To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

Projection (linear algebra)

From Wikipedia, the free encyclopedia

The transformation P is the orthogonal projection onto the line m.

In linear algebra and functional analysis, a projection is a linear transformation from a vector space to itself (an endomorphism) such that . That is, whenever is applied twice to any vector, it gives the same result as if it were applied once (i.e. is idempotent). It leaves its image unchanged.[1] This definition of "projection" formalizes and generalizes the idea of graphical projection. One can also consider the effect of a projection on a geometrical object by examining the effect of the projection on points in the object.

YouTube Encyclopedic

  • 1/5
    Views:
    211 040
    32 206
    470 426
    471 991
    87 187
  • Projections onto subspaces | Linear Algebra | Khan Academy
  • Projection into Subspaces
  • Introduction to projections | Matrix transformations | Linear Algebra | Khan Academy
  • 15. Projections onto Subspaces
  • Linear Algebra 6.2.2 Orthogonal Projections

Transcription

Many videos ago we introduced the idea of a projection. And in that case we dealt more particularly with projections onto lines that went through the origin. So if we had some line-- let's say L-- and let's say L is equal to the span of some vector v. Or you could say, alternately, that L is equal to the set of all multiples of v, such that the scalar factors are just any real numbers. These are both representations of lines that go through the origin. We defined a projection of any vector onto that line. Let me just draw it real fast. So let me see, we draw some axes. So that is my-- I want to draw it a little bit straighter than that-- that is my vertical axis and that is my horizontal axis. Just like that and let's say I have some line that goes through the origin. Let's say-- that doesn't go through the origin-- let's say that that line right there goes through the origin. So that is L. We knew visually that a projection of some vector x onto L-- so let's say that that is a vector x. Visually, if you were to draw-- if you have some light coming straight down it would be the shadow of x onto L. So this right here, that right there, was the projection onto the line L of the vector x. And we defined it more formally. We kind of took a perpendicular. We said that x minus the projection of x onto L is perpendicular to the line L, or perpendicular to everything-- orthogonal to everything-- on the line L. But this is how at least I visualize this. It's kind of the shadow as you go down onto the line L. And this was a special case, in general, of projections. You might notice that L is going to be a valid subspace. You could prove it to yourself. It contains the zero vector. It goes through the origin. It's closed under addition-- any member of it plus any other member of it is going to be another member of it. It's closed under scalar multiplication-- you can take any member of it and scale it up or down, it's still going to be an L. So this was a subspace when we defined this. And just as a bit of a reminder of what it was, we were able to figure out what this projection is for some line L. If you have some spanning vector, the projection onto this line L that goes through the origin of the vector x, we figured out was x dot your spanning vector for your line, so x dot v over v dot v, which is really just the length of v squared. So all of this was a number and you want it to be in the same direction as your line. It's going to be another vector in your line. So it's going to be times the vector v. So it's just going to be a scaled up or scaled down version of your spanning vector. Maybe your spanning vector is like that. And really any vector in your line could be a spanning vector. Any vector other than the zero vector. Now that was a projection onto a line which was a special kind of subspace. But now we're going to broaden our definition of a projection to any subspace. So we already know that if-- let me draw a little dividing line to show that we're doing something slightly different-- if v is a subspace of Rn then v complement is also a subspace space of Rn. So the orthogonal complement of v is also a subspace. And let's say we have some members, or let me write it this way. If we have these two subspaces-- you have a subspace and you have this orthogonal complement-- we already learned that if you have any member of Rn-- so let's say that x is a member of our Rn-- then x can be represented as a sum of a member of v and a member of the orthogonal complement of v. Where-- let me write this-- the vector v is a member of the subspace v and the vector w is a member of the orthogonal complement of the subspace v. Just like that. We saw this several videos ago. We proved that this was true for any member Rn. Now given that, we can define the projection of x onto the subspace v as being equal to, just the part of x -- these are two orthogonal parts of x-- we define the projection onto v as a part of x that came from v. It's equal to just that vector v. Alternately you could say that the projection of x onto the orthogonal complement of-- sorry I wrote transpose-- the orthogonal complement of v is going to be equal to w. So this piece