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

Rank (linear algebra)

From Wikipedia, the free encyclopedia

In linear algebra, the rank of a matrix A is the dimension of the vector space generated (or spanned) by its columns.[1][2][3] This corresponds to the maximal number of linearly independent columns of A. This, in turn, is identical to the dimension of the vector space spanned by its rows.[4] Rank is thus a measure of the "nondegenerateness" of the system of linear equations and linear transformation encoded by A. There are multiple equivalent definitions of rank. A matrix's rank is one of its most fundamental characteristics.

The rank is commonly denoted by rank(A) or rk(A);[2] sometimes the parentheses are not written, as in rank A.[i]

YouTube Encyclopedic

  • 1/5
    Views:
    477 214
    2 495 628
    41 811
    41 738
    93 572
  • Dimension of the column space or rank | Vectors and spaces | Linear Algebra | Khan Academy
  • Inverse matrices, column space and null space | Chapter 7, Essence of linear algebra
  • Linear Algebra - 22 - Rank
  • [Linear Algebra] Row Space and The Rank Theorem
  • How to Find the Rank of a Matrix (with echelon form) | Linear Algebra

Transcription

We've seen in several videos that the column space of a matrix is pretty straightforward to find. In this situation the column space of A is just equal to all of the linear combinations of the column vectors of A. Another way of saying all of the linear combinations is just the span of each of these column vectors. So if we call this one right here a1. This is a2, a3, a4. This is a5. Then the column space of A is just equal to the span of a1, a2, a3, a4, and a5. Fair enough. But a more interesting question is whether these guys form a basis for the column space. Or even more interesting, what is the basis for the column space of A? And in this video I'm going to show you a method for determining the basis, and along the way we'll get an intuition for maybe why it works. And if I have time, actually I probably won't have time in this video. In the next video I'll prove to you why it works. So we want to figure out the basis for the column space of A. Remember the basis just means that vectors span, C, A. Clearly these vectors span our column space. I mean the span of these vectors is the column space. But in order to be a basis, the vectors also have to be linearly, let me just write, linearly independent. And we don't know whether these guys or what subset of these guys are linearly independent. So what you do-- and I'm just really going to describe the process here, as opposed to the proof-- is you put this guy in reduced row echelon form. So let's do that. So let me see if we can do that. Let's keep our first row the same. 1, 0. Let me do it actually in the right side right here. So let's keep the first row the same. 1, 0, minus 1, 0, 4. And then let's replace our second row with the second row minus 2 times the first row. So then our second row. 2 minus 2 times 1 is 0. 1 minus 2 times 0 is 1. 0 minus 2 times negative 1, so that's 0 plus 2. 0 minus 2 times 0 is just 0. And then 9 minus 2 times 4 is 1. Fair enough. Now we want to zero out this guy. Well it seems like a pretty straightforward way. Just replace this row with this row plus the first row. So minus 1 plus 1 is 0. 2 plus 0 is 2. 5 minus 1 is 4. 1 plus 0 is 1. Minus 5 plus 4 is minus 1. And then finally we got this guy right here, and in order to zero him out, let's replace him with him minus the first row. So 1 minus 1 is 0. Minus 1 minus 0 is minus 1. Minus 3 minus negative 1, that's minus 3 plus 1, so that's minus 2. Minus 2 minus 0 is minus 2. And then 9 minus 4 is 5. So we did one round. We got our first pivot column going. Now let's do another round of row operations. Well we want to zero all of these guys out. Luckily this is already 0. So we don't have to change our first row or our second row. So we get 1, 0, minus 1, 0, 4. Our second row becomes 0, 1, 2, 0, 1. And now let us see if we can eliminate this guy right here. And let's do it by replacing our blue row, our third row, with the third row minus 2 times the second row. So 0 minus 2 times 0 is 0. 2 minus 2 times 1 is 0. 4 minus 2 times 2 is 0. 1 minus 2 times 0 is 1. Minus 1 minus 2 times 1 is minus 3. All right. Now this last guy we want to eliminate him, and we want turn this into a 0. Let's replace this fourth row with the fourth row plus the second row. So 0 plus 0 is 0. Minus 1 plus minus 1 is 0. Minus 2 plus minus 2 is 0. Minus 2 plus 0 is minus 2. And then 5 plus 1 is 6. We're getting close. So let's look at our pivot entries. We have this is a pivot entry. That's a pivot entry. And this is not a pivot entry, because it's following obviously another. This guy is a pivot entry right here, or will be. Zero this minus 2 out, and I think we'll be done. So let me write my first row just the way it is, because everything above it is 0, so we don't have to worry about it. So my first row I can just write as 1, 0, minus 1, 0, 4. I can write my second row, 0, 1, 2, 0, 1. I can write my third row as 0, 0, 0, 1 minus 3. And now let's replace my fourth row. Let's replace it with it plus 2 times the second row. So 0 plus 2 times 0, 0 plus 2 times 0, 0 plus 2 times 0, minus 2 plus 2 times 1 is just 0. 