You are on page 1of 34

Matrix Methods

These tutorials were developed by Mr. Roaldi in 2000 to assist students in taking circuit courses. On May 23, 2011, there was an e-mail inquiry on matrix inversion. Unfortunately, the e-mail was accidentally purged. Basically, to determine values of N independent variables, one needs N linear equations. Coefficients of these equations form an NxN matrix. The determinant, D, of the matrix must not be zero. Otherwise, the matrix is singular and can't be inverted. Either there exist no consistent solutions or values of some independent variables can't be determined. The two main methods of circuit analysis presented in the book, mesh and nodal analysis, both result in a number of unknowns and the same number of equations. These equations are usually much easier to solve using the various properties of matrices than they would be by hand. This section will first show how the equations with the unknowns can be put into vector-matrix form and then three different ways to solve the matrix and find the unknowns.

Matrices
A matrix is a rectangular list of values with m rows and n columns (m and n may be different). The values in each row and ordered from left to right according to their column number, 1 to n. The values in each column are ordered from top to bottom according to their row numbers, 1 to m. Each value in the matrix has two subscripts, row then column. For example, a22 is the value in the second row and second column.

Matrix Multiplication
In order to use one of the three methods of solving the unknown variables using matrix methods, the equations must be put in vector-matrix form. To understand this form it is necessary to understand how matrices are multiplied. Only conformable matrices can be multiplied. In order to be conformable, the number of columns in the left matrix must equal the number of rows in the right matrix. Matrix multiplication is carried out in the following way. To find each element of the resulting matrix, you look at each of the rows of the first matrix and the corresponding column of the second matrix. You then take the leftmost element in the row of the first matrix and multiply it with the topmost element of the column of the second matrix. Then move right one in the row and down one in the column and find their product. For example, when multiplying two matrices, A * B, the first

step is to multiply a11 by b11. The next step is to multiply a12 by b21. Continue until no more rows or columns remain in the corresponding matrices (this will happen at the same time for each matrix if they are conformable). Add all of the products together to get the element's value. If A and B each had three rows and columns, the resulting matrix's first element would be r11 = (a11)(b11) + (a12)(b21) + (a13)(b31) Here is an example.

To aid in the multiplication, write the second matrix above and to the right of the first and the resulting matrix at the intersection of the two.

Now, to find the first element of the resulting matrix, r11, take the leftmost number in the corresponding row of the first matrix, 4, multiply it with the topmost number in the corresponding column of the second matrix, 1, and then add the product of the next number to the right in the first matrix and the next number down in the second matrix. For the top left element, it would be the following. (4)(1) + (3)(3) = 13

For the top-right element of the resulting matrix, we will still use row 1 of the first matrix but now use column 2 of the second matrix. This element is solved below. (4)(2) + (3)(1) = 11

Doing the same with the rest of the numbers leaves the following matrix.

Vector-Matrix Form
When mesh or nodal analysis is used to create equations which describe a circuit, the analysis results is some number of unknowns with an equal number of equations. It will look something like this. 1. 3a - b = 7 2. -a + 7b - 6c = 2 3. -3b + 4c = 0

A vector is simply a matrix with one column. The preceding equations can also be expressed in the following way:

Since multiplication is inferred by the lack of a sign, the form is usually shown like this.

This is simply an equation created using matrix multiplication. If you multiply the first matrix with the vector containing the unknowns, you can get the left side of the equations back, as shown here.

Since each equation states that the left side is equal to the right, it is valid to substitute a vector containing the right side of each equation for the side with the variables, and the vector-matrix form results.

Cramer's Rule
The first way to solve for the unknowns in a matrix is called Cramer's Rule. It is usually used if there are only two or three unknowns to solve. The first step is to find the determinant. The determinant is found by taking a square matrix (the first matrix in the vector-matrix form should always be a square) and finding the result when the products of the downward left diagonal numbers is subtracted from the products of the downward right diagonals. Since we can't use Greek letters, we will represent the determinant by D. Note: In matrices with 3 or more rows and columns, the diagonals will wrap around. Take the matrix from the last example.

