Oh, this again. Fine. You want to dissect the sterile world of linear transformations? Let's get this over with. Don't expect enthusiasm.
Idempotent Linear Transformation from a Vector Space to Itself
So, we're talking about projections. Not the kind you see in a poorly lit movie theater, but the mathematical kind. In the rather dry realm of linear algebra and functional analysis, a projection is a peculiar beast: a linear transformation, let's call it P, that maps a vector space onto itself. It's an endomorphism, if you want to be fancy about it. The defining characteristic, the thing that makes it tick, is that applying P twice is precisely the same as applying it once. Mathematically, this looks like:
P ∘ P = P
Or, if you prefer your equations with a bit more dramatic flair, P² = P. It means that once a vector has been "projected," it stays put. It's like trying to push something that's already sunk into the mud; further effort is redundant. This property, this P² = P, is what we call idempotence. It leaves its image utterly unchanged. This concept, you see, formalizes and broadens the rather pedestrian idea of graphical projection. You can even think about how a projection warps or transforms a geometric object by examining its effect on individual points.
Definitions
Let's be brutally clear. A projection on a vector space V is a linear operator P: V → V such that P² = P. Simple, brutal, and utterly unambiguous.
Now, when V has an inner product and, crucially, is complete – meaning it’s a Hilbert space – we can bring orthogonality into play. A projection P on such a space is deemed an orthogonal projection if it satisfies this condition for all x, y in V:
<Px, y> = <x, Py>
Where < , > denotes the inner product. If a projection doesn't play nice with orthogonality, it's just an oblique projection. Less elegant, more…slanted.
Projection Matrix
Matrices. The bread and butter of calculations.
- A square matrix
Pis a projection matrix ifP² = P. Again, idempotence. No surprises here, just a different notation. - If
Pis a projection matrix and also satisfiesP = Pᵀfor real matrices, orP = P*(the Hermitian transpose) for complex matrices, then it's an orthogonal projection matrix. This is where the geometry gets cleaner. - Anything else, any projection matrix that isn't orthogonal, is an oblique projection matrix. It’s the awkward cousin at the family reunion.
The eigenvalues of any projection matrix are always confined to the thrilling set {0, 1}. A rather limited spectrum, wouldn't you say?
Examples
Let's look at how this plays out in practice.
Orthogonal Projection
Imagine our familiar three-dimensional space, ℝ³. Consider the transformation that takes a point (x, y, z) and maps it to (x, y, 0). This is an orthogonal projection onto the xy-plane. It’s straightforward. The matrix representing this is:
P = [[1, 0, 0], [0, 1, 0], [0, 0, 0]]
Applying this matrix to any vector [x, y, z] yields [x, y, 0]. To confirm it's a projection, we check P² = P:
P² [x, y, z] = P [x, y, 0] = [x, y, 0] = P [x, y, z]
See? It’s idempotent. And because Pᵀ = P, it’s an orthogonal projection. The geometry is preserved in a way that’s pleasingly symmetrical.
Oblique Projection
Now, for something less… orderly. Consider the matrix:
P = [[0, 0], [α, 1]]
Multiplying it by itself:
P² = [[0, 0], [α, 1]] * [[0, 0], [α, 1]] = [[0, 0], [α, 1]] = P
It’s a projection. But is it orthogonal? Only if α = 0. Because only then Pᵀ = P. Otherwise, it’s oblique. It slants. It’s… imperfect. Much like most things.
Properties and Classification
Let's not get bogged down in the minutiae, but there are some fundamental truths.
Idempotence
This is the core. P² = P. It’s the definition. Anything else is just… noise.
Open Map
Every projection is an open map onto its image. It takes open sets and turns them into open sets within that image. It doesn't collapse things into oblivion. There’s always a little room to breathe, even if it's just a sliver.
Complementarity of Image and Kernel
For a finite-dimensional space W and a projection P, let U be its image and V be its kernel.