6 plus 2 times minus 3, that's 6 minus 6, that's just 0. And there we've actually put our matrix in reduced row echelon form. So let me put brackets around it. It's not so bad if you just kind of go and just do the manipulations. And sometimes you kind of get a headache thinking about doing something like this, but this wasn't too bad. So this is let me just say the reduced row echelon form of A. Let me just call that matrix R. So this is matrix R right there. Now what do we see about matrix R? Well it has 3 pivot entries, or 3 pivot columns. Let me square them out, or circle them out. Column 1 is a pivot column, column 2 is a pivot column, and column 3 is a pivot column. And we've done this in previous videos. There's two things that you can see. These three columns are clearly linearly independent. How do we know that? And that's just with respect to each other. If we just took a set of, let's call this r1, r2, and this would be r3, this would be r4 right here. It's clear that the set r1, r2, and r4 is linearly independent. And you say why is that? Well look, our one's got a 1 here, while the other two have a 0 in that entry, right? And this is by definition of pivot entries. Pivot entries have 0's, or pivot columns have 0's everywhere except for where they have a 1. For any pivot column, it will be the only pivot column that has 0's there. Or it'll be the only pivot column that has a 1 there. So there's no way that you can add up combinations of these guys to get a 1. You can say 100 times 0, minus 3, times 0. You're just going to get a bunch of 0's. So no combination of these two guys is going to be equal to that guy. By the same reasoning, no combination of that and that is going to equal this. This is by definition of a pivot entry. When you put it in reduced row echelon form, it's very clear that any pivot column will be the only one to have 1 in that place. So it's very clear that these guys are linearly independent. Now it turns out, and I haven't proven it to you, that the corresponding columns in A-- this is r1, but it's A before we put it in reduced row echelon form-- that these guys right here, so a1, a2, and a4 are also linearly independent. So a1-- let me circle it-- a2, and a4. So if I write it like this, a1, a2, and a4. Let me write it in set notation. These guys are also linearly independant, which I haven't proven. But I think you can kind of get a sense that these row operations really don't change the sense of the matrix. And I'll do a better explanation of this, but I really just wanted you to understand how to develop a basis for the column space. So they're linearly independent. So the next question is do they span our column space? And in order for them to span, obviously all of these 5 vectors, if you have all of them, that's going to span your column space by definition. But if we can show, and I'm not going to show it in this video, but it turns out that you can always represent the non-pivot columns as linear combinations of the pivot columns. And we've kind of touched on that in previous videos where we find a solution for the null space and all that. So these guys can definitely be represented as linear combinations of these guys. I haven't shown you that, but if you take that on faith, then you don't need that column and that column to span. If you did then, or I guess a better way to think it, you don't need them to span, although they are part of the span. Because if you needed this guy, you can just construct him with linear combinations of these guys. So if you wanted to figure out a basis for the column space of A, you literally just take A into reduced row echelon form. You look at the pivot entries in the reduced row echelon form of A, and that's those three. And then you look at the corresponding columns to those pivot columns in your original A. And those form the basis. Because any linear combination of them, or linear combinations of them can be used to construct the non-pivot columns, and they're linearly independant. So I haven't shown you that. But for this case, if you want to know the basis, it's just a1, a2, and a4. And now we can answer another question. So a1, a2, and a4 form a basis for the column space of A, because you can construct the other two guys with linear combinations of our basis vectors, and they're also linearly independent. Now the next question is what is the dimension of the basis? Or what is the dimension-- not the dimension of the basis-- what is the dimension of the column space of A? Well the dimension is just the number of vectors in any basis for the column space. And all bases have the same number of vectors for any given subspace. So we have 1, 2, 3 vectors. So the dimension of our column space is equal to 3. And the dimension of a column space actually has a specific term for it, and that's called the rank. So the rank of A, which is the exact same thing as the dimension of the column space, it is equal to 3. And another way to think about it is, the rank of A is the number of linearly independent column vectors that you have that can span your entire column space. Or the number of linearly independent column vectors that can be used to construct all of the other column vectors. But hopefully this didn't confuse you too much, because the idea is very simple. Take A, put it into reduced row echelon form, see which columns are pivot columns. The corresponding columns are going to be a basis for your column space. If you want to know the rank for your matrix, you can just count them. Or if you don't want to count those, you could literally just count the number of pivot columns you have in your reduced row echelon form. So that's how you do it. In the next video I'll explain why this worked.