right here is a projection onto the subspace v. This piece right here is a projection onto the orthogonal complement of the subspace v. Now what I want to do in this video is show you that these two definitions-- that this definition right here which is then in conjunction with this right here-- this is the equivalent to what we learned up here if the subspace v that we're dealing with is a line. Because this was a valid subspace. But not all subspaces are going to be lines. And to see this we can revisit an example that we saw several videos ago. Several videos ago we had this matrix here A. This 2 by 2 matrix. And then we had this other vector b that was a member of the column space of A. We did this problem to show you that the shortest solution to this right here was a unique member of the row space. Hopefully that gets your memory on track for this problem when we first did it. But let me graph it and show you that for the solution of that problem we could have just as easily taken a projection onto a subspace. Let me graph everything in this problem. This might help you remember also about the problem. So let me draw my axes just like that. So the first thing we learned-- you know you could solve this but I already did this in a video. I think it was two or three videos ago-- the null space of A, or all of the x's that satisfy Ax is equal to zero, is a span of the vector 2, 3. So you go 2 to the right. 1, 2. And then you go 3 up. 1, 2, 3. And so it's the span of this vector. And so the span of that vector is just all the points. Well that vector specifies that point. But if you scale this vector up and down you're going to specify all of the point on this line. All the points on that line. Let me draw it like that. That's good enough. It shouldn't curve down like that at the end. So let me draw that a little straighter. So this is the null space. That is our null space of that matrix right there. And then the row space was a span of the vector 3, minus 2. You see that right here. 3, minus 2 is the first row. This guy is just a multiple of that one. That's why we don't have this guy right here in the span as well. And if we were to graph it, 3, minus 2. You go out 3, then you go down 1, 2. it would be the span of this vector right there. Let me draw it like that. Now you take all of the scalar multiples of that vector and you put those vectors in standard position. They're going to specify, or their tips are going to be on points along this line right there. Along that line right there. I'm trying to make sure I draw them orthogonally. So this right here is the row space. That right there is the row space of A which is the same thing as a column space of A transpose. And we know that these guys are each other's orthogonal complements. We know, we've seen this in multiple videos, that the null space of A is the orthogonal complement of the row space. And we also know that the orthogonal complement of the null space is equal to the row space. Everything in this is orthogonal to everything in that. Everything in that is orthogonal to everything in this. You can see it here in this graph. That these two spaces, which are represented by these lines that go through the origin, are orthogonal. And it makes sense that any-- we said at really the beginning of the video-- that anything in R2 in this situation, can be represented as some sum of a unique member of our row space and a unique member of its orthogonal complement. Let's say I have that point right there. How could I represent it as a sum of a member of this and a member of that? Well if I go along this guy, I have this vector right here. I have that vector right there along that line. And then I have this vector right here. If I were to shift it-- this is drawn in standard position, but I can draw a vector wherever I want. These lines are just all of the vectors drawn in standard position with their tails at the origin. But we learned, in really, I think, the first or second vector videos, that I can draw them wherever I want. So if I add this vector and that vector, I can shift this vector over and this vector will be right there. And there you have it. I took an arbitrary point in R2 and I can represent it as a sum of a member of my row space and a member of the row space's orthogonal complement or the null space. But just to review, what we originally did in that problem is we looked at the solution set of this. We said the solution set of this looks like this. It has a particular solution plus members of your null space, plus homogeneous solutions. We've seen that multiple videos ago. So 3, 0-- it looks like this-- plus members of the null space. So your solution set is going to be parallel to this but shifted to the right by 3. So it looks-- let me draw it a little neater than that-- Let me draw it like that. And then it goes down like, the second part I didn't draw-- there you go. Oh that's not good either. Maybe I'm being too picky. OK, so this is your solution set. And if you remember in that video we said, hey there's some member of this solution set that is also a member of our row space and that member of the solution set that is a member of our row space is going to be the shortest solution. And we saw that. You can see it visually right here. Right? This