Kernel of a linear map – Serlo

The kernel of a linear map intuitively contains the information that is "deleted" when applying the linear map. Further, the kernel can be used to characterize the injectivity of linear maps. It also plays a central role in solving systems of linear equations.

Introduction Bearbeiten

We have learned about special mappings between vector spaces, called linear maps. Those are structure-preserving; that is, they are compatible with addition and scalar multiplication of a vector space. We can therefore think of a linear map from   to   as something that transports the vector space structure from   to  .

Introductory examples Bearbeiten

We consider two accounts, each with the account balance   and   respectively. We can describe this information with a vector  . The total account balance is the sum of the two account balances. We can calculate it by using the map

 

This map is linear and therefore transports the vector space structure from   to  . In the process, information is lost: one no longer knows how the money is distributed among the accounts. For example, one can no longer distinguish the individual account balances   and   because they both map to the same total account balance  . In particular, the mapping is not injective. However, we get the information about how much money is in the accounts in total.

 
Rotation of the real plane by 90° against the clockwise direction

Next, we consider the map

 

Visually, this corresponds to a counterclockwise rotation of   by   degrees. By undoing this rotation, one can recover the original vector from any rotated vector in  . Formally speaking, this mapping is an isomorphism and no information is lost. In particular, the image of linearly independent vectors is linearly independent again (because an isomorphism is injective, see the article monomorphism) and the image of a generator of   is again a generator of   (because an isomorphism is surjective, see the article epimorphism).

Finally, we consider a rotation again, but then embed the rotated plane into the  :

 

Although this mapping is no longer bijective, no information is lost here when transporting the vector space structure of the   into the  : As in the previous example, different vectors in the   are mapped to different vectors in the   because of injectivity. Linear independence of vectors is also preserved. However, a generating system of   is not mapped to a generator of  . For example, the linear map sends the standard basis   to  , which is not a generator of  . The property of a set of vectors to be a generator depends on the ambient space. This is not the case with linear independence; it is an "intrinsic" property of sets of vectors.

Derivation Bearbeiten

We have seen various examples of linear maps that transport a  -vector space into another  -vector space, while preserving the structure. In the process, varying amounts of "intrinsic" information from the original vector space (such as differences of vectors or linear independence) were lost. The last example suggests that injective mappings preserve such intrinsic properties. On the other hand, we see: If   is not injective, then there are vectors   with  . So in that case,   "eliminates" the difference   of   and  . The difference   is again an element in  . Since   is linear, we can reformulate:

 

Intuitively,   is injective if and only if differences   of vectors under   are not eliminated (i.e., mapped to zero). Because   is structure-preserving, we have that for all   and  , that   implies

 

If the difference of   and   is eliminated under  , so is that of   and  . In the same way, if  : if   and  , then also

 

So the difference of   and   is also eliminated. The differences eliminated by   are themselves vectors in  . These are send by   to the zero element   of   and thus, the eliminated vectors are in the preimage  . Conversely, any vector   can be written as a difference  ; that is, the difference   between   and the zero vector is eliminated by  . The preimage   measures exactly what differences of vectors (how much "information") is lost in the transport from   to  . Our considerations show that   is even a subspace of  . We give a name to this subspace: the kernel of  .

Definition Bearbeiten

The kernel of a linear map intuitively measures how much "intrinsic" information about vectors from   (differences of vectors or linear independence) is lost when applying the map. Mathematically, the kernel is the preimage of the zero vector.

Definition (Kernel of a linear map)

Let   and   be two  -vector spaces and   linear. Then we call   the kernel of  .

In the derivation we claimed that the kernel of a linear map from   to   is a subspace of  . We will now prove this in detail.

Theorem (The kernel is a vector space)

Let   be a linear map between the  -vector spaces   and  . Then   is a subspace of  .

Proof (The kernel is a vector space)

To verify the claim, we need to show four things:

  1.  
  2.  
  3. For all   we have that  .
  4. For all   and all   we have that  .

Proof step:  

The first assertion follows directly from the definition.

Proof step:  

Since   is linear, we know that   holds. So  .

Proof step: For all  , we have that  .

Now we show the third point: for all   it holds that

 

So also   is in the kernel of  .

Proof step: For all   and all   we have that  .

The fourth step works analogously to the third step: For all   and all   it is true that

 

Thus,  .

