MathIsimple
LA-5.2
Available

Properties of Determinants

Determinants possess remarkable algebraic properties that make them indispensable tools for matrix analysis. These properties—multiplicativity, behavior under transpose, and effects of elementary operations—form the computational backbone of linear algebra.

2-3 hours Core Level 10 Objectives
Learning Objectives
  • Prove and apply the multiplicativity property: det(AB) = det(A)det(B)
  • Master the transpose property: det(A^T) = det(A) and its consequences
  • Understand the deep connection between determinant and invertibility
  • Compute det(A^{-1}) = 1/det(A) for invertible matrices
  • Apply row and column operation effects on determinants systematically
  • Calculate determinants of all three types of elementary matrices
  • Master determinants of triangular matrices (upper, lower, diagonal)
  • Compute determinants of block triangular and block diagonal matrices
  • Understand det(kA) = k^n det(A) for n×n matrices
  • Recognize and avoid common mistakes involving determinant properties
Prerequisites
  • Determinant definition via axioms, permutations, and cofactors (LA-5.1)
  • Matrix multiplication and the associative law
  • Elementary matrices and their relationship to row operations
  • Matrix inverse and the equation AA⁻¹ = I
  • Basic properties of transpose
  • Understanding of linear independence and rank

1. The Multiplicativity Property

The most fundamental property of determinants is multiplicativity: the determinant of a product equals the product of determinants. This connects matrix algebra to scalar algebra and has geometric meaning—composing transformations multiplies volume scaling factors.

Theorem 5.4: Multiplicativity of Determinant

For any n×nn \times n matrices AA and BB over a field FF:

det(AB)=det(A)det(B)\det(AB) = \det(A) \cdot \det(B)
Proof:

Method 1 (Axiomatic): Fix matrix AA and define f(B)=det(AB)f(B) = \det(AB).

We verify ff satisfies the three axioms:

  • Multilinearity: ff is multilinear in rows of BB (since ABAB has rows that are linear in BB's rows)
  • Alternating: Swapping rows of BB swaps corresponding rows of ABAB
  • Normalization: f(I)=det(AI)=det(A)f(I) = \det(AI) = \det(A)

By uniqueness, f(B)=det(A)det(B)f(B) = \det(A) \cdot \det(B).

Method 2 (Elementary matrices): If BB is invertible, write B=E1EkB = E_1 \cdots E_k. Since det(AEi)=det(A)det(Ei)\det(AE_i) = \det(A) \cdot \det(E_i) for each elementary matrix, by induction det(AB)=det(A)det(B)\det(AB) = \det(A) \cdot \det(B).

If BB is not invertible, then ABAB is also not invertible, so both sides equal 0.

Example 5.10: Verifying Multiplicativity

Let A=(2113)A = \begin{pmatrix} 2 & 1 \\ 1 & 3 \end{pmatrix} and B=(1204)B = \begin{pmatrix} 1 & 2 \\ 0 & 4 \end{pmatrix}.

det(A)=61=5,det(B)=40=4\det(A) = 6-1 = 5, \quad \det(B) = 4-0 = 4
AB=(28114),det(AB)=288=20=54AB = \begin{pmatrix} 2 & 8 \\ 1 & 14 \end{pmatrix}, \quad \det(AB) = 28-8 = 20 = 5 \cdot 4 \checkmark
Corollary 5.2: Power of Determinant

For any n×nn \times n matrix AA and positive integer kk:

det(Ak)=(detA)k\det(A^k) = (\det A)^k
Proof:

By induction using multiplicativity: det(Ak)=det(Ak1)det(A)=(detA)k1det(A)=(detA)k\det(A^k) = \det(A^{k-1}) \cdot \det(A) = (\det A)^{k-1} \cdot \det(A) = (\det A)^k.

Remark 5.1: Geometric Interpretation

If TAT_A and TBT_B are linear transformations with matrices AA and BB:

  • det(A)|\det(A)| = volume scaling factor of TAT_A
  • det(AB)=det(A)det(B)|\det(AB)| = |\det(A)| \cdot |\det(B)| = volume scaling of TATBT_A \circ T_B

2. Determinant of Transpose

The transpose exchanges rows and columns. Remarkably, this preserves the determinant, establishing a powerful duality: every row property has a column counterpart.

