Inner products generalize the dot product to abstract vector spaces, providing a way to measure lengths, angles, and orthogonality. They form the foundation for the geometric structure of Hilbert spaces and underpin applications from quantum mechanics to signal processing.
This module introduces the axiomatic definition of inner products, explores key examples on various spaces, and establishes fundamental inequalities that govern the geometry of inner product spaces.
An inner product equips a vector space with geometric structure—the ability to measure lengths and angles. We begin with the abstract definition, then explore concrete examples.
Let be a vector space over . An inner product on is a function satisfying:
For real inner products, symmetry plus linearity in the first argument implies linearity in the second argument:
Thus real inner products are bilinear (linear in both arguments).
Let be a vector space over . An inner product on is a function satisfying:
Complex inner products are sesquilinear (conjugate-linear in the second argument):
Note: Some texts use the opposite convention (linear in second argument). We follow the physicist's convention.
Conjugate symmetry ensures is always real:
Without this, positive definiteness wouldn't make sense for complex spaces.
A vector space equipped with an inner product is called an inner product space (or pre-Hilbert space).
A Hilbert space is a complete inner product space—one where every Cauchy sequence converges. Finite-dimensional inner product spaces are automatically complete (and hence Hilbert spaces).
The inner product axioms can be stated in several equivalent ways:
These are all equivalent up to which argument is linear vs. conjugate-linear.
For any positive definite Hermitian matrix , the function defines an inner product on .
Linearity:
Conjugate symmetry:
Positive definiteness: Since is positive definite, for all .
Show that is an inner product on .
Linearity:
Symmetry:
Positive definiteness: , with equality iff (for continuous ).
If we relax positive definiteness to (allowing for ), we get a semi-inner product or pseudo-inner product. This induces a seminorm rather than a norm.
The concept of inner product was developed in the early 20th century. Key contributors include:
Consider on .
For : but .
This violates positive definiteness, so it's NOT an inner product. (It's a symmetric bilinear form, but indefinite.)
If a norm satisfies the parallelogram law, then it comes from an inner product defined by the polarization identity:
for real spaces. For complex spaces, use the full polarization formula.
Inner products provide:
The standard (Euclidean, dot) inner product on :
For and :
The standard inner product on :
where is the conjugate transpose (Hermitian transpose).
For and :
For positive weights , define on :
This is useful when different coordinates have different importance or units.
On continuous functions :
For and on :
On (polynomials of degree ≤ n):
Legendre polynomials are orthogonal with respect to this inner product.
On :
This treats a matrix as a vector of entries.
For any positive definite matrix :
This defines a valid inner product. When , we recover the standard inner product.
The space of square-summable sequences :
This is an infinite-dimensional Hilbert space, fundamental in functional analysis.
For a positive weight function :
Different weights give different orthogonal polynomial families:
On , the Frobenius inner product:
This induces the Frobenius norm:
Inner products can be classified by their domain:
Every inner product induces a norm (notion of length):
The induced norm satisfies:
(1) Follows directly from positive definiteness of inner product.
(2)
(3) Follows from Cauchy-Schwarz (proven next).
On with standard inner product:
For :
On :
For on :
The 1-norm and ∞-norm do NOT come from any inner product. Only norms satisfying the parallelogram law (see below) are induced by inner products.
The induced norm defines a metric (distance function):
This satisfies: (1) , (2) , (3) , (4) .
In with standard inner product, find the distance between and :
The distance between and on :
For any vectors :
Expand using the definition of norm:
By conjugate symmetry:
For real inner product spaces:
A vector with is called a unit vector. For any nonzero , the vector is the normalization of (a unit vector in the same direction).
Normalize :
Verify: ✓
For any vectors in an inner product space:
Equality holds if and only if and are linearly dependent.
If , both sides are 0. Assume .
For any scalar , positive definiteness gives:
Choose (the optimal value):
Rearranging:
Taking square roots gives the result. Equality holds iff for some scalar .
By Cauchy-Schwarz:
For and :
Indeed, ✓
For :
The Cauchy-Schwarz inequality is one of the most important inequalities in mathematics:
For sequences and :
Example: ✓
Consider the quadratic .
Since for all real , the discriminant must be non-positive:
This gives , and similarly for the imaginary part.
Equality holds if and only if:
In other words, equality holds iff and are linearly dependent.
For and :
Equality holds because (linearly dependent).
For nonzero vectors in a real inner product space:
when the denominator is nonzero, where is the angle between and .