First, find the sum of the downward right diagonals, then subtract the sum of the downward left diagonals.

D = (3)(7)(4) + (-1)(-6)(0) + (0)(-1)(-3) - (0)(7)(0) - (-3)(-6)(3) - (-1)(-1)(4) = 84 - 58 = 26

The determinant is 26. Next, to find each variable, replace the corresponding column with the solution vector (the vector on the right side of the equal sign in vector-matrix form) and perform the same operation used to find the determinant. Then divide the result by the determinant to find the answer. For example, to find a, the first variable in the example, replace column one with the solution vector and perform the diagonal multiplication.

Compute the determinant of this matrix

(7)(7)(4) + (-1)(-6)(0) + (0)(2)(-3) - (0)(7)(0) - (2)(-1)(4) (7)(-3)(-6) = 78

Then divide by the original determinant.

a = 78 / 26 = 3

The other two variables can be computed in a similar way.

b = ( (3)(2)(4) + (7)(-6)(0) + (0)(-1)(0) - (0)(2)(0) - (-1)(7)(4) (0)(-6)(3) ) / 26 = 52 / 26 = 2

c = ( (3)(7)(0) + (-1)(-3)(7) + (-1)(2)(0) - (0)(7)(7) - (-1)(-1)(0) (3)(-3)(2) ) / 26 = 39 / 26 = 3/2

Matrix Inversion
The second method of solving for the unknowns is called matrix inversion. The first step in this method is also finding the determinant of the original matrix. The next step involves finding the cofactors of all the matrix's elements. The cofactor is found by eliminating the row and column corresponding to the indices of the desired element and then finding the

determinant of the resulting matrix. For example, to find the cofactor of the first element, eliminate the first row and column and then find the determinant of the resulting matrix. There is also an additional step of multiplying the result by -1 raised to the sum of the indices. This basically means if the sum of the indices is even, the sign of the determinant is unchanged, if the sum is odd, the sign is changed. In finding the cofactor of the first row and first column the sign of the number would not be changed since 1+1 equals 2 and 2 is even. Taking the previous problem.

First, find the determinant of the original matrix. We already know this to be 26. Next find each cofactor.

C11 = -12 * ( (7)(4) - (-3)(-6) ) = (1) (10) = 10 Similarly, find the C12 by eliminating the first row and second column, finding the determinant of the resulting matrix and then multiplying by -1(1+2)

C12 = -13 * ( (-1)(4) - (0)(-6) ) = (-1) (-4) = 4 After finding all the cofactors, put them in a matrix according to their indices (C11 in the top left corner, C12, to its right and so on). The matrix of cofactors for this problem looks like this.

The next step is to find the transpose of the cofactor matrix. The transpose of a matrix is the matrix flipped around the main axis (the diagonal from top left to bottom right). The transpose of the cofactor matrix is:

Dividing this matrix by the determinant yields the following equation.

The unknowns can now be found by matrix multiplication and division by the original determinant, 26.

a = 78 / 26 = 3

b = 52 / 26 = 2 c = 39 / 26 = 3/2

These results agree with the results from Cranmer's Method.

Gauss Elimination
The final matrix method uses the basic idea of equations to solve for each unknown. Since one side of the equation is equal to the other (hence the term, equation) subtracting one equation from another must yield a third valid equation. The basic idea behind the gauss elimination method it to subtract the equations or multiples of the equations from each other until each equation has only one non-zero unknown. Here is an example.

First, we want to get rid of the first unknown in every row but the first row. The third row already has a zero there, but the second does not. Therefore, we will add 1/3 of the first row to the second row.

As you can see, 1/3 of 3 gave 1 which left us with 0 when added to the second row. 1/3 of -1 added to 7 gave 20/3, 1/3 of 0 added to -6 left 0, and 1/3 of 7 added to 6 gave 13/3. Next, we want to eliminate the second unknown in the last equation so that the last equation only has one unknown. This can be accomplished by adding 9/20 of the second row to the third.

