arXiv:2010.12551v3 [math.RA] 16 Aug 2021 h otitrsigaetoecrepnigt h clrfunct scalar the to theo corresponding mathematicians This those many are by mathematics. interesting studied most applied extensively the and been pure has and in developed applications various in used nes so nrni neet thsipratpoete htmak that properties important has it applic interest, (se of inverse intrinsic range Drazin of the wide and is a matrices inverse have of powers and the importance between relation considerable of are trices bene be should exponential. formulas these the addition, explicit involving In present to practice. is and purpose sentation our applications, equations. important other these and of space r which solution method each a for provided [5] a Liz solvin of E. avoid equations, To differential coefficients. these constant for me th with alternative exactly equations an requires differential presented approach ear Leonard His E. I. prerequisites. [4], mathematical In the [8]. Putzer of rithm whe roots. case multiple general has the method produce matrix not a the g does of generalizes that method approach His that prese manageable cases. and procedure authors method special elegant [6], a the an In presented all presented [1] describe [7] 7]). Apostol to 6, Howland attempted 5, L. have they 4, J. algorithms efforts 3, such 2, and various [1, methods on example several for attention calculation, (see great decades such attracted For has and ature. researchers many by studied matrices. of k od arcs rjcin fmatrices. of projections matrices; monde OAIH,THE LOGARITHM, OEEPII OMLSFRMTI XOETA,MATRIX EXPONENTIAL, MATRIX FOR FORMULAS EXPLICIT SOME sa nee.I hsatce eaeitrse ntecmuainof computation the in interested are we article, this In integer. an is h hoyo ucin fmtie ly eta oei ieralge linear in role central a plays matrices of functions of theory The 2020 eto sdvtdt h oesadteDai nes fmatrice of inverse Drazin the and powers the to devoted is 3 Section in exponential matrix the of appearance ubiquitous the of Because is exponential matrix the of computation the of difficulty the of Much t of computation The exponential. matrix the to devoted is 2 Section e od n phrases. and words Key ahmtc ujc Classification. Subject Mathematics xmlsaefruae oilsrt h ehd presente methods obtai the is illustrate matrices to pr formulated of are are logarithm inverses examples for Drazin formula their explicit closed and Chevallelegant addition, matrices In the of matrix. obtain powers any positive easily More of we projections subject. consequence, spectral the essential a this As on m and generalized. literature tractable extensive gives it the elementary, in and direct is method Our Abstract. nti ok e lsdfr omlsfrtemti expone matrix the for formulas closed-form new work, this In arxepnnil arxlgrtm rznivre;po inverses; Drazin logarithm; matrix exponential; Matrix OAMDMU¸OFADSI ZRIAA SAID MOUC¸ AND MOHAMMED OUF n HPWRO ARCSADTERDRAZIN THEIR AND MATRICES OF POWER TH 1. 51,1A9 15B99. 15A09, 15A16, INVERSES Introduction 1 fr xrsin o h arbitrary the for expressions -form e.Svrlpriua ae and cases particular Several ned. vr tesaercprtdand recuperated are others over, nti paper. this in d sne.Uigteersls an results, these Using esented. ngal omlsntcurrent not formulas anageable yJra eopsto and decomposition ey–Jordan ose ions h hrceitcpolynomial characteristic the n ouino ooeeu lin- homogeneous of solution e ndvreppr nteliter- the in papers diverse in v xlctfrua o some for formulas explicit ave ,Mucu 1].TeDrazin The Mou¸couf [19]). e, ca nftr investigations future in ficial htse ob practical. be to seem that s tdacrflinvestigation careful a nted omlst ipiyispre- its simplify to formulas h nta au problems value initial the g hditnigt minimize to intending thod qie nyteknowledge the only equires x teteeyueu and useful extremely it e perdi h atfew last the in appeared mn hs functions, these Among . ierdnmclsystems dynamical linear log( , ecie n[] .M. T. [6]. in described ta r provided. are ntial tos hr sadeep a is There ations. bv itdfunctions listed above .Tepwr fma- of powers The s. eso arcs Vander- matrices; of wers yasdb h algo- the by bypassed r n ti widely is it and bra ywssubsequently was ry i arxhsbeen has matrix his x ), x 