vector right here. It is in our row space. It is a member of our row space. And it also specifies a point on our solution set. And you could see visually that it's going to be the shortest solution. And one way you could think about it is, this is the projection-- let me pick a good, different, new color-- any solution on our solution set-- let me see right there-- let's say that that is some arbitrary solution on our solution set. Right? That's going to be a point in R2 and any point in R2 can be represented as a sum of some vector in our row space and some vector in our null space. And so if I have this vector right here, how can I do that? Well, I could represent it as a sum of this guy right here and then this vector right here. That vector right here. And this vector right here is clearly a member of my null space. I just shifted over. This line is only when I draw in standard position. This vector right here-- I'm just showing it heads to tails-- if I add this member of my row space to this member of my null space, I get an arbitrary solution to my solution set. And if you think about it, the projection of my arbitrary solution onto my row space will be this guy right here. And that just comes from our-- well there are two ways to think about it-- we could say that this is the solution right here. We could say our solution right here is equal to some member of my row space plus some member of my null space. This is the row space. That is the null space. And so by the definition of a projection onto a subspace I just gave you, we know that the projection of this solution onto my-- let me write a little bit-- onto my row space of my solution, is just equal to this first thing. It's equal to the component of it that's in my row space. It's other component, we could call it, is in the orthogonal complement of my row space. Or it's in my null space. So this is just going to be equal to the R vector. Now, I want to show you that that is essentially equal to the definition that we did before. That this is completely identical to the definition of a projection onto a line because in this case the subspace is a line. So let's find a solution set. And the easiest one, the easiest solution that we could find is if we set C as equal to 0 here. We know that x equals 3, 0 is one of these solutions. So x equals 3, 0 looks like that. So we know x equals 3, 0 is a solution. And what we want to do is we want to find the shortest solution. Or we want to find the projection of x onto the row space. Or if we wanted, we could also think of it is a projection of x onto this line. This line is equal to the row space. So let's do that. And I'm doing this to show you that this definition of a projection onto a subspace that I've just introduced you to in this video, it is completely identical to the definition, or it's not identical, it's consistent with the definition of a projection onto a line. Although this is more general because a subspace doesn't have to be a line. But in this case it is a line. So let's do that. So the projection of the vector 3, 0 onto our row space, which is a line so we can use that formula, it is equal to 3, 0 dot the spanning vector for our row space, right? Dot the spanning vector for our row space. So it's 3, minus 2. There's a bunch of spanning vectors for your row space. This is just the one we happened to pick. So dot 3, minus 2 all over the spanning vector dotted with itself. 3, minus 2 dot 3, minus 2. And then this is just going to be one big scalar and then we want to multiply that-- or essentially scale up-- our actual spanning vector by that. So this is a projection of this solution onto my row space, which should give me this vector right here. Because we're just taking a projection onto a line, because a row space in this subspace is a line. And so we used the linear projections that we first got introduced to, I think, when I first started doing linear transformations. So let's see this is 3 times 3 plus 0 times minus 2. This right here is equal to 9. This is 3 times 3 plus minus 2 times minus 2. So that's 9 plus 4. That's 13. So it's 9/13 times this vector right here. So it's going to be equal to 9/13 times the vector 3, minus 2. Which is equal to the vector 27/13 and then minus 18/13, which is this vector right here. We got this exact answer when we first did it, although we just didn't use the projection onto a line. But now we see that this is exactly consistent with what we did before. We just used the projection onto a line. And we see that this is consistent with our new, broader definition of a projection. Here we were able to do it because we did it onto a line. But here I'm calling a projection onto any subspace. We know how to do it if it's a line, but so far I've just kind of defined it onto an arbitrary subspace. But I haven't giving you a nice mathematical, I guess, or computational way to figure out what this is going to be if this isn't a line. In fact, I haven't even shown you when this is general, whether this is definitely a linear transformation. We know that when you take the projection onto the line it's a linear transformation. But I haven't shown you that when we take a projection onto an arbitrary subspace that it is a linear projection. I'll do that in the next video.