For random variables with finite second moments:
This is Cauchy-Schwarz with , the inner product.
Equality holds iff and are linearly related (one is a constant multiple of the other).
The correlation coefficient satisfies:
This is a direct consequence of Cauchy-Schwarz applied to centered random variables.
Cauchy-Schwarz generalizes to:
For positive :
Apply Cauchy-Schwarz to and .
For nonzero vectors in a real inner product space, the angle between them is defined by:
By Cauchy-Schwarz, , so is well-defined.
Vectors and are orthogonal (perpendicular), written , if:
This corresponds to (or radians).
Vectors and :
So .
For and :
So .
If , then:
since .
If are pairwise orthogonal:
For a subset , the orthogonal complement is:
For any subset , is a subspace of .
Zero vector: for all , so .
Closure: If and are scalars:
On with :
The functions and are orthogonal for all integers :
This orthogonality is the foundation of Fourier series.
Find the angle between and on :
Since , the angle is — they are orthogonal!
A set of nonzero pairwise orthogonal vectors is linearly independent.
Suppose with for .
Take inner product with :
Since , we have , so .
In an -dimensional inner product space, any orthogonal set has at most nonzero vectors.
A set is orthonormal if:
That is, the vectors are pairwise orthogonal and each has unit length.
The standard basis of is orthonormal:
We have for the standard inner product.
The rotated basis where:
Verify: and .
If is orthonormal, coordinates are easy to find:
The coefficient of is simply —no system of equations needed!
If is an orthonormal basis and :
The squared norm equals the sum of squared coefficients—a generalized Pythagorean theorem.
In any inner product space:
Expand both sides using the inner product:
Adding these gives the result.
The parallelogram law states that the sum of the squares of the diagonals of a parallelogram equals the sum of the squares of all four sides. This is a fundamental property that characterizes inner product spaces.
A normed vector space is an inner product space (with the norm induced by that inner product) if and only if the norm satisfies the parallelogram law.
On with 1-norm, take :
Since , the 1-norm doesn't come from an inner product.
In a real inner product space:
In a complex inner product space:
The polarization identity shows that the inner product is completely determined by the norm. If you know all the lengths, you can compute all the inner products (and hence all the angles).
Given in a real inner product space, find :
The vectors are orthogonal! (This is the 3-4-5 right triangle.)
For any vectors and their midpoint :
This relates distances from a point to the endpoints and midpoint of a segment.
Apply the parallelogram law to vectors and :
The parallelogram law completely characterizes inner product spaces among normed spaces. This deep result (Jordan-von Neumann, 1935) shows that the parallelogram law is the only additional axiom needed to get from a normed space to an inner product space.
In with :
Both sides equal 30 ✓
In a real inner product space, the polarization identity can be written as:
In complex inner product spaces, we need all four terms:
The inner product is continuous: if and (in norm), then:
By Cauchy-Schwarz:
By triangle inequality:
So . By symmetry, .
For subspaces of inner product space :
Let (xy-plane in ).
Then (z-axis).
And .
In finite-dimensional spaces, always. In infinite dimensions, we only have , with equality iff is closed.
For a subspace of finite-dimensional inner product space :
Since (direct sum), dimensions add.
In , let . Find .
A vector iff .
So .
Check: ✓
Let be a closed subspace of inner product space . For any , there exists a unique minimizing . This is characterized by:
The vector in the theorem is called the orthogonal projection of onto , written or . It minimizes distance to .
Project onto the xy-plane :
The residual (the z-axis).
If is an orthonormal basis for subspace :
We need . For any :
Let be an -dimensional real inner product space with basis . Define the Gram matrix:
Then for and :
A symmetric matrix is positive definite if:
Equivalently, all eigenvalues of are positive.
defines an inner product on if and only if is symmetric and positive definite.
Let . Check positive definiteness:
for . So is a valid inner product.
For the standard inner product on with standard basis, . A basis where is called an orthonormal basis.
On , defines an inner product iff is Hermitian () and positive definite.
Is positive definite?
Check : Yes (conjugate transpose equals itself).
Eigenvalues:
Both positive, so defines a valid inner product.
A matrix is positive definite if and only if it can be written as:
where is lower triangular with positive diagonal entries. This is the Cholesky decomposition.
Factor :
The positive diagonal in confirms is positive definite.
A symmetric matrix is positive definite iff all leading principal minors are positive:
where is the upper-left submatrix.
Is positive definite?