Theorem 5.5: Transpose Property

For any n×nn \times n matrix AA:

det(AT)=det(A)\det(A^T) = \det(A)
Proof:

Using permutation formula: We have (AT)ij=aji(A^T)_{ij} = a_{ji}, so:

det(AT)=σSn(1)τ(σ)aσ(1),1aσ(n),n\det(A^T) = \sum_{\sigma \in S_n} (-1)^{\tau(\sigma)} a_{\sigma(1),1} \cdots a_{\sigma(n),n}

The map σσ1\sigma \mapsto \sigma^{-1} is a bijection on SnS_n with τ(σ)=τ(σ1)\tau(\sigma) = \tau(\sigma^{-1}).

The products are identical (just reordered), so the sums are equal.

Remark 5.2: Row-Column Duality

Since det(AT)=det(A)\det(A^T) = \det(A), all row properties have column versions:

  • Swapping two columns negates det
  • Scaling a column by cc multiplies det by cc
  • Adding a multiple of one column to another preserves det
  • Two identical columns means det = 0
Example 5.11: Transpose Verification

A=(1204)A = \begin{pmatrix} 1 & 2 \\ 0 & 4 \end{pmatrix} (upper triangular), det(A)=4\det(A) = 4.

AT=(1024)A^T = \begin{pmatrix} 1 & 0 \\ 2 & 4 \end{pmatrix} (lower triangular), det(AT)=4\det(A^T) = 4 \checkmark

Corollary 5.3: Products with Transpose
det(AAT)=det(ATA)=(detA)2\det(A \cdot A^T) = \det(A^T \cdot A) = (\det A)^2

3. Determinant and Invertibility

The connection between determinants and invertibility is fundamental. A matrix is invertible if and only if its determinant is nonzero—a simple scalar test for an otherwise complex property.

Theorem 5.6: Determinant of Inverse

If AA is an invertible n×nn \times n matrix:

det(A1)=1det(A)=(detA)1\det(A^{-1}) = \frac{1}{\det(A)} = (\det A)^{-1}
Proof:

Apply multiplicativity to AA1=IA \cdot A^{-1} = I:

det(A)det(A1)=det(I)=1\det(A) \cdot \det(A^{-1}) = \det(I) = 1

Since AA is invertible, det(A)0\det(A) \neq 0, so det(A1)=1/det(A)\det(A^{-1}) = 1/\det(A).

Theorem 5.7: Invertibility Criterion

For any n×nn \times n matrix AA:

A is invertible    det(A)0A \text{ is invertible} \iff \det(A) \neq 0
Proof:

(⟹) If AA is invertible, write A=E1E2EkA = E_1 E_2 \cdots E_k. Each det(Ei)0\det(E_i) \neq 0, so det(A)0\det(A) \neq 0.

(⟸) If AA is not invertible, columns are dependent. By multilinearity, any column can be written as a linear combination of others, giving det = 0.

Example 5.12: Checking Invertibility

Is A=(123456789)A = \begin{pmatrix} 1 & 2 & 3 \\ 4 & 5 & 6 \\ 7 & 8 & 9 \end{pmatrix} invertible?

det(A)=1(4548)2(3642)+3(3235)=3+129=0\det(A) = 1(45-48) - 2(36-42) + 3(32-35) = -3 + 12 - 9 = 0

Since det(A)=0\det(A) = 0, AA is not invertible.

Remark 5.3: Equivalent Conditions

For n×nn \times n matrix AA, these are equivalent:

  • AA is invertible
  • det(A)0\det(A) \neq 0
  • rank(A)=n\text{rank}(A) = n
  • Columns (or rows) are linearly independent
  • Ax=0Ax = 0 has only the trivial solution

4. Effects of Row and Column Operations

Understanding how elementary operations affect determinants is crucial for computation. These effects follow from the axiomatic definition and enable efficient algorithms.

Theorem 5.8: Effects of Row Operations

Let AA be n×nn \times n and AA' result from a row operation:

OperationNotationEffect
Swap rows ii, jjRiRjR_i \leftrightarrow R_jdet(A)=det(A)\det(A') = -\det(A)
Scale row ii by ccRicRiR_i \to c \cdot R_idet(A)=cdet(A)\det(A') = c \cdot \det(A)
Add kk × row jj to row iiRiRi+kRjR_i \to R_i + kR_jdet(A)=det(A)\det(A') = \det(A)
Proof:

Type I (Swap): Follows from alternating property.