Definitions

A projection on a vector space is a linear operator such that .

When has an inner product and is complete, i.e. when is a Hilbert space, the concept of orthogonality can be used. A projection on a Hilbert space is called an orthogonal projection if it satisfies for all . A projection on a Hilbert space that is not orthogonal is called an oblique projection.

Projection matrix

  • A square matrix is called a projection matrix if it is equal to its square, i.e. if .[2]: p. 38 
  • A square matrix is called an orthogonal projection matrix if for a real matrix, and respectively for a complex matrix, where denotes the transpose of and denotes the adjoint or Hermitian transpose of .[2]: p. 223 
  • A projection matrix that is not an orthogonal projection matrix is called an oblique projection matrix.

The eigenvalues of a projection matrix must be 0 or 1.

Examples

Orthogonal projection

For example, the function which maps the point in three-dimensional space to the point is an orthogonal projection onto the xy-plane. This function is represented by the matrix

The action of this matrix on an arbitrary vector is

To see that is indeed a projection, i.e., , we compute

Observing that shows that the projection is an orthogonal projection.

Oblique projection

A simple example of a non-orthogonal (oblique) projection is

Via matrix multiplication, one sees that

showing that is indeed a projection.

The projection is orthogonal if and only if because only then

Properties and classification

The transformation T is the projection along k onto m. The range of T is m and the kernel is k.

Idempotence

By definition, a projection is idempotent (i.e. ).

Open map

Every projection is an open map, meaning that it maps each open set in the domain to an open set in the subspace topology of the image.[citation needed] That is, for any vector and any ball (with positive radius) centered on , there exists a ball (with positive radius) centered on that is wholly contained in the image .

Complementarity of image and kernel

Let be a finite-dimensional vector space and be a projection on . Suppose the subspaces and are the image and kernel of respectively. Then has the following properties:

  1. is the identity operator on :
  2. We have a direct sum . Every vector may be decomposed uniquely as with and , and where

The image and kernel of a projection are complementary, as are and . The operator is also a projection as the image and kernel of become the kernel and image of and vice versa. We say is a projection along onto (kernel/image) and is a projection along onto .

Spectrum

In infinite-dimensional vector spaces, the spectrum of a projection is contained in as

Only 0 or 1 can be an eigenvalue of a projection. This implies that an orthogonal projection is always a positive semi-definite matrix. In general, the corresponding eigenspaces are (respectively) the kernel and range of the projection. Decomposition of a vector space into direct sums is not unique. Therefore, given a subspace , there may be many projections whose range (or kernel) is .

If a projection is nontrivial it has minimal polynomial , which factors into distinct linear factors, and thus is diagonalizable.

Product of projections

The product of projections is not in general a projection, even if they are orthogonal. If two projections commute then their product is a projection, but the converse is false: the product of two non-commuting projections may be a projection.

If two orthogonal projections commute then their product is an orthogonal projection. If the product of two orthogonal projections is an orthogonal projection, then the two orthogonal projections commute (more generally: two self-adjoint endomorphisms commute if and only if their product is self-adjoint).

Orthogonal projections

When the vector space has an inner product and is complete (is a Hilbert space) the concept of orthogonality can be used. An orthogonal projection is a projection for which the range and the kernel are orthogonal subspaces. Thus, for every and in , . Equivalently:

A projection is orthogonal if and only if it is self-adjoint. Using the self-adjoint and idempotent properties of , for any and in we have , , and

where is the inner product associated with . Therefore, and are orthogonal projections.[3] The other direction, namely that if is orthogonal then it is self-adjoint, follows from the implication from to
for every and in ; thus .

The existence of an orthogonal projection onto a closed subspace follows from the Hilbert projection theorem.

Properties and special cases

An orthogonal projection is a bounded operator. This is because for every in the vector space we have, by the Cauchy–Schwarz inequality:

Thus .

For finite-dimensional complex or real vector spaces, the standard inner product can be substituted for .

Formulas

A simple case occurs when the orthogonal projection is onto a line. If is a unit vector on the line, then the projection is given by the outer product

(If is complex-valued, the transpose in the above equation is replaced by a Hermitian transpose). This operator leaves u invariant, and it annihilates all vectors orthogonal to , proving that it is indeed the orthogonal projection onto the line containing u.[4] A simple way to see this is to consider an arbitrary vector as the sum of a component on the line (i.e. the projected vector we seek) and another perpendicular to it, . Applying projection, we get
by the properties of the dot product of parallel and perpendicular vectors.

This formula can be generalized to orthogonal projections on a subspace of arbitrary dimension. Let be an orthonormal basis of the subspace , with the assumption that the integer , and let denote the matrix whose columns are , i.e., . Then the projection is given by:[5]

which can be rewritten as

The matrix is the partial isometry that vanishes on the orthogonal complement of , and is the isometry that embeds into the underlying vector space. The range of is therefore the final space of . It is also clear that is the identity operator on .

The orthonormality condition can also be dropped. If is a (not necessarily orthonormal) basis with , and is the matrix with these vectors as columns, then the projection is:[6][7]

The matrix still embeds into the underlying vector space but is no longer an isometry in general. The matrix is a "normalizing factor" that recovers the norm. For example, the rank-1 operator is not a projection if After dividing by we obtain the projection onto the subspace spanned by .