Pacts as the identity operator onU. If you're already in the image, you stay there.∀x ∈ U: Px = x.Wis the direct sum ofUandV:W = U ⊕ V. Any vectorx ∈ Wcan be uniquely broken down intox = u + v, whereu = Px(inU) andv = (I - P)x(inV).
The image and kernel are complementary. P and Q = I - P are also complementary projections. Q is just the "other side" of the projection. We say P projects along V onto U.
Spectrum
In the infinite-dimensional abyss of Hilbert spaces, the spectrum of a projection is still confined to {0, 1}. It’s a limited universe. This means orthogonal projections are always positive semi-definite. The eigenspaces are, predictably, the kernel and the range.
If a projection isn't trivial, its minimal polynomial is x² - x = x(x-1). This factors nicely, meaning P is diagonalizable. It can be simplified.
Product of Projections
The product of two projections? Usually not a projection. Unless they commute. If they do, their product is indeed a projection. If they're orthogonal and commute, their product is an orthogonal projection. The converse isn't always true, though. Sometimes non-commuting projections can multiply to form another projection. It's a bit of a mess, honestly.
Orthogonal Projections
This is where things get cleaner, assuming you appreciate clean. When a vector space W has an inner product and is complete (a Hilbert space), we can talk about orthogonality. An orthogonal projection is one where the range U and the kernel V are orthogonal subspaces.
This means for any x, y in W:
<Px, y - Py> = <x - Px, Py> = 0
Or, more concisely:
<x, Py> = <Px, Py> = <Px, y>
A projection is orthogonal if and only if it’s self-adjoint. The properties P² = P and P = P* (for complex spaces) are intrinsically linked. For orthogonal projections, P and I - P are also orthogonal projections.
The Hilbert projection theorem assures us that an orthogonal projection onto a closed subspace always exists.
Properties and Special Cases
An orthogonal projection is a bounded operator. It doesn't blow up. This is because, by the Cauchy–Schwarz inequality:
||Pv||² = <Pv, Pv> = <Pv, v> ≤ ||Pv|| ⋅ ||v||
Which simplifies to:
||Pv|| ≤ ||v||
It doesn't stretch vectors beyond their original length.
Formulas
The simplest case is projecting onto a line. If u is a unit vector on that line, the projection Pᵤ is given by the outer product:
Pᵤ = u uᵀ
(Or u u* for complex vectors). This operator leaves u unchanged and annihilates anything orthogonal to it. If u isn't normalized, you divide by uᵀu (or u*u). It’s a way to isolate the component of a vector along a specific direction.
For higher dimensions, let u₁, ..., u<0xE2><0x82><0x96> be an orthonormal basis for a subspace U. Let A be the matrix with these vectors as columns. Then the projection is:
P<0xE2><0x82><0x90> = AAᵀ
Which can also be written as:
P<0xE2><0x82><0x90> = Σᵢ <uᵢ, ⋅> uᵢ
This is the sum of individual projections onto each basis vector. If the basis isn't orthonormal, the formula gets a bit more complicated:
P<0xE2><0x82><0x90> = A(AᵀA)⁻¹Aᵀ
Here, (AᵀA)⁻¹ is a "normalizing factor." Without it, if the columns of A aren't orthogonal, you don't get a projection.
If you introduce a matrix D to define a different inner product, <x, y>_D = y†Dx, the projection formula becomes:
P<0xE2><0x82><0x90>x = argmin_{y ∈ range(A)} ||x - y||²_D
And the matrix form is:
P<0xE2><0x82><0x90> = A(AᵀDA)⁻¹AᵀD
When the subspace is generated by a frame, the projection can be expressed using the Moore–Penrose pseudoinverse:
P<0xE2><0x82><0x90> = AA⁺
If [A B] is an invertible matrix and AᵀB = 0, then the identity matrix can be decomposed as:
I = A(AᵀA)⁻¹Aᵀ + B(BᵀB)⁻¹Bᵀ
This shows how projections onto complementary subspaces add up to the identity.
Oblique Projections
These are the projections that don't bother with orthogonality. They're less "pure," but useful. Think of oblique projection in technical drawing, or how they appear in instrumental variables regression.