Examples Bearbeiten

We determine the kernel of the examples from the introduction.

Vector is mapped to the sum of entries Bearbeiten

We consider the mapping

 

The kernel of   is made up by the vectors   with  , so  . In other words

 

Thus the kernel of   is a one-dimensional subspace of  . More generally, for   we can consider the mapping

 

Again, by definition, a vector   lies in the kernel of   if and only if   holds. So we can freely choose   and then set  . Thus

 

Hence, the kernel of   is a  -dimensional subspace of  . It is also called a hyperplane in  .

Rotation in   Bearbeiten

We consider the rotation

 

Suppose   lies in the kernel of  , i.e. it holds that

 

From this we obtain  . So only the zero vector lies in the kernel of   and we have that  .

  is rotated and embedded into the   Bearbeiten

Next we consider

 

As in the previous example, we determine the kernel by choosing any vector  . Thus it holds that

 

Again it follows that  , so that also for this mapping   holds.

Derivatives of polynomials Bearbeiten

Finally, we consider a linear map that did not appear in the introduction:

 

which maps a real polynomial to its derivative. That is, a polynomial

 

with coefficients   is mapped to the polynomial

 

Graphically, we associate with   a polynomial   that indicates the gradient of   at each point. From this information, we still learn what the shape of the polynomial is (just as if we were given a stencil). However, we no longer know where it is positioned on the  -axis, because the information about the constant part of the polynomial is lost when taking the derivative. Polynomials that just differ by a displacement along the  -axis can no longer be distinguished after derivation. For example, both   and   have the derivative  . So the mapping   maps them to the same polynomial.

The kernel of   thus contains exactly the constant polynomials:

 

The inclusion " " is clear, because the derivative of a constant polynomial is always the zero polynomial. For the converse inclusion " ", we consider any polynomial   and show that it is constant. We can always write such a polynomial as   for some   and certain coefficients  . Because of   it holds that

 

and by comparison of the coefficients, we obtain  . So   is constant.

To-Do:

Once the polynomial ring article is written, link to the coefficient comparison in it

Kernel and injectivity Bearbeiten

In the derivation above, we saw that a linear map preserves all differences of vectors (i.e., no vector is eliminated) if and only if the kernel consists only of the zero vector. We also saw there that linearity implies: A linear map is injective if and only if no difference of vectors is eliminated. So we have the following theorem:

Theorem (Relationship between kernel and injectivity)

Let   and   be two  -vector spaces and let   be linear. Then   is injective if and only if  . In particular,   is injective if and only if  .

Summary of proof (Relationship between kernel and injectivity)

For establishing the theorem we have to show two directions:

  • If   is injective, then  .
  • From   it follows that   is injective.

The first direction we directly be shown. For the other direction, we assume   and show that for any   and   with   we must have  . Here, we can use that for two vectors   with  , we have  . Further,   is equivalent to  .

Proof (Relationship between kernel and injectivity)

Proof step: If   is injective, then  .

Let us first assume that   is injective. We already know that  . Since   is injective, it can map at most one argument to one function value. So only   is mapped to  . Thus  , because the kernel is defined as the set of all vectors that meet the zero vector.

Proof step: From   we get that   is injective.

Let  . In order to show that   is injective, we consider two vectors   and   from   with  . Then

 

So  . Since we have assumed  , it follows that   and thus  . Hence, we have the implication   for all  . But this is exactly the definition for   being injective.

Proof step:   is injective if and only if  .

We have already shown that   is injective if and only if  . It remains to show that this is equivalent to  . The kernel of   is a subspace of  . A subspace of   is exactly equal to   if its dimension is zero. So   is indeed injective if and only if  .

Alternative proof (Relationship between kernel and injectivity)

One can also show this theorem with only one chain of equivalent statements:

 


The larger the kernel is, the more differences between vectors are "eliminated" and the more the mapping "fails to be injective". The kernel is thus a measure of the "non-injectivity" of a linear map.

Injective maps and subspaces Bearbeiten

In the introductory examples we conjectured that injective linear maps preserve "intrinsic" properties of vector spaces. By this, we mean properties that do not depend on the ambient vector space, such as the linear independence of vectors or vectors being distinct. The property of being a generator can be lost in injective linear maps, as we have seen in the example of the twisted embedding of   into  : The mapping is injective, but the standard basis of   is not mapped to a generator of  .