Type II (Scale): By multilinearity: det(,cαi,)=cdet(,αi,)\det(\ldots, c\alpha_i, \ldots) = c \cdot \det(\ldots, \alpha_i, \ldots).

Type III (Add): det(,αi+kαj,,αj,)=det(A)+k0=det(A)\det(\ldots, \alpha_i + k\alpha_j, \ldots, \alpha_j, \ldots) = \det(A) + k \cdot 0 = \det(A).

Example 5.13: Row Reduction to Compute Determinant

Compute det(246125314)\det\begin{pmatrix} 2 & 4 & 6 \\ 1 & 2 & 5 \\ 3 & 1 & 4 \end{pmatrix}.

Step 1: Factor 2 from row 1: =2det(123125314)= 2 \cdot \det\begin{pmatrix} 1 & 2 & 3 \\ 1 & 2 & 5 \\ 3 & 1 & 4 \end{pmatrix}

Step 2: R2R1,R33R1R_2 - R_1, R_3 - 3R_1: =2det(123002055)= 2 \cdot \det\begin{pmatrix} 1 & 2 & 3 \\ 0 & 0 & 2 \\ 0 & -5 & -5 \end{pmatrix}

Step 3: R2R3R_2 \leftrightarrow R_3: =2det(123055002)= -2 \cdot \det\begin{pmatrix} 1 & 2 & 3 \\ 0 & -5 & -5 \\ 0 & 0 & 2 \end{pmatrix}

Step 4: Upper triangular: =21(5)2=20= -2 \cdot 1 \cdot (-5) \cdot 2 = 20

Remark 5.4: Column Operations

By transpose property, column operations have identical effects on determinant.

5. Determinants of Elementary Matrices

Elementary matrices are obtained by performing one row operation on the identity. Their determinants follow directly from the effect on det(I) = 1.

Theorem 5.9: Determinants of Elementary Matrices

The three types have determinants:

  • Row swap: det(Eij)=1\det(E_{ij}) = -1
  • Row scaling: det(Ei(c))=c\det(E_i(c)) = c
  • Row addition: det(Eij(k))=1\det(E_{ij}(k)) = 1
Example 5.14: Elementary Matrix Determinants
E12=(0110)E_{12} = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}

det=1\det = -1

E1(3)=(3001)E_1(3) = \begin{pmatrix} 3 & 0 \\ 0 & 1 \end{pmatrix}

det=3\det = 3

E12(2)=(1201)E_{12}(2) = \begin{pmatrix} 1 & 2 \\ 0 & 1 \end{pmatrix}

det=1\det = 1

Corollary 5.4: Left Multiplication Effect

det(EA)=det(E)det(A)\det(E \cdot A) = \det(E) \cdot \det(A), confirming row operation effects.

Remark 5.6: Right Multiplication (Column Operations)

Right multiplication AEA \cdot E performs column operations:

  • AEijA \cdot E_{ij} swaps columns ii and jj
  • AEi(c)A \cdot E_i(c) scales column ii by cc
  • AEij(k)A \cdot E_{ij}(k) adds kk times column ii to column jj

Same determinant effects: det(AE)=det(A)det(E)\det(A \cdot E) = \det(A) \cdot \det(E).

Example 5.18: Factoring a Matrix via Elementary Matrices

If A=(2312)A = \begin{pmatrix} 2 & 3 \\ 1 & 2 \end{pmatrix}, we can row reduce:

R1R2R_1 \leftrightarrow R_2: E12A=(1223)E_{12} \cdot A = \begin{pmatrix} 1 & 2 \\ 2 & 3 \end{pmatrix}

R22R1R_2 - 2R_1: E21(2)E12A=(1201)E_{21}(-2) \cdot E_{12} \cdot A = \begin{pmatrix} 1 & 2 \\ 0 & -1 \end{pmatrix}