Main definitions

In this section, we give some definitions of the rank of a matrix. Many definitions are possible; see Alternative definitions for several of these.

The column rank of A is the dimension of the column space of A, while the row rank of A is the dimension of the row space of A.

A fundamental result in linear algebra is that the column rank and the row rank are always equal. (Three proofs of this result are given in § Proofs that column rank = row rank, below.) This number (i.e., the number of linearly independent rows or columns) is simply called the rank of A.

A matrix is said to have full rank if its rank equals the largest possible for a matrix of the same dimensions, which is the lesser of the number of rows and columns. A matrix is said to be rank-deficient if it does not have full rank. The rank deficiency of a matrix is the difference between the lesser of the number of rows and columns, and the rank.

The rank of a linear map or operator is defined as the dimension of its image:[5][6][7][8]

where is the dimension of a vector space, and is the image of a map.

Examples

The matrix

has rank 2: the first two columns are linearly independent, so the rank is at least 2, but since the third is a linear combination of the first two (the first column plus the second), the three columns are linearly dependent so the rank must be less than 3.

The matrix

has rank 1: there are nonzero columns, so the rank is positive, but any pair of columns is linearly dependent. Similarly, the transpose
of A has rank 1. Indeed, since the column vectors of A are the row vectors of the transpose of A, the statement that the column rank of a matrix equals its row rank is equivalent to the statement that the rank of a matrix is equal to the rank of its transpose, i.e., rank(A) = rank(AT).

Computing the rank of a matrix

Rank from row echelon forms

A common approach to finding the rank of a matrix is to reduce it to a simpler form, generally row echelon form, by elementary row operations. Row operations do not change the row space (hence do not change the row rank), and, being invertible, map the column space to an isomorphic space (hence do not change the column rank). Once in row echelon form, the rank is clearly the same for both row rank and column rank, and equals the number of pivots (or basic columns) and also the number of non-zero rows.

For example, the matrix A given by

can be put in reduced row-echelon form by using the following elementary row operations:
The final matrix (in reduced row echelon form) has two non-zero rows and thus the rank of matrix A is 2.

Computation

When applied to floating point computations on computers, basic Gaussian elimination (LU decomposition) can be unreliable, and a rank-revealing decomposition should be used instead. An effective alternative is the singular value decomposition (SVD), but there are other less computationally expensive choices, such as QR decomposition with pivoting (so-called rank-revealing QR factorization), which are still more numerically robust than Gaussian elimination. Numerical determination of rank requires a criterion for deciding when a value, such as a singular value from the SVD, should be treated as zero, a practical choice which depends on both the matrix and the application.