Next, we want to find the actual value of the third unknown, so multiply the third row by 10/13 to simplify it.

From this we see that c is equal to 3/2. Next we need to get rid of the third unknown in the second row. Adding 6 times the 3rd row to the second will do this.

Multiply the second row by 3/20 to simplify.

Now add the first and second rows to get rid of the second unknown in the first row.

Finally, divide the first row by 3.

a = 3 b = 2 c = 3/2

These results agree with both the other matrix methods.

Determinant
In linear algebra, the determinant is a value associated with a square matrix. It can be computed from the entries of the matrix by a specific arithmetic expression, while other ways to determine its value exist as well. The determinant provides important information about a matrix of coefficients of a system of linear equations, or about a matrix that corresponds to a linear transformation of a vector space. In the first case the system has a unique solution exactly when the determinant is nonzero; when the determinant is zero there are either no solutions or many solutions. In the second case the transformation has an inverse operation exactly when the determinant is nonzero. A geometric interpretation can be given to the value of the determinant of a square matrix with real entries: the absolute value of the determinant gives the scale factor by which area or volume (or a higher dimensional analogue) is multiplied under the associated linear transformation, while its sign indicates whether the transformation preserves orientation. Thus a 2 2 matrix with determinant 2, when applied to a region of the plane with finite area, will transform that region into one with twice the area, while reversing its orientation. Determinants occur throughout mathematics. The use of determinants in calculus includes the Jacobian determinant in the substitution rule for integrals of functions of several variables. They are used to define the characteristic polynomial of a matrix that is an essential tool in eigenvalue problems in linear algebra. In some cases they are used just as a compact notation for expressions that would otherwise be unwieldy to write down. The determinant of a matrix A is denoted det(A), det A, or |A|.[1] In the case where the matrix entries are written out in full, the determinant is denoted by surrounding the matrix entries by vertical bars instead of the brackets or parentheses of the matrix. For instance, the determinant of the matrix

is written

and has the value

Although most often used for matrices whose entries are real or complex numbers, the definition of the determinant only involves addition, subtraction and multiplication, and so it can be defined for square matrices with entries taken from any commutative ring. Thus for instance the determinant of a matrix with integer coefficients will be an integer, and the matrix has an inverse with integer coefficients if and only if this determinant is 1 or 1 (these being the only invertible elements of the integers). For square matrices with entries in a non-commutative ring, for instance the quaternions, there is no unique definition for the determinant, and no definition that has all the usual properties of determinants over commutative rings. There are various ways to define the determinant of a square matrix A, i.e. one with the same number of rows and columns. Perhaps the most natural way is expressed in terms of the columns of the matrix. If we write an n n matrix in terms of its column vectors

where the so that

are vectors of size n, then the determinant of A is defined

where b and c are scalars, v is any vector of size n and I is the identity matrix of size n. These properties state that the determinant is an alternating multilinear function of the columns, and they suffice to uniquely calculate the determinant of any square matrix. Provided the underlying scalars form a field (more generally, a commutative ring with unity), the definition below shows that such a function exists, and it can be shown to be unique.[2]

Equivalently, the determinant can be expressed as a sum of products of entries of the matrix where each product has n terms and the coefficient of each product is 1 or 1 or 0 according to a given rule: it is a polynomial expression of the matrix entries. This expression grows rapidly with the size of the matrix (an n n matrix contributes n! terms), so it will first be given explicitly for the case of 22 matrices and 3 3 matrices, followed by the rule for arbitrary size matrices, which subsumes these two cases. Assume A is a square matrix with n rows and n columns, so that it can be written as

The entries can be numbers or expressions (as happens when the determinant is used to define a characteristic polynomial); the definition of the determinant depends only on the fact that they can be added and multiplied together in a commutative manner. The determinant of A is denoted as det(A), or it can be denoted directly in terms of the matrix entries by writing enclosing bars instead of brackets:

2 matrices

The area of the parallelogram is the absolute value of the determinant of the matrix formed by the vectors representing the parallelogram's sides. The determinant of a 2 2 matrix is defined by