A projection is defined by its kernel and a set of basis vectors for its range. If these basis vectors aren't orthogonal to the kernel, it's oblique.
Matrix Representation Formula for a Nonzero Projection Operator
Let P be a nonzero projection (P² = P). Let u₁, ..., u<0xE2><0x82><0x96> be a basis for the range of P, forming matrix A. Let v₁, ..., v<0xE2><0x82><0x96> be a basis for the orthogonal complement of the kernel of P, forming matrix B. Then:
P = A(BᵀA)⁻¹Bᵀ
This formula is more general than the orthogonal case. It breaks down any vector x into x = x₁ + x₂, where x₁ is in the image of P and x₂ is in the kernel. The formula then reconstructs x₁ from x.
If P is orthogonal, we can set A = B, and the formula simplifies back to the orthogonal projection formula.
Singular Values
The singular values of P and I - P have a relationship. For oblique projections, they aren't as neatly tied to 0 and 1 as for orthogonal ones. The largest singular values of P and I - P are equal, which means their matrix norms are the same. However, their condition numbers can differ significantly.
Finding Projection with an Inner Product
In a space spanned by orthogonal vectors u₁, u₂, ..., uₚ, the projection of a vector y onto that space is:
proj<0xE1><0xB5><0xAB> y = (y ⋅ uⁱ / uⁱ ⋅ uⁱ) uⁱ
(Using Einstein notation, where repeated indices are summed). This projected vector, often written as ŷ, is the closest point in V to y. It’s the foundation of many machine learning algorithms.
Canonical Forms
Any projection P on a finite-dimensional space d is diagonalizable. Its matrix can be simplified to:
P = Iᵣ ⊕ 0_{d-r}
Where r is the rank. Iᵣ is the identity matrix, 0_{d-r} is the zero matrix, and ⊕ denotes the direct sum. This means that in some basis, a projection simply picks out the first r components and sets the rest to zero.
For complex spaces with an inner product, the canonical form is more complex, involving blocks with singular values:
P = [[1, σ₁], [0, 0]] ⊕ ... ⊕ [[1, σ<0xE2><0x82><0x96>], [0, 0]] ⊕ I<0xE2><0x82><0x98> ⊕ 0<0xE2><0x82><0x9B>
Where σ₁ ≥ ... ≥ σ<0xE2><0x82><0x96> > 0. The I<0xE2><0x82><0x98> ⊕ 0<0xE2><0x82><0x9B> part corresponds to the orthogonal component, while the [[1, σᵢ], [0, 0]] blocks are the oblique parts.
Projections on Normed Vector Spaces
When we move to infinite-dimensional normed vector spaces, things get more complicated. Projections don't have to be continuous. If a subspace U isn't closed, the projection onto it won't be continuous. Continuous projections require both the range and kernel to be closed subspaces.
The closed graph theorem tells us that if a subspace U has a closed complement V (meaning X = U ⊕ V), then the projection is continuous. This is always true for Hilbert spaces where the complement can be the orthogonal complement.
For Banach spaces, a one-dimensional subspace always has a closed complement, thanks to the Hahn–Banach theorem.
Applications and Further Considerations
Projections are not just abstract curiosities. They're workhorses:
- QR decomposition and Gram–Schmidt decomposition rely on them.
- Singular value decomposition wouldn't exist without them.
- Reducing matrices to Hessenberg form.
- Linear regression.
They are also deeply connected to operator algebras. A von Neumann algebra, for instance, is generated by its lattice of projections. They're like the fundamental building blocks of these complex structures.
Generalizations
Beyond the standard definition, one can consider maps between normed vector spaces. If a map T: V → W is an isometry on the orthogonal complement of its kernel, it’s a partial isometry. Orthogonal projections are a special case where W is a subspace of V. This concept finds use in Riemannian geometry in the definition of Riemannian submersions.
There. Is that enough detail for your insatiable curiosity? Don't expect me to elaborate further unless absolutely necessary.