Yes, is positive definite.
If we change basis with invertible matrix (new coordinates ), the Gram matrix transforms as:
Inner product is preserved: .
Every real symmetric positive definite can be diagonalized by an orthogonal matrix :
where with .
For , use , NOT . Without conjugation, can be negative!
Only norms satisfying the parallelogram law are induced by inner products. The 1-norm and ∞-norm are NOT inner product norms.
Positive definite: for . Positive semidefinite allows for some . Only positive definite forms are inner products.
Some texts use linearity in the second argument. Be consistent! We use linearity in the first argument (physicist's convention).
The inequality is , NOT . Don't forget the absolute value!
The inner product takes two arguments and can be negative. The norm takes one argument and is always non-negative.
When defining a new "inner product," always verify all three axioms. Positive definiteness is often the hardest to check—make sure only when .
If and , it does NOT follow that . Example: and , but .
Triangle inequality: . Reverse triangle: . Don't confuse them!
Quantum states live in complex Hilbert spaces. The inner product gives probability amplitudes. Orthogonal states are distinguishable.
The inner product measures signal correlation. Orthogonal signals don't interfere. Fourier analysis uses orthogonal basis functions.
Covariance is an inner product. Correlation = cosine of angle. Kernel methods use inner products in feature spaces.
Lighting calculations use dot products (inner products). Surface normals and view directions determine shading.
Least squares uses orthogonal projections. Krylov methods (like conjugate gradient) exploit inner products for efficient solving.
Best approximations minimize distance (norm). Orthogonal polynomials (Legendre, Chebyshev) arise from different inner products.
Given an inconsistent system , least squares finds minimizing . The solution satisfies the normal equations:
This is the orthogonal projection of onto the column space of .
Fit a line to points :
Normal equations: ,
Solving: , so the best-fit line is .
Any periodic function can be written as:
The coefficients are inner products: .
In machine learning and data science:
Measures "length" of vectors
Most important inequality!
Generalizes perpendicularity
Inner products generalize the familiar dot product to abstract vector spaces, providing the geometric concepts of length, angle, and orthogonality. The Cauchy-Schwarz inequality is the cornerstone result that makes everything work.
20+
Theorems
25+
Examples
12
Quiz Questions
10
FAQs
Adjoint operators, unitary/orthogonal matrices
Spectral theorem, orthogonal diagonalization
SVD, least squares, Fourier analysis
With inner products mastered, you're ready for:
The dot product is a specific inner product on ℝⁿ. An inner product is a more general concept that can be defined on any vector space satisfying the axioms (linearity, conjugate symmetry, positive definiteness). Every inner product induces a notion of 'dot product' in its space.
Without conjugation, ⟨x,x⟩ could be negative or complex for nonzero x. Conjugate symmetry ensures ⟨x,x⟩ is always real, and positive definiteness makes it positive for x ≠ 0. This is essential for defining a valid norm.
It guarantees that ||x|| = √⟨x,x⟩ is a valid norm: (1) ||x|| ≥ 0, (2) ||x|| = 0 iff x = 0, (3) ||cx|| = |c|||x||. Without it, we couldn't measure 'length' properly.
It's fundamental! It proves the triangle inequality, defines angles between vectors (since |cos θ| ≤ 1), bounds correlations in statistics, and appears throughout analysis, probability, and physics.
No! A norm comes from an inner product if and only if it satisfies the parallelogram law: ||x+y||² + ||x-y||² = 2||x||² + 2||y||². The 1-norm and ∞-norm fail this test.
It shows that if you know the norm, you can recover the inner product (when it exists). This means all geometric information is encoded in lengths alone—angles are derived quantities.
Function spaces like L²[a,b] are infinite-dimensional inner product spaces crucial for Fourier analysis, quantum mechanics, and differential equations. The same geometric intuition (length, angle, orthogonality) extends to these spaces.
A complete inner product space—meaning every Cauchy sequence converges. ℝⁿ and ℂⁿ are finite-dimensional Hilbert spaces. L²[a,b] is an infinite-dimensional example, fundamental in functional analysis.
For finite-dimensional spaces, ⟨x,y⟩ = xᵀAy for some positive definite matrix A. The standard inner product uses A = I. Changing A changes the geometry (lengths and angles) of the space.
Bilinear means linear in both arguments. Sesquilinear (Latin: 'one-and-a-half linear') means linear in one argument, conjugate-linear in the other. Complex inner products are sesquilinear; real inner products are bilinear.