Rank–nullity theorem
   HOME

TheInfoList



OR:

The rank–nullity theorem is a theorem in
linear algebra Linear algebra is the branch of mathematics concerning linear equations such as: :a_1x_1+\cdots +a_nx_n=b, linear maps such as: :(x_1, \ldots, x_n) \mapsto a_1x_1+\cdots +a_nx_n, and their representations in vector spaces and through matrices. ...
, which asserts that the
dimension In physics and mathematics, the dimension of a Space (mathematics), mathematical space (or object) is informally defined as the minimum number of coordinates needed to specify any Point (geometry), point within it. Thus, a Line (geometry), lin ...
of the
domain Domain may refer to: Mathematics *Domain of a function, the set of input values for which the (total) function is defined **Domain of definition of a partial function **Natural domain of a partial function **Domain of holomorphy of a function * Do ...
of a
linear map In mathematics, and more specifically in linear algebra, a linear map (also called a linear mapping, linear transformation, vector space homomorphism, or in some contexts linear function) is a Map (mathematics), mapping V \to W between two vect ...
is the sum of its
rank Rank is the relative position, value, worth, complexity, power, importance, authority, level, etc. of a person or object within a ranking, such as: Level or position in a hierarchical organization * Academic rank * Diplomatic rank * Hierarchy * ...
(the dimension of its
image An image is a visual representation of something. It can be two-dimensional, three-dimensional, or somehow otherwise feed into the visual system to convey information. An image can be an artifact, such as a photograph or other two-dimensiona ...
) and its ''nullity'' (the dimension of its
kernel Kernel may refer to: Computing * Kernel (operating system), the central component of most operating systems * Kernel (image processing), a matrix used for image convolution * Compute kernel, in GPGPU programming * Kernel method, in machine learn ...
). p. 70, §2.1, Theorem 2.3


Stating the theorem

Let T : V \to W be a linear transformation between two vector spaces where T's domain V is finite dimensional. Then \operatorname(T) ~+~ \operatorname(T) ~=~ \dim V, where \operatorname(T) ~:=~ \dim(\operatorname(T)) \qquad \text \qquad \operatorname(T) ~:=~ \dim(\operatorname (T)). In other words, \dim (\operatorname T) + \dim (\ker T) = \dim (\operatorname T). This theorem can be refined via the
splitting lemma In mathematics, and more specifically in homological algebra, the splitting lemma states that in any abelian category, the following statements are equivalent for a short exact sequence : 0 \longrightarrow A \mathrel B \mathrel C \longrightarro ...
to be a statement about an
isomorphism In mathematics, an isomorphism is a structure-preserving mapping between two structures of the same type that can be reversed by an inverse mapping. Two mathematical structures are isomorphic if an isomorphism exists between them. The word is ...
of spaces, not just dimensions. Explicitly, since induces an isomorphism from V / \operatorname (T) to \operatorname (T), the existence of a basis for that extends any given basis of \operatorname(T) implies, via the splitting lemma, that \operatorname(T) \oplus \operatorname(T) \cong V. Taking dimensions, the rank–nullity theorem follows.


Matrices

Since \operatorname_(\mathbb) \cong \operatorname\left(\mathbb^n, \mathbb^m\right), one can represent linear maps as
matrices Matrix most commonly refers to: * ''The Matrix'' (franchise), an American media franchise ** ''The Matrix'', a 1999 science-fiction action film ** "The Matrix", a fictional setting, a virtual reality environment, within ''The Matrix'' (franchis ...
. In the case of an m \times n matrix, the dimension of the domain is n, the number of columns in the matrix. Thus the rank–nullity theorem for a given matrix M \in \operatorname_(\mathbb) immediately becomes \operatorname(M) + \operatorname(M) = n.


Proofs

Here we provide two proofs. The first operates in the general case, using linear maps. The second proof looks at the homogeneous system \mathbf = \mathbf for \mathbf \in \operatorname_(\mathbb) with
rank Rank is the relative position, value, worth, complexity, power, importance, authority, level, etc. of a person or object within a ranking, such as: Level or position in a hierarchical organization * Academic rank * Diplomatic rank * Hierarchy * ...
r and shows explicitly that there exists a set of n-r
linearly independent In the theory of vector spaces, a set of vectors is said to be if there is a nontrivial linear combination of the vectors that equals the zero vector. If no such linear combination exists, then the vectors are said to be . These concepts are ...
solutions that span the kernel of \mathbf. While the theorem requires that the domain of the linear map be finite-dimensional, there is no such assumption on the codomain. This means that there are linear maps not given by matrices for which the theorem applies. Despite this, the first proof is not actually more general than the second: since the image of the linear map is finite-dimensional, we can represent the map from its domain to its image by a matrix, prove the theorem for that matrix, then compose with the inclusion of the image into the full codomain.