Proofs that column rank = row rank

Proof using row reduction

The fact that the column and row ranks of any matrix are equal forms is fundamental in linear algebra. Many proofs have been given. One of the most elementary ones has been sketched in § Rank from row echelon forms. Here is a variant of this proof:

It is straightforward to show that neither the row rank nor the column rank are changed by an elementary row operation. As Gaussian elimination proceeds by elementary row operations, the reduced row echelon form of a matrix has the same row rank and the same column rank as the original matrix. Further elementary column operations allow putting the matrix in the form of an identity matrix possibly bordered by rows and columns of zeros. Again, this changes neither the row rank nor the column rank. It is immediate that both the row and column ranks of this resulting matrix is the number of its nonzero entries.

We present two other proofs of this result. The first uses only basic properties of linear combinations of vectors, and is valid over any field. The proof is based upon Wardlaw (2005).[9] The second uses orthogonality and is valid for matrices over the real numbers; it is based upon Mackiw (1995).[4] Both proofs can be found in the book by Banerjee and Roy (2014).[10]

Proof using linear combinations

Let A be an m × n matrix. Let the column rank of A be r, and let c1, ..., cr be any basis for the column space of A. Place these as the columns of an m × r matrix C. Every column of A can be expressed as a linear combination of the r columns in C. This means that there is an r × n matrix R such that A = CR. R is the matrix whose ith column is formed from the coefficients giving the ith column of A as a linear combination of the r columns of C. In other words, R is the matrix which contains the multiples for the bases of the column space of A (which is C), which are then used to form A as a whole. Now, each row of A is given by a linear combination of the r rows of R. Therefore, the rows of R form a spanning set of the row space of A and, by the Steinitz exchange lemma, the row rank of A cannot exceed r. This proves that the row rank of A is less than or equal to the column rank of A. This result can be applied to any matrix, so apply the result to the transpose of A. Since the row rank of the transpose of A is the column rank of A and the column rank of the transpose of A is the row rank of A, this establishes the reverse inequality and we obtain the equality of the row rank and the column rank of A. (Also see Rank factorization.)

Proof using orthogonality

Let A be an m × n matrix with entries in the real numbers whose row rank is r. Therefore, the dimension of the row space of A is r. Let x1, x2, …, xr be a basis of the row space of A. We claim that the vectors Ax1, Ax2, …, Axr are linearly independent. To see why, consider a linear homogeneous relation involving these vectors with scalar coefficients c1, c2, …, cr:

where v = c1x1 + c2x2 + ⋯ + crxr. We make two observations: (a) v is a linear combination of vectors in the row space of A, which implies that v belongs to the row space of A, and (b) since Av = 0, the vector v is orthogonal to every row vector of A and, hence, is orthogonal to every vector in the row space of A. The facts (a) and (b) together imply that v is orthogonal to itself, which proves that v = 0 or, by the definition of v,
But recall that the xi were chosen as a basis of the row space of A and so are linearly independent. This implies that c1 = c2 = ⋯ = cr = 0. It follows that Ax1, Ax2, …, Axr are linearly independent.

Now, each Axi is obviously a vector in the column space of A. So, Ax1, Ax2, …, Axr is a set of r linearly independent vectors in the column space of A and, hence, the dimension of the column space of A (i.e., the column rank of A) must be at least as big as r. This proves that row rank of A is no larger than the column rank of A. Now apply this result to the transpose of A to get the reverse inequality and conclude as in the previous proof.

Alternative definitions

In all the definitions in this section, the matrix A is taken to be an m × n matrix over an arbitrary field F.

Dimension of image

Given the matrix , there is an associated linear mapping

defined by
The rank of is the dimension of the image of . This definition has the advantage that it can be applied to any linear map without need for a specific matrix.

Rank in terms of nullity

Given the same linear mapping f as above, the rank is n minus the dimension of the kernel of f. The rank–nullity theorem states that this definition is equivalent to the preceding one.

Column rank – dimension of column space

The rank of A is the maximal number of linearly independent columns of A; this is the dimension of the column space of A (the column space being the subspace of Fm generated by the columns of A, which is in fact just the image of the linear map f associated to A).

Row rank – dimension of row space

The rank of A is the maximal number of linearly independent rows of A; this is the dimension of the row space of A.