So det(A)=det(E21(2))1det(E12)11(1)=1(1)(1)=1\det(A) = \det(E_{21}(-2))^{-1} \cdot \det(E_{12})^{-1} \cdot 1 \cdot (-1) = 1 \cdot (-1) \cdot (-1) = 1

Verify: det(A)=2231=1\det(A) = 2 \cdot 2 - 3 \cdot 1 = 1 \checkmark

6. Scalar Multiplication

Theorem 5.10: Scalar Multiple Property

For an n×nn \times n matrix AA and scalar kk:

det(kA)=kndet(A)\det(kA) = k^n \det(A)
Proof:

Writing AA with rows α1,,αn\alpha_1, \ldots, \alpha_n:

det(kA)=det(kα1,,kαn)=kndet(A)\det(kA) = \det(k\alpha_1, \ldots, k\alpha_n) = k^n \det(A)

Each of nn rows contributes a factor of kk by multilinearity.

Example 5.15: Scalar Multiplication

If AA is 4×44 \times 4 with det(A)=3\det(A) = 3:

det(2A)=243=48,det(A)=(1)43=3\det(2A) = 2^4 \cdot 3 = 48, \quad \det(-A) = (-1)^4 \cdot 3 = 3
Remark 5.5: Common Error

Warning: det(kA)kdet(A)\det(kA) \neq k \cdot \det(A) unless n=1n = 1!

Example 5.19: Sign of det(-A)

For an n×nn \times n matrix:

det(A)=(1)ndet(A)\det(-A) = (-1)^n \det(A)
  • nn even: det(A)=det(A)\det(-A) = \det(A)
  • nn odd: det(A)=det(A)\det(-A) = -\det(A)

Example: For 3×3 matrix with det(A) = 5: det(A)=(1)35=5\det(-A) = (-1)^3 \cdot 5 = -5

Corollary 5.5: Determinant of Similar Matrices

If B=P1APB = P^{-1}AP (similar matrices), then:

det(B)=det(P1AP)=det(P1)det(A)det(P)=1det(P)det(A)det(P)=det(A)\det(B) = \det(P^{-1}AP) = \det(P^{-1}) \det(A) \det(P) = \frac{1}{\det(P)} \det(A) \det(P) = \det(A)

Conclusion: Similar matrices have the same determinant.

7. Determinants of Special Matrices

Theorem 5.11: Triangular Matrices

For upper/lower triangular or diagonal matrices:

det(A)=a11a22ann=i=1naii\det(A) = a_{11} \cdot a_{22} \cdots a_{nn} = \prod_{i=1}^{n} a_{ii}
Example 5.16: Triangular Determinants
det(237015004)=2(1)4=8\det\begin{pmatrix} 2 & 3 & 7 \\ 0 & -1 & 5 \\ 0 & 0 & 4 \end{pmatrix} = 2 \cdot (-1) \cdot 4 = -8
Theorem 5.12: Block Triangular Matrices

For block triangular (A is k×k, B is r×r):

det(ACOB)=det(A)det(B)\det\begin{pmatrix} A & C \\ O & B \end{pmatrix} = \det(A) \cdot \det(B)
det(AOCB)=det(A)det(B)\det\begin{pmatrix} A & O \\ C & B \end{pmatrix} = \det(A) \cdot \det(B)
Proof:

Expand along the first column/row repeatedly. The block of zeros means only one cofactor is nonzero in each expansion step.

Example 5.17: Block Triangular

det(1234056700890012)\det\begin{pmatrix} 1 & 2 & 3 & 4 \\ 0 & 5 & 6 & 7 \\ 0 & 0 & 8 & 9 \\ 0 & 0 & 1 & 2 \end{pmatrix}

With A=(1205)A = \begin{pmatrix} 1 & 2 \\ 0 & 5 \end{pmatrix} and B=(8912)B = \begin{pmatrix} 8 & 9 \\ 1 & 2 \end{pmatrix}:

det=5(169)=57=35\det = 5 \cdot (16-9) = 5 \cdot 7 = 35
Theorem 5.13: Anti-Block Diagonal
det(OABO)=(1)krdet(A)det(B)\det\begin{pmatrix} O & A \\ B & O \end{pmatrix} = (-1)^{kr} \det(A) \cdot \det(B)

where A is k×r and B is r×k.