First proof

Let V, W be vector spaces over some field \mathbb and T defined as in the statement of the theorem with \dim V = n. As \operatornameT \subset V is a subspace, there exists a basis for it. Suppose \dim\operatornameT = k and let \mathcal := \ \subset \operatorname(T) be such a basis. We may now, by the
Steinitz exchange lemma The Steinitz exchange lemma is a basic theorem in linear algebra used, for example, to show that any two bases for a finite-dimensional vector space have the same number of elements. The result is named after the German mathematician Ernst Steinit ...
, extend \mathcal with n-k linearly independent vectors w_1, \ldots, w_ to form a full basis of V. Let \mathcal := \ \subset V \setminus \operatorname(T) such that \mathcal := \mathcal \cup \mathcal = \ \subset V is a basis for V. From this, we know that \operatorname T = \operatornameT(\mathcal) = \operatorname\ = \operatorname\ = \operatornameT(\mathcal) . We now claim that T(\mathcal) is a basis for \operatorname T. The above equality already states that T(\mathcal) is a generating set for \operatorname T; it remains to be shown that it is also linearly independent to conclude that it is a basis. Suppose T(\mathcal) is not linearly independent, and let \sum_^ \alpha _j T(w_j) = 0_W for some \alpha _j \in \mathbb. Thus, owing to the linearity of T, it follows that T \left(\sum_^ \alpha _j w_j \right) = 0_W \implies \left(\sum_^ \alpha _j w_j \right) \in \operatorname T = \operatorname \mathcal \subset V . This is a contradiction to \mathcal being a basis, unless all \alpha _j are equal to zero. This shows that T(\mathcal) is linearly independent, and more specifically that it is a basis for \operatornameT. To summarize, we have \mathcal, a basis for \operatornameT, and T(\mathcal), a basis for \operatornameT. Finally we may state that \operatorname(T) + \operatorname(T) = \dim \operatorname T + \dim \operatornameT = , T(\mathcal), + , \mathcal, = (n-k) + k = n = \dim V . This concludes our proof.


Second proof

Let \mathbf \in \operatorname_(\mathbb) with r
linearly independent In the theory of vector spaces, a set of vectors is said to be if there is a nontrivial linear combination of the vectors that equals the zero vector. If no such linear combination exists, then the vectors are said to be . These concepts are ...
columns (i.e. \operatorname(\mathbf) = r). We will show that: To do this, we will produce a matrix \mathbf \in \operatorname_(\mathbb) whose columns form a
basis Basis may refer to: Finance and accounting * Adjusted basis, the net cost of an asset after adjusting for various tax-related items *Basis point, 0.01%, often used in the context of interest rates * Basis trading, a trading strategy consisting ...
of the null space of \mathbf. Without loss of generality, assume that the first r columns of \mathbf are linearly independent. So, we can write \mathbf = \begin \mathbf_1 & \mathbf_2\end , where *\mathbf_1 \in \operatorname_(\mathbb) with r linearly independent column vectors, and *\mathbf_2 \in \operatorname_(\mathbb), each of whose n-r columns are linear combinations of the columns of \mathbf_1. This means that \mathbf_2 = \mathbf_1\mathbf for some \mathbf \in \operatorname _ (see
rank factorization In mathematics, given a field \mathbb F, nonnegative integers m,n, and a matrix A\in\mathbb F^, a rank decomposition or rank factorization of is a factorization of of the form , where C\in\mathbb F^ and F\in\mathbb F^, where r=\operatorname A is ...
) and, hence, \mathbf = \begin \mathbf_1 & \mathbf_1\mathbf\end . Let \mathbf = \begin -\mathbf \\ \mathbf_ \end , where \mathbf_ is the (n-r)\times (n-r)
identity matrix In linear algebra, the identity matrix of size n is the n\times n square matrix with ones on the main diagonal and zeros elsewhere. Terminology and notation The identity matrix is often denoted by I_n, or simply by I if the size is immaterial o ...
. We note that \mathbf \in \operatorname_(\mathbb) satisfies \mathbf\mathbf = \begin\mathbf_1 & \mathbf_1\mathbf \end\begin -\mathbf \\ \mathbf_ \end = -\mathbf_1\mathbf + \mathbf_1\mathbf = \mathbf_. Therefore, each of the n-r columns of \mathbf are particular solutions of \mathbf = \mathbf_. Furthermore, the n-r columns of \mathbf are
linearly independent In the theory of vector spaces, a set of vectors is said to be if there is a nontrivial linear combination of the vectors that equals the zero vector. If no such linear combination exists, then the vectors are said to be . These concepts are ...
because \mathbf = \mathbf_ will imply \mathbf = \mathbf_ for \mathbf \in \mathbb^: \mathbf\mathbf = \mathbf_ \implies \begin -\mathbf \\ \mathbf_ \end\mathbf = \mathbf_ \implies \begin -\mathbf\mathbf \\ \mathbf \end = \begin \mathbf_ \\ \mathbf_ \end \implies \mathbf = \mathbf_. Therefore, the column vectors of \mathbf constitute a set of n-r linearly independent solutions for \mathbf = \mathbf_. We next prove that ''any'' solution of \mathbf = \mathbf_ must be a linear combination of the columns of \mathbf. For this, let \mathbf = \begin \mathbf_1 \\ \mathbf_2 \end \in \mathbb^ be any vector such that \mathbf = \mathbf_. Note that since the columns of \mathbf_1 are linearly independent, \mathbf_1\mathbf = \mathbf_ implies \mathbf = \mathbf_. Therefore, \begin \mathbf\mathbf & = & \mathbf_ \\ \implies \begin\mathbf_1 & \mathbf_1\mathbf\end \begin \mathbf_1 \\ \mathbf_2 \end & = & \mathbf_1\mathbf_1 + \mathbf_1\mathbf\mathbf_2 & = & \mathbf_1(\mathbf_1 + \mathbf\mathbf_2) & = & \mathbf_ \\ \implies \mathbf_1 + \mathbf\mathbf_2 & = & \mathbf_ \\ \implies \mathbf_1 & = & -\mathbf\mathbf_2 \end \implies \mathbf = \begin \mathbf_1 \\ \mathbf_2 \end = \begin -\mathbf \\ \mathbf_ \end\mathbf_2 = \mathbf\mathbf_2. This proves that any vector \mathbf that is a solution of \mathbf = \mathbf must be a linear combination of the n-r special solutions given by the columns of \mathbf. And we have already seen that the columns of \mathbf are linearly independent. Hence, the columns of \mathbf constitute a basis for the
null space In mathematics, the kernel of a linear map, also known as the null space or nullspace, is the linear subspace of the Domain of a function, domain of the map which is mapped to the zero vector. That is, given a linear map between two vector space ...
of \mathbf. Therefore, the nullity of \mathbf is n - r. Since r equals rank of \mathbf, it follows that \operatorname(\mathbf) + \operatorname(\mathbf) = n. This concludes our proof.