In the general case, we can have an arbitrary positive definite matrix defining an inner product , and the projection is given by . Then

When the range space of the projection is generated by a frame (i.e. the number of generators is greater than its dimension), the formula for the projection takes the form: . Here stands for the Moore–Penrose pseudoinverse. This is just one of many ways to construct the projection operator.

If is a non-singular matrix and (i.e., is the null space matrix of ),[8] the following holds:

If the orthogonal condition is enhanced to with non-singular, the following holds:

All these formulas also hold for complex inner product spaces, provided that the conjugate transpose is used instead of the transpose. Further details on sums of projectors can be found in Banerjee and Roy (2014).[9] Also see Banerjee (2004)[10] for application of sums of projectors in basic spherical trigonometry.

Oblique projections

The term oblique projections is sometimes used to refer to non-orthogonal projections. These projections are also used to represent spatial figures in two-dimensional drawings (see oblique projection), though not as frequently as orthogonal projections. Whereas calculating the fitted value of an ordinary least squares regression requires an orthogonal projection, calculating the fitted value of an instrumental variables regression requires an oblique projection.

A projection is defined by its kernel and the basis vectors used to characterize its range (which is a complement of the kernel). When these basis vectors are orthogonal to the kernel, then the projection is an orthogonal projection. When these basis vectors are not orthogonal to the kernel, the projection is an oblique projection, or just a projection.

A matrix representation formula for a nonzero projection operator

Let be a linear operator, such that and assume that is not the zero operator. Let the vectors form a basis for the range of , and assemble these vectors in the matrix . Therefore the integer , otherwise and is the zero operator. The range and the kernel are complementary spaces, so the kernel has dimension . It follows that the orthogonal complement of the kernel has dimension . Let form a basis for the orthogonal complement of the kernel of the projection, and assemble these vectors in the matrix . Then the projection (with the condition ) is given by

This expression generalizes the formula for orthogonal projections given above.[11][12] A standard proof of this expression is the following. For any vector in the vector space , we can decompose , where vector is in the image of , and vector So , and then is in the kernel of , which is the null space of In other words, the vector is in the column space of so for some dimension vector and the vector satisfies by the construction of . Put these conditions together, and we find a vector so that . Since matrices and are of full rank by their construction, the -matrix is invertible. So the equation gives the vector In this way, for any vector and hence .

In the case that is an orthogonal projection, we can take , and it follows that . By using this formula, one can easily check that . In general, if the vector space is over complex number field, one then uses the Hermitian transpose and has the formula . Recall that one can define the Moore–Penrose inverse of the matrix by since has full column rank, so .

Singular values

Note that is also an oblique projection. The singular values of and can be computed by an orthonormal basis of . Let be an orthonormal basis of and let be the orthogonal complement of . Denote the singular values of the matrix by the positive values . With this, the singular values for are:[13]

and the singular values for are
This implies that the largest singular values of and are equal, and thus that the matrix norm of the oblique projections are the same. However, the condition number satisfies the relation , and is therefore not necessarily equal.

Finding projection with an inner product

Let be a vector space (in this case a plane) spanned by orthogonal vectors . Let be a vector. One can define a projection of onto as

where repeated indices are summed over (Einstein sum notation). The vector can be written as an orthogonal sum such that . Note that is sometimes denoted as . There is a theorem in linear algebra that states that this is the smallest distance (the orthogonal distance) from to and is commonly used in areas such as machine learning.
y is being projected onto the vector space V.

Canonical forms

Any projection on a vector space of dimension over a field is a diagonalizable matrix, since its minimal polynomial divides , which splits into distinct linear factors. Thus there exists a basis in which has the form

where is the rank of . Here is the identity matrix of size , is the zero matrix of size , and is the direct sum operator. If the vector space is complex and equipped with an inner product, then there is an orthonormal basis in which the matrix of P is[14]

where . The integers and the real numbers are uniquely determined. Note that . The factor corresponds to the maximal invariant subspace on which acts as an orthogonal projection (so that P itself is orthogonal if and only if ) and the -blocks correspond to the oblique components.

Projections on normed vector spaces

When the underlying vector space is a (not necessarily finite-dimensional) normed vector space, analytic questions, irrelevant in the finite-dimensional case, need to be considered. Assume now is a Banach space.

Many of the algebraic results discussed above survive the passage to this context. A given direct sum decomposition of into complementary subspaces still specifies a projection, and vice versa. If is the direct sum , then the operator defined by is still a projection with range and kernel . It is also clear that . Conversely, if is projection on , i.e. , then it is easily verified that . In other words, is also a projection. The relation implies and is the direct sum .