Definition 5.5: Schur Complement

For block matrix M=(ABCD)M = \begin{pmatrix} A & B \\ C & D \end{pmatrix} with AA invertible:

M/A=DCA1BM/A = D - CA^{-1}B

is called the Schur complement of AA in MM.

Theorem 5.14: Determinant via Schur Complement

If AA is invertible:

det(ABCD)=det(A)det(DCA1B)=det(A)det(M/A)\det\begin{pmatrix} A & B \\ C & D \end{pmatrix} = \det(A) \cdot \det(D - CA^{-1}B) = \det(A) \cdot \det(M/A)
Proof:

Block LU decomposition (Gaussian block elimination):

(IOCA1I)(ABCD)=(ABODCA1B)\begin{pmatrix} I & O \\ -CA^{-1} & I \end{pmatrix} \begin{pmatrix} A & B \\ C & D \end{pmatrix} = \begin{pmatrix} A & B \\ O & D - CA^{-1}B \end{pmatrix}

The left matrix has det = 1 (block lower triangular with I on diagonal), so:

det(ABCD)=det(ABODCA1B)=det(A)det(DCA1B)\det\begin{pmatrix} A & B \\ C & D \end{pmatrix} = \det\begin{pmatrix} A & B \\ O & D-CA^{-1}B \end{pmatrix} = \det(A) \cdot \det(D - CA^{-1}B)
Example 5.20: Schur Complement Calculation

Compute det(210131012)\det\begin{pmatrix} 2 & 1 & 0 \\ 1 & 3 & 1 \\ 0 & 1 & 2 \end{pmatrix} using Schur complement.

Take A=(2113)A = \begin{pmatrix} 2 & 1 \\ 1 & 3 \end{pmatrix}, B=(01)B = \begin{pmatrix} 0 \\ 1 \end{pmatrix}, C=(0,1)C = (0, 1), D=2D = 2.

A1=15(3112),CA1B=15(0,1)(3112)(01)=25A^{-1} = \frac{1}{5}\begin{pmatrix} 3 & -1 \\ -1 & 2 \end{pmatrix}, \quad CA^{-1}B = \frac{1}{5}(0,1)\begin{pmatrix} 3 & -1 \\ -1 & 2 \end{pmatrix}\begin{pmatrix} 0 \\ 1 \end{pmatrix} = \frac{2}{5}
det=det(A)(DCA1B)=5(225)=585=8\det = \det(A) \cdot (D - CA^{-1}B) = 5 \cdot (2 - \frac{2}{5}) = 5 \cdot \frac{8}{5} = 8
Remark 5.7: Vandermonde Determinant

A classic special determinant is the Vandermonde determinant:

det(1x1x12x1n11x2x22x2n11xnxn2xnn1)=1i<jn(xjxi)\det\begin{pmatrix} 1 & x_1 & x_1^2 & \cdots & x_1^{n-1} \\ 1 & x_2 & x_2^2 & \cdots & x_2^{n-1} \\ \vdots & \vdots & \vdots & \ddots & \vdots \\ 1 & x_n & x_n^2 & \cdots & x_n^{n-1} \end{pmatrix} = \prod_{1 \leq i < j \leq n} (x_j - x_i)

This is nonzero iff all xix_i are distinct, which is key in polynomial interpolation.

8. Common Mistakes to Avoid

Determinants have counterintuitive properties. Here are critical errors to avoid.

Mistake 1: det(A+B) = det(A) + det(B)

Wrong! Determinant is NOT additive.

Example: A = B = I₂. det(A) = det(B) = 1, but det(A+B) = det(2I) = 4 ≠ 2.

Mistake 2: det(cA) = c·det(A)

Wrong! It's cndet(A)c^n \det(A) for n×n matrices.

Example: det(2I₃) = 2³ = 8, not 2.

Mistake 3: Forgetting Sign in Row Swaps

Each row swap multiplies det by -1.

Tip: Count swaps during row reduction and adjust sign at the end.

Mistake 4: Confusing Row vs Matrix Scaling

Scaling ONE row by c → multiply det by c

Scaling WHOLE matrix by c → multiply det by c^n

Mistake 5: det(AB) = det(BA)?

True! This one is actually correct: det(AB) = det(A)det(B) = det(B)det(A) = det(BA).