What exactly does it mean that a property of a family   of vectors does not depend on the ambient space  ? Often, properties of vectors from   (for example, linear independence) depend on the vector space structure of  , that is, addition and scalar multiplication. To make dependences as small as possible, we restrict our attention to the smallest subspace of   containing  , that is, we restrict to  . Now, we call a property of   intrinsic if it depends only on   but not on  .

Example (Intrinsic and non-intrinsic properties)

Let   be a vector space and   a subset of vectors.

  • Linear independence of vectors in   is an intrinsic property, because the definition of linear independence can also be checked in   and does not refer to the ambient vector space  .
  • Differences of vectors in   are also intrinsic properties: all that is needed to examine it are vectors   and their difference  .
  • Not intrinsic, on the other hand, is the property of   of being a generator of  : The set   is always a generator of  . But if the ambient space   is larger than  , then   is not a generator of  .

What do intrinsic properties of a family of vectors have to do with injectivity? Let   be a linear map. Suppose   preserves intrinsic properties of vectors, that is, if a family   has some intrinsic property, then its image   under   also has this property. Then   also preserves the property of vectors being different, since this is an intrinsic property. That means, if   are different, i.e.,  , then their image under   is also different, i.e.,  . So   is injective.

Conversely, if   is injective, then   is isomorphic to the subspace   of  : If we restrict the target space of   to its image, we obtain an injective and surjective linear map  , that is, an isomorphism. In particular, for any family   in  , it holds that the subspace   of   is isomorphic to  . Thus, the latter has the same properties as   and hence,   preserves intrinsic properties of subsets of  .

So we have seen that   is injective if and only if   preserves intrinsic properties of subsets of  .

Kernel and linear independence Bearbeiten

In the previous section we have seen that injective linear maps   are exactly those linear maps which preserve intrinsic properties of  . The linear independence of a family of vectors is such an intrinsic property, as they either hold for any choice of an ambient space or do not hold for any choice of an ambient space.

So, injective linear maps should preserve linear independence of vectors, i.e., the image of linearly independent vectors is again linearly independent. Conversely, a linear map cannot be injective if it does not preserve the linear independence of vectors, since the intrinsic information of "being linearly independent" is lost.

Overall, we get the following theorem, which has already been proved in the article on monomorphisms:

Theorem (Injective linear maps preserve linear independence)

Let   and   be two  -vector spaces and   a linear map. Then   holds if and only if the image of every linearly independent subset of   is again linearly independent.

In particular, for any linear map  , the vector space   is a  -dimensional subspace of  . In the finite-dimensional case, there cannot exist an injective linear map from   to   if  . This has also already been shown in the article on monomorphisms.

Kernel and linear systems Bearbeiten

The kernel of a linear map is an important concept in the study of systems of linear equations.

Let   be a field and let  . We consider a linear system of equations

 

with   variables   and   rows. We have  , where   and  . We can also write this system of equations using matrix multiplication:

 

where  ,   and  . We denote the set of solutions by

 

Determining a solution to the linear system of equations   for a given right-hand side   is the same as finding a preimage of   under the linear map

 
To-Do:

Link where the map "multiply matrices by a given fixed matrix" is studied? Especially where it is explained that it is linear. Possibly also to the article where it is explained how to determine the kernel of a matrix (Gauss), if this is written.

The system of equations   has solutions if the preimage   is not empty. In this case, we may ask whether there are multiple solutions, that is, whether the solution is not unique. In other words, we are interested in how many preimages a   has under  .

By definition of injectivity, every point   has at most one element in its preimage if and only if   is injective. This means that the linear system of equations   has at most one solution for each  , that is,  . Because   is linear, injectivity is equivalent to  . So we can already state:

Theorem (Uniqueness of solutions)

Let   be a field and let  ,   and  . Then

 

Hint

The set of solutions of   can be empty. This occurs, for example, when   is the zero matrix and  . Consequently, the kernel makes no statement about the existence of solutions, only about their uniqueness. To say something about the existence of solutions, we need to consider the image of  .

