v1 + · · · + 1·vi + · · · + 0·vn ) = 0·w1 + · · · + 1·wi + · · · + 0·wn = wi .
Check that T is linear. Let x, y ∈ V . Say x = ni=1 ai vi and y = ni=1 bi vi . Then
P P
Xn Xn n
X n
X
T (x + y) = T ( (ai + bi )vi ) = (ai + bi )wi = ai w i + bi wi = T x + T y.
i=1 i=1 i=1 i=1
3
MATH 146 February 8 Section 2