If the matrix entries are real numbers, the matrix A can be used to represent two linear mappings: one that maps the standard basis vectors to the rows of A, and one that maps them to the columns of A. In either case, the images of the basis vectors form a parallelogram that represents the image of the unit square under the mapping. The parallelogram defined by the rows of the above matrix is the one with vertices at (0, 0), (a, b), (a + c, b + d), and (c, d), as shown in the accompanying diagram. The absolute value of ad bc is the area of the parallelogram, and thus represents the scale factor by which areas are transformed by A. (The parallelogram formed by the columns of A is in general a different parallelogram, but since the determinant is symmetric with respect to rows and columns, the area will be the same.) The absolute value of the determinant together with the sign becomes the oriented area of the parallelogram. The oriented area is the same as the usual area, except that it is negative when the angle from the first to the second vector defining the parallelogram turns in a clockwise direction (which is opposite to the direction one would get for the identity matrix).

Thus the determinant gives the scaling factor and the orientation induced by the mapping represented by A. When the determinant is equal to one, the linear mapping defined by the matrix is equi-areal and orientation-preserving. The object known as the bivector is related to these ideas. In 2d, it can be interpreted as an oriented plane segment formed by imagining two vectors each with origin (0, 0), and coordinates (a, b) and (c, d). The bivector magnitude (denoted (a, b) (c, d)) is the signed area, which is also the determinant ad bc.[3]

3 matrices

The volume of this Parallelepiped is the absolute value of the determinant of the matrix formed by the rows r1, r2, and r3. The determinant of a 33 matrix is defined by

The determinant of a 3

3 matrix can be calculated by its diagonals.

The rule of Sarrus is a mnemonic for this formula: the sum of the products of three diagonal north-west to south-east lines of matrix elements, minus the sum of the products of three diagonal south-west to north-east lines of elements, when the copies of the first two columns of the matrix are written beside it as in the illustration. For example, the determinant of

is calculated using this rule:

This scheme for calculating the determinant of a 3 carry over into higher dimensions.

3 matrix does not

n n matrices
The determinant of a matrix of arbitrary size can be defined by the Leibniz formula or the Laplace formula. The Leibniz formula for the determinant of an n

n matrix A is

Here the sum is computed over all permutations of the set {1, 2, ..., n}. A permutation is a function that reorders this set of integers. The value in the ith position after the reordering is denoted i. For example, for n = 3, the original sequence 1, 2, 3 might be reordered to = [2, 3, 1], with 1 = 2, 2 = 3, and 3 = 1. The set of all such permutations (also known as the symmetric group on n elements) is denoted Sn. For each permutation , sgn() denotes the signature of , a value that is +1 whenever the reordering given by can be achieved by successively interchanging two entries an even number of times, and 1 whenever it can be achieved by an odd number of such interchanges. In any of the summands, the term

is notation for the product of the entries at positions (i, i), where i ranges from 1 to n:

For example, the determinant of a 3

3 matrix A (n = 3) is

Levi-Civita symbol It is sometimes useful to extend the Leibniz formula to a summation in which not only permutations, but all sequences of n indices in the range

1,...,n occur, ensuring that the contribution of a sequence will be zero unless it denotes a permutation. Thus the totally antisymmetric Levi-Civita symbol extends the signature of a permutation, by setting for any permutation of n, and when no permutation exists such that for

(or equivalently, whenever some pair of indices is equal). The determinant for an n n-fold summation as

n matrix can then be expressed using an

Properties of the determinant


The determinant has many properties. Some basic properties of determinants are: 1. 2. where In is the n

n identity matrix

3. 4. For square matrices A and B of equal size,

5.

for an n

n matrix.