However, in contrast to the finite-dimensional case, projections need not be continuous in general. If a subspace of is not closed in the norm topology, then the projection onto is not continuous. In other words, the range of a continuous projection must be a closed subspace. Furthermore, the kernel of a continuous projection (in fact, a continuous linear operator in general) is closed. Thus a continuous projection gives a decomposition of into two complementary closed subspaces: .

The converse holds also, with an additional assumption. Suppose is a closed subspace of . If there exists a closed subspace such that X = UV, then the projection with range and kernel is continuous. This follows from the closed graph theorem. Suppose xnx and Pxny. One needs to show that . Since is closed and {Pxn} ⊂ U, y lies in , i.e. Py = y. Also, xnPxn = (IP)xnxy. Because is closed and {(IP)xn} ⊂ V, we have , i.e. , which proves the claim.

The above argument makes use of the assumption that both and are closed. In general, given a closed subspace , there need not exist a complementary closed subspace , although for Hilbert spaces this can always be done by taking the orthogonal complement. For Banach spaces, a one-dimensional subspace always has a closed complementary subspace. This is an immediate consequence of Hahn–Banach theorem. Let be the linear span of . By Hahn–Banach, there exists a bounded linear functional such that φ(u) = 1. The operator satisfies , i.e. it is a projection. Boundedness of implies continuity of and therefore is a closed complementary subspace of .

Applications and further considerations

Projections (orthogonal and otherwise) play a major role in algorithms for certain linear algebra problems:

As stated above, projections are a special case of idempotents. Analytically, orthogonal projections are non-commutative generalizations of characteristic functions. Idempotents are used in classifying, for instance, semisimple algebras, while measure theory begins with considering characteristic functions of measurable sets. Therefore, as one can imagine, projections are very often encountered in the context of operator algebras. In particular, a von Neumann algebra is generated by its complete lattice of projections.

Generalizations

More generally, given a map between normed vector spaces one can analogously ask for this map to be an isometry on the orthogonal complement of the kernel: that be an isometry (compare Partial isometry); in particular it must be onto. The case of an orthogonal projection is when W is a subspace of V. In Riemannian geometry, this is used in the definition of a Riemannian submersion.

See also

Notes

  1. ^ Meyer, pp 386+387
  2. ^ a b Horn, Roger A.; Johnson, Charles R. (2013). Matrix Analysis, second edition. Cambridge University Press. ISBN 9780521839402.
  3. ^ Meyer, p. 433
  4. ^ Meyer, p. 431
  5. ^ Meyer, equation (5.13.4)
  6. ^ Banerjee, Sudipto; Roy, Anindya (2014), Linear Algebra and Matrix Analysis for Statistics, Texts in Statistical Science (1st ed.), Chapman and Hall/CRC, ISBN 978-1420095388
  7. ^ Meyer, equation (5.13.3)
  8. ^ See also Linear least squares (mathematics) § Properties of the least-squares estimators.
  9. ^ Banerjee, Sudipto; Roy, Anindya (2014), Linear Algebra and Matrix Analysis for Statistics, Texts in Statistical Science (1st ed.), Chapman and Hall/CRC, ISBN 978-1420095388
  10. ^ Banerjee, Sudipto (2004), "Revisiting Spherical Trigonometry with Orthogonal Projectors", The College Mathematics Journal, 35 (5): 375–381, doi:10.1080/07468342.2004.11922099, S2CID 122277398
  11. ^ Banerjee, Sudipto; Roy, Anindya (2014), Linear Algebra and Matrix Analysis for Statistics, Texts in Statistical Science (1st ed.), Chapman and Hall/CRC, ISBN 978-1420095388
  12. ^ Meyer, equation (7.10.39)
  13. ^ Brust, J. J.; Marcia, R. F.; Petra, C. G. (2020), "Computationally Efficient Decompositions of Oblique Projection Matrices", SIAM Journal on Matrix Analysis and Applications, 41 (2): 852–870, doi:10.1137/19M1288115, OSTI 1680061, S2CID 219921214
  14. ^ Doković, D. Ž. (August 1991). "Unitary similarity of projectors". Aequationes Mathematicae. 42 (1): 220–224. doi:10.1007/BF01818492. S2CID 122704926.

References

External links

This page was last edited on 28 April 2024, at 21:53
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.