Even if   is not injective, i.e.,   holds, we can still say more about the set of solutions by exploiting the kernel: The difference of two vectors   and  , which   maps to the same vector, lies in the kernel of  . Therefore, the preimage of some   under   can be written as

 

where   is any element of  . This is shown by the following theorem:

Theorem (Solution set of linear system and kernel)

Let   be a field and let  ,   and  . further, let   be a solution of the linear system of equations  . Then

 

In particular, a solution   of the system of equations is unique if and only if the linear map   induced by   has a kernel that only consists of the zero vector.

Proof (Solution set of linear system and kernel)

We have to prove the equality  . For this we need to establish two subset relations.

Proof step:  

Let  . Then  . The only possible candidate for   to satisfy the equation   is  . Since

 

we have  .

Proof step:  

We show that   holds for any  . Let   be arbitrary. Then   holds. Since by assumption,   is a solution of  , we have that

 

So   is also a solution of   and thus lies in the set  .

We have thus even extended the statement of the theorem above. The larger the kernel of   is, that is, the "less injective" the mapping   is, the "less unique" are solutions of  , if any exist. The set of solutions of a linear system of equations   is the kernel of the induced linear map   shifted by a particular solution  . Furthermore,

 

The set of solutions of the homogeneous system of equations   (that is, with right-hand side zero) is exactly the kernel of  .

Hint

As with the previous theorem, no statement is made about whether solutions of   exist at all for a given  . The kernel only characterizes uniqueness.

Exercises Bearbeiten

Exercise (Injectivity and dimension of   and  )

Let   and   be two finite-dimensional vector spaces. Show that there exists an injective linear map   if and only if  .

How to get to the proof? (Injectivity and dimension of   and  )

To prove equivalence, we need to show two implications. For the execution, we use that every monomorphism   preserves linear independence: If   is a basis of  , then the   vectors   are linearly independent. For the converse direction, we need to construct a monomorphism from   to   using the assumption  . To do this, we choose bases in   and   and then use the principle of linear continuation to define a monomorphism by the images of the basis vectors.

Solution (Injectivity and dimension of   and  )

Proof step: There is a monomorphism  

Let   be a monomorphism and   a basis of  . Then   is in particular linearly independent and therefore   is linearly independent. Thus, it follows that  . So   is a necessary criterion for the existence of a monomorphism from   to  .

Proof step:   there is a monomorphism

Conversely, in the case   we can construct a monomorphism: Let   be a basis of   and   be a basis of  . Then  . We define a linear map   by setting

 

for all  . According to the principle of linear continuation, such a linear map exists and is uniquely determined. We now show that   is injective by proving that   holds. Let  . Because   is a basis of  , there exist some   with

 

Thus, we get

 

Since   are linearly independent,   must hold for all  . So it follows for   that

 

We have shown that   holds and thus   is a monomorphism.

Exercise

We consider the linear map  . Determine the kernel of  .

Solution

We are looking for vectors   such that  . Let   be any vector in   for which   is true. We now examine what properties this vector must have. It holds that

 

So   and  . From this we conclude  . So any vector   in the kernel of   satisfies the condition  . Now take a vector   with  . Then

 

We see that  . In total

 

Check your understanding: Can you visualize   in the plane? What does the image of   look like? How do the kernel and the image relate to each other?

 
The kernel of f

We have already seen that

 

Now we determine the image of   by applying   to the canonical basis.

 

So   holds. We see that the two vectors are linearly dependent. That is, we can generate the image with only one vector:  .

In our example, the image and the kernel of the linear map   are straight lines through the origin. The two straight lines intersect only at the zero and together span the whole  .

Exercise

Let   be a vector space,  , and   be a nilpotent linear map, i.e., there is some   such that

 

is the zero mapping. Show that   holds.

Does the converse also hold, that is, is any linear map   with   nilpotent?

Solution

Proof step:   nilpotent  

We prove the statement by contraposition. That is we show: If  , then   is not nilpotent.

Let  . Then   is injective, and as a concatenation of injective functions,   is also injective. By induction it follows that for all   the function   is injective. But then also   for all  . Since the kernel of the zero mapping would be all of  , the map   could not be the zero mapping for any  . Consequently,   is not nilpotent.

Proof step: The converse implication

The converse implication does not hold. There are mappings that are neither injective nor nilpotent. For example we can define

 

This mapping is not injective, because  . But it is also not nilpotent, because we have   for all  .