Note: Even though AB ≠ BA in general, their determinants are always equal.

9. Key Takeaways

Multiplicative

det(AB) = det(A)det(B)

Transpose

det(A^T) = det(A)

Invertibility

A invertible ⟺ det(A) ≠ 0

Scalar Multiple

det(kA) = k^n det(A)

10. Practice Problems

Problem 1

If det(A)=3\det(A) = 3 and det(B)=2\det(B) = -2, find det(A2B1)\det(A^2 B^{-1}).

Problem 2

Let AA be 3×3 with det(A)=4\det(A) = 4. Find det(2AT)\det(2A^T) and det(A1)\det(A^{-1}).

Problem 3

Compute det(3000120045102364)\det\begin{pmatrix} 3 & 0 & 0 & 0 \\ 1 & 2 & 0 & 0 \\ 4 & 5 & -1 & 0 \\ 2 & 3 & 6 & 4 \end{pmatrix}.

Problem 4

Prove: If A2=AA^2 = A (idempotent), then det(A){0,1}\det(A) \in \{0, 1\}.

Problem 5

If AA is orthogonal (ATA=IA^T A = I), prove det(A)=±1\det(A) = \pm 1.

Problem 6

Compute det(1+a1111+b1111+c)\det\begin{pmatrix} 1+a & 1 & 1 \\ 1 & 1+b & 1 \\ 1 & 1 & 1+c \end{pmatrix}.

Solutions

Solution 1

det(A2B1)=det(A)2det(B)1=9(12)=92\det(A^2 B^{-1}) = \det(A)^2 \cdot \det(B)^{-1} = 9 \cdot (-\frac{1}{2}) = -\frac{9}{2}

Solution 2

det(2AT)=23det(AT)=84=32\det(2A^T) = 2^3 \cdot \det(A^T) = 8 \cdot 4 = 32
det(A1)=1det(A)=14\det(A^{-1}) = \frac{1}{\det(A)} = \frac{1}{4}

Solution 3

Lower triangular, so det = product of diagonal:

det=32(1)4=24\det = 3 \cdot 2 \cdot (-1) \cdot 4 = -24

Solution 4

From A2=AA^2 = A: det(A2)=det(A)\det(A^2) = \det(A)

So (detA)2=detA(\det A)^2 = \det A, meaning detA(detA1)=0\det A (\det A - 1) = 0.

Therefore detA=0\det A = 0 or detA=1\det A = 1.

Solution 5

From ATA=IA^T A = I:

det(ATA)=det(I)=1\det(A^T A) = \det(I) = 1
det(AT)det(A)=(detA)2=1\det(A^T) \cdot \det(A) = (\det A)^2 = 1

Therefore detA=±1\det A = \pm 1.

Solution 6

Use row/column operations. Subtract row 1 from rows 2, 3:

det=det(1+a11ab0a0c)\det = \det\begin{pmatrix} 1+a & 1 & 1 \\ -a & b & 0 \\ -a & 0 & c \end{pmatrix}

Expand along column 3:

=1det(aba0)+cdet(1+a1ab)= 1 \cdot \det\begin{pmatrix} -a & b \\ -a & 0 \end{pmatrix} + c \cdot \det\begin{pmatrix} 1+a & 1 \\ -a & b \end{pmatrix}
=1(0(ab))+c((1+a)b(a))=ab+c(b+ab+a)= 1 \cdot (0 - (-ab)) + c \cdot ((1+a)b - (-a)) = ab + c(b + ab + a)
=ab+bc+abc+ac=ab+bc+ca+abc= ab + bc + abc + ac = ab + bc + ca + abc

11. Quick Reference Summary

Core Properties

PropertyFormula
Multiplicativitydet(AB)=det(A)det(B)\det(AB) = \det(A) \det(B)
Transposedet(AT)=det(A)\det(A^T) = \det(A)
Inversedet(A1)=1/det(A)\det(A^{-1}) = 1/\det(A)
Powerdet(Ak)=(detA)k\det(A^k) = (\det A)^k
Scalar multipledet(kA)=kndet(A)\det(kA) = k^n \det(A)
InvertibilityAA invertible     det(A)0\iff \det(A) \neq 0