6. If A is a triangular matrix, i.e. ai,j = 0 whenever i > j or, alternatively, whenever i < j, then its determinant equals the product of the diagonal entries:

History
Historically, determinants were considered without reference to matrices: originally, a determinant was defined as a property of a system of linear equations. The determinant "determines" whether the system has a unique solution (which occurs precisely if the determinant is non-zero). In this sense, determinants were first used in the Chinese mathematics textbook The Nine Chapters on the Mathematical Art (, Chinese scholars, around the 3rd century BC). In Europe, 22 determinants were considered by Cardano at the end of the 16th century and larger ones by Leibniz. In Europe, Cramer (1750) added to the theory, treating the subject in relation to sets of equations. The recurrence law was first announced by Bzout (1764). It was Vandermonde (1771) who first recognized determinants as independent functions. Laplace (1772) gave the general method of expanding a determinant in terms of its complementary minors: Vandermonde had already given a special case. Immediately following, Lagrange (1773) treated determinants of the second and third order. Lagrange was the first to apply determinants to questions of elimination theory; he proved many special cases of general identities. Gauss (1801) made the next advance. Like Lagrange, he made much use of determinants in the theory of numbers. He introduced the word determinants (Laplace had used resultant), though not in the present signification, but rather as applied to the discriminant of a quantic. Gauss also arrived at the notion of reciprocal (inverse) determinants, and came very near the multiplication theorem. The next contributor of importance is Binet (1811, 1812), who formally stated the theorem relating to the product of two matrices of m columns and n rows, which for the special case of m = n reduces to the multiplication theorem. On the same day (November 30, 1812) that Binet presented his paper to the Academy, Cauchy also presented one on the subject. (See CauchyBinet formula.) In this he used the word determinant in its present sense, summarized and simplified what was then known on the subject, improved the notation, and gave the multiplication theorem with a proof more satisfactory than Binet's. With him begins the theory in its generality. The next important figure was Jacobi(from 1827). He early used the functional determinant which Sylvester later called the Jacobian, and in his memoirs in Crelle for 1841 he specially treats this subject, as well

as the class of alternating functions which Sylvester has called alternants. About the time of Jacobi's last memoirs, Sylvester (1839) and Cayley began their work. The study of special forms of determinants has been the natural result of the completion of the general theory. Axisymmetric determinants have been studied by Lebesgue, Hesse, and Sylvester; persymmetric determinants by Sylvester and Hankel; circulants by Catalan, Spottiswoode, Glaisher, and Scott; skew determinants and Pfaffians, in connection with the theory of orthogonal transformation, by Cayley; continuants by Sylvester; Wronskians (so called by Muir) by Christoffel and Frobenius; compound determinants by Sylvester, Reiss, and Picquet; Jacobians and Hessians by Sylvester; and symmetric gauche determinants by Trudi. Of the textbooks on the subject Spottiswoode's was the first. In America, Hanus (1886), Weld (1893), and Muir/Metzler (1933) published treatises.

Applications
Linear independence
As mentioned above, the determinant of a matrix (with real or complex entries, say) is zero if and only if the column vectors of the matrix are linearly dependent. Thus, determinants can be used to characterize linearly dependent vectors. For example, given two linearly independent vectors v1, v2 in R3, a third vector v3 lies in the plane spanned by the former two vectors exactly if the determinant of the 3 3 matrix consisting of the three vectors is zero. The same idea is also used in the theory of differential equations: given n functions f1(x), ..., fn(x) (supposed to be n1 times differentiable), the Wronskian is defined to be

It is non-zero (for some x) in a specified interval if and only if the given functions and all their derivatives up to order n1 are linearly independent. If it can be shown that the Wronskian is zero everywhere on an interval then, in the case of analytic functions, this implies the given

functions are linearly dependent. See the Wronskian and linear independence.

Orientation of a basis
Main article: Orientation (vector space) The determinant can be thought of as assigning a number to every sequence of n vectors in Rn, by using the square matrix whose columns are the given vectors. For instance, an orthogonal matrix with entries in Rn represents an orthonormal basis in Euclidean space. The determinant of such a matrix determines whether the orientation of the basis is consistent with or opposite to the orientation of the standard basis. If the determinant is +1, the basis has the same orientation. If it is 1, the basis has the opposite orientation. More generally, if the determinant of A is positive, A represents an orientation-preserving linear transformation (if A is an orthogonal 22 or 3 3 matrix, this is a rotation), while if it is negative, A switches the orientation of the basis.