Reformulations and generalizations

This theorem is a statement of the
first isomorphism theorem In mathematics, specifically abstract algebra, the isomorphism theorems (also known as Noether's isomorphism theorems) are theorems that describe the relationship between quotients, homomorphisms, and subobjects. Versions of the theorems exist fo ...
of algebra for the case of vector spaces; it generalizes to the
splitting lemma In mathematics, and more specifically in homological algebra, the splitting lemma states that in any abelian category, the following statements are equivalent for a short exact sequence : 0 \longrightarrow A \mathrel B \mathrel C \longrightarro ...
. In more modern language, the theorem can also be phrased as saying that each short exact sequence of vector spaces splits. Explicitly, given that 0 \rightarrow U \rightarrow V \mathbin R \rightarrow 0 is a
short exact sequence An exact sequence is a sequence of morphisms between objects (for example, groups, rings, modules, and, more generally, objects of an abelian category) such that the image of one morphism equals the kernel of the next. Definition In the context ...
of vector spaces, then U \oplus R \cong V , hence \dim(U) + \dim(R) = \dim(V) . Here ''R'' plays the role of im ''T'' and ''U'' is ker ''T'', i.e. 0 \rightarrow \ker T \mathbin V \mathbin \operatorname T \rightarrow 0 In the finite-dimensional case, this formulation is susceptible to a generalization: if is an
exact sequence An exact sequence is a sequence of morphisms between objects (for example, groups, rings, modules, and, more generally, objects of an abelian category) such that the image of one morphism equals the kernel of the next. Definition In the context ...
of finite-dimensional vector spaces, then \sum_^r (-1)^i\dim(V_i) = 0. The rank–nullity theorem for finite-dimensional vector spaces may also be formulated in terms of the ''index'' of a linear map. The index of a linear map T \in \operatorname(V,W), where V and W are finite-dimensional, is defined by \operatorname T = \dim \operatorname(T) - \dim \operatorname T . Intuitively, \dim \operatorname T is the number of independent solutions v of the equation Tv = 0, and \dim \operatorname T is the number of independent restrictions that have to be put on w to make Tv = w solvable. The rank–nullity theorem for finite-dimensional vector spaces is equivalent to the statement \operatorname T = \dim V - \dim W . We see that we can easily read off the index of the linear map T from the involved spaces, without any need to analyze T in detail. This effect also occurs in a much deeper result: the Atiyah–Singer index theorem states that the index of certain differential operators can be read off the geometry of the involved spaces.


Citations


References

* * * *. * * {{DEFAULTSORT:Rank-nullity theorem Theorems in linear algebra Isomorphism theorems Articles containing proofs