Elementary Matrix Determinants

TypeMatrixDeterminant
Row swapEijE_{ij}1-1
Row scaleEi(c)E_i(c)cc
Row additionEij(k)E_{ij}(k)11

Special Matrix Determinants

TypeFormula
TriangularProduct of diagonal: a11anna_{11} \cdots a_{nn}
Block triangulardet(A)det(B)\det(A) \cdot \det(B)
Anti-block diagonal(1)krdet(A)det(B)(-1)^{kr} \det(A) \det(B)
Schur complementdet(A)det(DCA1B)\det(A) \cdot \det(D - CA^{-1}B)
Similar matricesdet(P1AP)=det(A)\det(P^{-1}AP) = \det(A)

Row Operation Effects

OperationEffect on det
Swap two rowsMultiply by 1-1
Scale row by ccMultiply by cc
Add multiple of rowNo change

12. Historical Notes

Cauchy (1812): Proved the multiplicative property det(AB) = det(A)det(B), establishing determinants as a proper algebraic theory beyond just computational formulas. His work laid the foundation for understanding determinants as algebraic objects with consistent rules.

Jacobi (1841): Systematically developed properties of determinants and introduced the Jacobian determinant, connecting linear algebra to multivariable calculus. The Jacobian measures how volume changes under coordinate transformations.

Sylvester (1851): Coined the term "matrix" and developed the theory of matrix determinants alongside Arthur Cayley. Introduced the concept of minors and contributed to the understanding of block matrices.

Weierstrass: Contributed to the axiomatic understanding, showing that the three axioms (multilinear, alternating, normalized) uniquely determine the determinant. This abstract approach became foundational for modern linear algebra.

Modern view: Determinant is now understood as the unique alternating multilinear form on n vectors that equals 1 on the standard basis—connecting to exterior algebra and differential forms. In this view, det is a volume form.

13. Applications Preview

Determinant properties are used throughout mathematics and applications. Here are key areas where these properties are essential.

Cramer's Rule

Solve Ax=bAx = b using determinants: xi=det(Ai)/det(A)x_i = \det(A_i)/\det(A). Requires det(A) ≠ 0 (invertibility).

Eigenvalue Equations

Eigenvalues satisfy det(AλI)=0\det(A - \lambda I) = 0. The characteristic polynomial uses det properties extensively.

Change of Variables

In multivariable integrals: dV=detJdVdV' = |\det J| \, dV where JJ is the Jacobian matrix.

Linear Independence

Vectors v1,,vnv_1, \ldots, v_n are linearly independent iffdet[v1vn]0\det[v_1 \cdots v_n] \neq 0.

Area and Volume

Area of parallelogram: det[v,w]|\det[v, w]|. Volume of parallelepiped:det[u,v,w]|\det[u, v, w]|.

Orientation

Sign of det determines orientation. Positive = preserves handedness, negative = reverses (like reflection).

What's Next?

Now that you've mastered determinant properties, the next topics explore:

  • Computation Methods: Efficient techniques including row reduction and cofactor tricks
  • Laplace Expansion: Expansion along any row or column, generalized
  • Adjugate Matrix: The transpose of the cofactor matrix and Cramer's Rule

Key Skills You've Learned

  • Apply multiplicativity: det(AB) = det(A)det(B)
  • Use transpose property for row-column duality
  • Check invertibility via determinant
  • Track determinant changes through row operations
  • Compute determinants of triangular and block matrices
  • Use Schur complement for general block matrices

Chapter Summary

This module covered the fundamental algebraic properties of determinants. The key insight is that determinant behaves like a multiplicative homomorphism from matrices to scalars: it respects products but not sums. Combined with the transpose property, this gives determinants their computational power.