Volume and Jacobian determinant


As pointed out above, the absolute value of the determinant of real vectors is equal to the volume of the parallelepiped spanned by those vectors. As a consequence, if f: Rn Rn is the linear map represented by the matrix A, and S is any measurable subset of Rn, then the volume of f(S) is given by |det(A)| times the volume of S. More generally, if the linear map f: Rn Rm is represented by the m n matrix A, then the n-dimensional volume of f(S) is given by:

By calculating the volume of the tetrahedron bounded by four points, they can be used to identify skew lines. The volume of any tetrahedron, given its vertices a, b, c, and d, is (1/6)|det(a b, b c, c d)|, or any other combination of pairs of vertices that would form a spanning tree over the vertices. For a general differentiable function, much of the above carries over by considering the Jacobian matrix of f. For

the Jacobian is the n

n matrix whose entries are given by

Its determinant, the Jacobian determinant appears in the higher-dimensional version of integration by substitution: for suitable functions f and an open subset U of R'n (the domain of f), the integral over f(U) of some other function : Rn Rm is given by

The Jacobian also occurs in the inverse function theorem.

Vandermonde determinant (alternant)


Main article: Vandermonde matrix Third order

In general, the nth-order Vandermonde determinant is

where the right-hand side is the continued product of all the differences that can be formed from the n(n1)/2 pairs of numbers taken from x1, x2, ..., xn, with the order of the differences taken in the reversed order of the suffixes that are involved.

Circulants
Main article: Circulant matrix Second order

Third order

where and 2 are the complex cube roots of 1. In general, the nth-order circulant determinant is

where j is an nth root of 1.

Introduction to Matrices / Matrix Size

Augmented matrices
Matrices are incredibly useful things that crop up in many different applied areas. For now, you'll probably only do some elementary manipulations with matrices, and then you'll move on to the next topic. But you should not be surprised to encounter matrices again in, say, physics or engineering. (The plural "matrices" is pronounced as "MAY-truh-seez".) Matrices were initially based on systems of linear equations. Given the following system of equations, write the associated augmented matrix.

2x + 3y z = 6 x y z = 9 x + y + 6z = 0

Write down the coefficients and the answer values, including all "minus" signs. If there is "no" coefficient, then the coefficient is "1".

That is, given a system of (linear) equations, you can relate to it the matrix (the grid of numbers inside the brackets) which contains only the coefficients of the linear system. This is called "an augmented matrix": the grid containing the coefficients from the left-hand side of each equation has been "augmented" with the answers from the right-hand side of each equation. The entries of (that is, the values in) the matrix correspond to the x-, y- and z-values in the original system, as long as the original system is arranged properly in the first place. Sometimes, you'll need to rearrange terms or insert zeroes as place-holders in your matrix. Given the following system of equations, write the associated augmented matrix.

x+y=0 y+z=3 zx=2


I first need to rearrange the system as:

x+y=0 y+z=3 -x + z = 2
Then I can write the associated matrix as:

When forming the augmented matrix, use a zero for any entry where the corresponding spot in the system of linear equations is blank.

Coefficient matrices
If you form the matrix only from the coefficient values, the matrix would look like this:

This is called "the coefficient matrix". 2003-2011 All Rights Reserved Above, we went from a linear system to an augmented matrix. You can go the other way, too. Given the following augmented matrix, write the associated linear system.

Remember that matrices require that the variables be all lined up nice and neat. And it is customary, when you have three variables, to use x, y, and z, in that order. So the associated linear system must be:

x + 3y = 4 2y z = 5 3x + z = 2

The Size of a matrix


Matrices are often referred to by their sizes. The size of a matrix is given in the form of a dimension, much as a room might be referred to as "a ten-by-twelve room". The dimensions for a matrix are the rows and columns, rather than the width and length. For instance, consider the following matrix A:

Since A has three rows and four columns, the size of A is 3 "three-by-four").