Decomposition rank

The rank of A is the smallest integer k such that A can be factored as , where C is an m × k matrix and R is a k × n matrix. In fact, for all integers k, the following are equivalent:

  1. the column rank of A is less than or equal to k,
  2. there exist k columns of size m such that every column of A is a linear combination of ,
  3. there exist an matrix C and a matrix R such that (when k is the rank, this is a rank factorization of A),
  4. there exist k rows of size n such that every row of A is a linear combination of ,
  5. the row rank of A is less than or equal to k.

Indeed, the following equivalences are obvious: . For example, to prove (3) from (2), take C to be the matrix whose columns are from (2). To prove (2) from (3), take to be the columns of C.

It follows from the equivalence that the row rank is equal to the column rank.

As in the case of the "dimension of image" characterization, this can be generalized to a definition of the rank of any linear map: the rank of a linear map f : VW is the minimal dimension k of an intermediate space X such that f can be written as the composition of a map VX and a map XW. Unfortunately, this definition does not suggest an efficient manner to compute the rank (for which it is better to use one of the alternative definitions). See rank factorization for details.

Rank in terms of singular values

The rank of A equals the number of non-zero singular values, which is the same as the number of non-zero diagonal elements in Σ in the singular value decomposition .

Determinantal rank – size of largest non-vanishing minor

The rank of A is the largest order of any non-zero minor in A. (The order of a minor is the side-length of the square sub-matrix of which it is the determinant.) Like the decomposition rank characterization, this does not give an efficient way of computing the rank, but it is useful theoretically: a single non-zero minor witnesses a lower bound (namely its order) for the rank of the matrix, which can be useful (for example) to prove that certain operations do not lower the rank of a matrix.

A non-vanishing p-minor (p × p submatrix with non-zero determinant) shows that the rows and columns of that submatrix are linearly independent, and thus those rows and columns of the full matrix are linearly independent (in the full matrix), so the row and column rank are at least as large as the determinantal rank; however, the converse is less straightforward. The equivalence of determinantal rank and column rank is a strengthening of the statement that if the span of n vectors has dimension p, then p of those vectors span the space (equivalently, that one can choose a spanning set that is a subset of the vectors): the equivalence implies that a subset of the rows and a subset of the columns simultaneously define an invertible submatrix (equivalently, if the span of n vectors has dimension p, then p of these vectors span the space and there is a set of p coordinates on which they are linearly independent).

Tensor rank – minimum number of simple tensors

The rank of A is the smallest number k such that A can be written as a sum of k rank 1 matrices, where a matrix is defined to have rank 1 if and only if it can be written as a nonzero product of a column vector c and a row vector r. This notion of rank is called tensor rank; it can be generalized in the separable models interpretation of the singular value decomposition.

Properties

We assume that A is an m × n matrix, and we define the linear map f by f(x) = Ax as above.

  • The rank of an m × n matrix is a nonnegative integer and cannot be greater than either m or n. That is,
    A matrix that has rank min(m, n) is said to have full rank; otherwise, the matrix is rank deficient.
  • Only a zero matrix has rank zero.
  • f is injective (or "one-to-one") if and only if A has rank n (in this case, we say that A has full column rank).
  • f is surjective (or "onto") if and only if A has rank m (in this case, we say that A has full row rank).
  • If A is a square matrix (i.e., m = n), then A is invertible if and only if A has rank n (that is, A has full rank).
  • If B is any n × k matrix, then
  • If B is an n × k matrix of rank n, then
  • If C is an l × m matrix of rank m, then
  • The rank of A is equal to r if and only if there exists an invertible m × m matrix X and an invertible n × n matrix Y such that
    where Ir denotes the r × r identity matrix.
  • Sylvester’s rank inequality: if A is an m × n matrix and B is n × k, then[ii]
    This is a special case of the next inequality.
  • The inequality due to Frobenius: if AB, ABC and BC are defined, then[iii]
  • Subadditivity:
    when A and B are of the same dimension. As a consequence, a rank-k matrix can be written as the sum of k rank-1 matrices, but not fewer.
  • The rank of a matrix plus the nullity of the matrix equals the number of columns of the matrix. (This is the rank–nullity theorem.)
  • If A is a matrix over the real numbers then the rank of A and the rank of its corresponding Gram matrix are equal. Thus, for real matrices
    This can be shown by proving equality of their null spaces. The null space of the Gram matrix is given by vectors x for which If this condition is fulfilled, we also have [11]
  • If A is a matrix over the complex numbers and denotes the complex conjugate of A and A the conjugate transpose of A (i.e., the adjoint of A), then