13

Theorems Covered

15

Practice Questions

12

FAQs Answered

Determinant Properties Practice
15
Questions
0
Correct
0%
Accuracy
1
det(AB)=\det(AB) =
Easy
Not attempted
2
det(AT)=\det(A^T) =
Easy
Not attempted
3
If AA is invertible, det(A1)=\det(A^{-1}) =
Medium
Not attempted
4
Determinant of an upper triangular matrix equals:
Easy
Not attempted
5
det(cA)\det(cA) for an n×nn \times n matrix AA equals:
Medium
Not attempted
6
If AA has two identical rows, then:
Easy
Not attempted
7
Adding a multiple of row jj to row ii (with iji \neq j):
Easy
Not attempted
8
det(Ak)=\det(A^k) =
Medium
Not attempted
9
The determinant of the elementary matrix EijE_{ij} (row swap) is:
Easy
Not attempted
10
The determinant of Ei(c)E_i(c) (scale row ii by cc) is:
Easy
Not attempted
11
The determinant of Eij(k)E_{ij}(k) (add kk times row jj to row ii) is:
Easy
Not attempted
12
For block diagonal matrix (AOOB)\begin{pmatrix} A & O \\ O & B \end{pmatrix}, det=\det =
Medium
Not attempted
13
If AA is 3×33 \times 3 and det(A)=2\det(A) = 2, then det(2A)=\det(2A) =
Medium
Not attempted
14
If det(A)=0\det(A) = 0, then:
Easy
Not attempted
15
For any square matrix AA: det(AAT)=\det(A \cdot A^T) =
Medium
Not attempted

Frequently Asked Questions

Why is det(AB) = det(A)det(B)?

The function f(B) = det(AB) is multilinear and alternating in the rows of B, with f(I) = det(A). By the uniqueness of determinants, f(B) must equal det(A)·det(B). Geometrically, composing linear transformations multiplies their volume scaling factors.

How do row operations affect the determinant?

There are three types: (1) Swap two rows → multiply det by -1; (2) Scale a row by c → multiply det by c; (3) Add a multiple of one row to another → det unchanged. These follow directly from the axiomatic definition (alternating, multilinear, normalized).

What's special about triangular matrix determinants?

For both upper and lower triangular matrices, det = product of diagonal entries. This makes Gaussian elimination extremely efficient: reduce to triangular form (tracking sign changes from swaps), then multiply the diagonal. This is O(n³) vs O(n!) for direct computation.

How does det relate to invertibility?

A square matrix A is invertible if and only if det(A) ≠ 0. Geometrically, det = 0 means the transformation collapses space (zero volume). Algebraically, det = 0 means the columns are linearly dependent, so A has a non-trivial kernel and cannot be injective.

Does det(A+B) = det(A) + det(B)?

NO! This is a common mistake. Determinant is NOT additive. It's only linear in each row separately (multilinear). Counterexample: A = B = I₂ gives det(A) = det(B) = 1, but det(A+B) = det(2I) = 4 ≠ 2.

Why does det(A^T) = det(A)?

In the permutation formula, transposing swaps row and column indices. Each permutation σ appears paired with its inverse σ⁻¹ (which has the same sign), so the sum is unchanged. This duality means all row properties have column versions.

How do I compute det of a block matrix?

For block triangular matrices: det([A,B;O,D]) = det(A)·det(D) and det([A,O;C,B]) = det(A)·det(B). For general block matrices [A,B;C,D], if A is invertible: det = det(A)·det(D - CA⁻¹B). The quantity D - CA⁻¹B is called the Schur complement.

What are the determinants of elementary matrices?

det(E_ij) = -1 (row swap), det(E_i(c)) = c (scale row by c), det(E_ij(k)) = 1 (add k times row j to row i). These follow from applying each operation to the identity matrix and using the determinant axioms.

Why is det(kA) = k^n·det(A) and not k·det(A)?

When you multiply a matrix by scalar k, EVERY row gets multiplied by k. Since det is linear in each row, multiplying n rows by k gives k^n factor. For example, det(2·I₃) = 2³ = 8, not 2.

Can determinant be negative?

Yes! The sign of det indicates orientation. det > 0 means the transformation preserves orientation, det < 0 means it reverses orientation (like a reflection). Only |det| measures volume scaling.

How is det(A⁻¹) related to det(A)?

For invertible A: det(A⁻¹) = 1/det(A). This follows from det(A)·det(A⁻¹) = det(A·A⁻¹) = det(I) = 1. Note that A must be invertible (det(A) ≠ 0) for this to make sense.

What happens to det when I multiply a single row by a scalar?

Multiplying ONE row by scalar c multiplies det by c (multilinearity). This is different from multiplying the WHOLE matrix by c, which gives c^n·det(A). Common mistake: confusing these two operations.