4 (pronounced as

The rows go side to side; the columns go up and down. "Row" and "column" are technical terms, and are not interchangable. Matrix dimensions are always given with the number of rows first, followed by the number of columns. Following this convention, the following matrix

B:

...is 2 3. If the matrix has the same number of rows as columns, the matrix is said to be a "square" matrix. For instance, the coefficient matrix from above:

...is a 3

3 square matrix.

Types of matrices
Sometimes matrices are categorized according to the configurations of their entries. For instance, a matrix like this one, with all-zero entries below the top-left-to-lower-right diagonal ("the diagonal") is called "upper triangular". (You can have lower triangular matrices, too, but they aren't of much use, so "triangular", without the "upper" or "lower", is generally taken to mean "upper triangular".)

A matrix with non-zero entries only on the diagonal is called "diagonal".

A diagonal matrix whose non-zero entries are all 1's is called an "identity" matrix, for reasons which will become clear when you learn how to multiply matrices.

There are many identity matrices. The previous example was the 3

3 identity; this is the 4 4 identity:

The 3

3 identity is denoted by I3 (pronounced as "eye-three" or "eye-sub-three"); similarly, the 4 4 identity is I4 and the 2 2 identity matrix is I2:

Note that triangular matrices are square, that diagonals are triangular and therefore are square, and that identities are diagonals and therefore are triangular and square. When describing a matrix, you usually just give its most specific classification, as this implies all the others. Classify the following matrix:

This isn't a square matrix, so it can't be an identity or anything. This is just a...

3 4 matrix.
Classify the following matrix:

This is a square matrix, but there isn't anything else special about it.

3 3 square matrix
Classify the following matrix:

This is a 3

3 upper triangular matrix, but it is not diagonal.

3 3 upper triangular matrix


Note: In a triangular matrix, you can have additional zeroes on or above the diagonal. Classify the following matrix:

This is a diagonal matrix, and, more than that, the diagonal entries are all 1's. Then this is... the 3

3 identity, I3.

Since identity matrices are, by definition, square matrices, you only need to use one subscript to give their dimensions.

Did you know . . .?

The history of matrices goes back to ancient times! But the term "matrix" was not applied to the concept until 1850.

"Matrix" is the Latin word for womb, and it retains that sense in English. It can also mean more generally any place in which something is formed or produced.

The orgins of mathematical matrices lie with the study of systems of simultaneous linear equations. An important Chinese text from between 300 BC and AD 200, Nine Chapters of the Mathematical Art (Chiu Chang Suan Shu),

gives the first known example of the use of matrix methods to solve simultaneous equations. In the treatise's seventh chapter, "Too much and not enough," the concept of a determinant first appears, nearly two millennia before its supposed invention by the Japanese mathematician Seki Kowa in 1683 or his German contemporary Gottfried Leibnitz (who is also credited with the invention of differential calculus, separately from but simultaneously with Isaac Newton). More uses of matrix-like arrangements of numbers appear in chapter eight, "Methods of rectangular arrays," in which a method is given for solving simultaneous equations using a counting board that is mathematically identical to the modern matrix method of solution outlined by Carl Friedrich Gauss (1777-1855), also known as Gaussian elimination.

The term "matrix" for such arrangements was introduced in 1850 by James Joseph Sylvester. Sylvester, incidentally, had a (very) brief career at the University of Virginia, which came to an abrupt end after an enraged Sylvester hit a newspaper-reading student with a sword stick and fled the country, believing he had killed the student!

James Joseph Sylvester.

Since their first appearance in ancient China, matrices have remained important mathematical tools. Today, they are used not simply for solving systems of simultaneous linear equations, but also for describing the quantum mechanics of atomic structure, designing computer game graphics, analyzing relationships, and even plotting complicated dance steps! The elevation of the matrix from mere tool to important mathematical theory owes a lot to the work of female mathematician Olga Taussky Todd (1906-1995), who began by using matrices to analyze vibrations on airplanes during World War II and became the torchbearer for matrix theory.

You might also like