Applications

One useful application of calculating the rank of a matrix is the computation of the number of solutions of a system of linear equations. According to the Rouché–Capelli theorem, the system is inconsistent if the rank of the augmented matrix is greater than the rank of the coefficient matrix. If on the other hand, the ranks of these two matrices are equal, then the system must have at least one solution. The solution is unique if and only if the rank equals the number of variables. Otherwise the general solution has k free parameters where k is the difference between the number of variables and the rank. In this case (and assuming the system of equations is in the real or complex numbers) the system of equations has infinitely many solutions.

In control theory, the rank of a matrix can be used to determine whether a linear system is controllable, or observable.

In the field of communication complexity, the rank of the communication matrix of a function gives bounds on the amount of communication needed for two parties to compute the function.

Generalization

There are different generalizations of the concept of rank to matrices over arbitrary rings, where column rank, row rank, dimension of column space, and dimension of row space of a matrix may be different from the others or may not exist.

Thinking of matrices as tensors, the tensor rank generalizes to arbitrary tensors; for tensors of order greater than 2 (matrices are order 2 tensors), rank is very hard to compute, unlike for matrices.

There is a notion of rank for smooth maps between smooth manifolds. It is equal to the linear rank of the derivative.

Matrices as tensors

Matrix rank should not be confused with tensor order, which is called tensor rank. Tensor order is the number of indices required to write a tensor, and thus matrices all have tensor order 2. More precisely, matrices are tensors of type (1,1), having one row index and one column index, also called covariant order 1 and contravariant order 1; see Tensor (intrinsic definition) for details.

The tensor rank of a matrix can also mean the minimum number of simple tensors necessary to express the matrix as a linear combination, and that this definition does agree with matrix rank as here discussed.

See also

Notes

  1. ^ Alternative notation includes from Katznelson & Katznelson (2008, p. 52, §2.5.1) and Halmos (1974, p. 90, § 50).
  2. ^ Proof: Apply the rank–nullity theorem to the inequality
  3. ^ Proof. The map
    is well-defined and injective. We thus obtain the inequality in terms of dimensions of kernel, which can then be converted to the inequality in terms of ranks by the rank–nullity theorem. Alternatively, if is a linear subspace then ; apply this inequality to the subspace defined by the orthogonal complement of the image of in the image of , whose dimension is ; its image under has dimension .

References

  1. ^ Axler (2015) pp. 111-112, §§ 3.115, 3.119
  2. ^ a b Roman (2005) p. 48, § 1.16
  3. ^ Bourbaki, Algebra, ch. II, §10.12, p. 359
  4. ^ a b Mackiw, G. (1995), "A Note on the Equality of the Column and Row Rank of a Matrix", Mathematics Magazine, 68 (4): 285–286, doi:10.1080/0025570X.1995.11996337
  5. ^ Hefferon (2020) p. 200, ch. 3, Definition 2.1
  6. ^ Katznelson & Katznelson (2008) p. 52, § 2.5.1
  7. ^ Valenza (1993) p. 71, § 4.3
  8. ^ Halmos (1974) p. 90, § 50
  9. ^ Wardlaw, William P. (2005), "Row Rank Equals Column Rank", Mathematics Magazine, 78 (4): 316–318, doi:10.1080/0025570X.2005.11953349, S2CID 218542661
  10. ^ 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
  11. ^ Mirsky, Leonid (1955). An introduction to linear algebra. Dover Publications. ISBN 978-0-486-66434-7.

Sources

Further reading

  • Roger A. Horn and Charles R. Johnson (1985). Matrix Analysis. Cambridge University Press. ISBN 978-0-521-38632-6.
  • Kaw, Autar K. Two Chapters from the book Introduction to Matrix Algebra: 1. Vectors [1] and System of Equations [2]
  • Mike Brookes: Matrix Reference Manual. [3]
This page was last edited on 22 March 2024, at 09:07
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.