1 and x k where 2 MOHAMMED MOUC¸OUF AND SAID ZRIAA interesting in various applications including cryptography, Markov chains, functional analysis, differential-algebraic equations, stochastic process, multibody system, Control theory (see for example [9, 10, 11] and references therein). Many authors have investigated the Drazin inverse from different viewpoints. Here, we propose an interesting explicit expression for the Drazin inverse for matrices and their powers in terms of the eigenvalues. The advantage of our method is that it gives explicit tractable representations for the Drazin inverse of a matrix and needs only calculating certain polynomials of this matrix. In section 4 we consider the logarithm of matrices. This function of matrices appears in many parts of mathematics, applied natural sciences, and engineering. It has many applications in the study of Systems Theory and has received much interest in Control Theory (see [12, 13] and references therein). If the matrix A has no eigenvalues on the closed negative real axis R−, then A has a unique logarithm with eigenvalues in the set z C/ π < Im(z) < π . This unique logarithm is called the principal logarithm of A. This{ logarithm∈ − is of great interest} and is needed in many applications. While there are various methods for the computation of the matrix exponential, relatively few ones exist for the matrix logarithm. The exact computation of the matrix logarithm re- veals considerable difficulties. Most of the methods presented for computing such matrix are approximation methods [12, 13]. Our motivation comes from the fact that few works have been considered for exhibiting compact formulas for the matrix logarithm. It is important to em- phasize that other methods for computing the matrix logarithm require advanced theory such as matrix square roots, Schur decompositions, and Pad´eapproximants (see [14, 15, 12, 13] and references therein). In a previous work, the first-named author of this paper has presented a method for obtain- ing easily from any given -canonical form of an arbitrary nonsingular k k matrix A with P × eigenvalues α1,...,αk, both a logarithm B of A, the eigenvalues of B may be chosen in advance arbitrarily as log(α1),..., log(αk), and one of its -canonical forms (see, Mou¸couf [16]). This result is the key for obtaining our logarithm matricesP formulas. We propose simple, direct, and compact formulas to compute the logarithms and the principal logarithm of matrices without any restrictions on the norm. We note that the Putzer matrix representation of the logarithm of complex matrices [13], or real matrices [15], requires the exact computations of some rational integrals. We also note that in a recent work [14], the authors are interested in developing the exact computation for the principal logarithm of ma- trices. More precisely, under some conditions on the norm, they compute exactly the principal logarithm matrix, but in this method it is necessary to solve a system of linear recursive equa- tions, determine the Fibonacci-H¨orner decomposition of the matrix, investigate the properties of generalized Fibonacci sequences and their Binet formula. In addition, their approach requires laborious calculations of some functions. Contrary to these works and other many works on this subject, our method does not require any cumbersome and laborious calculations. Additionally, our approach is general. The attractive feature of our proposed method lies in the possibility of choosing in advance the eigenvalues of logarithms of a matrix, and hence easily finding the principal logarithm of matrices. Among the essential tools used in this paper are the Vandermonde matrices and their in- verses and certain polynomials. So it is convenient to fix some notations. Let α1, α2,...,αs be distinct elements of C and m1,m2,...,ms be nonnegative integers. For m1 m2 m a non-constant polynomial P (x) = (x α ) (x α ) (x αs) s of degree n, we denote − 1 − 2 ··· − by Ljkj (x)[P ] the following polynomial

mj 1 kj − − kj 1 (i) i Ljk (x)[P ]= Pj (x)(x αj ) g (αj )(x αj ) (1.1) j − i! j − Xi=0 where 1 j s, 0 kj mj 1 and ≤ ≤ ≤ ≤ − s mi P (x) Pj (x)= (x αi) = m , 1 j s − (x αj ) j ≤ ≤ i=1Y,i=j 6 − THE EXPLICIT MATRIX EXPONENTIAL 3 and 1 gj (x) = (Pj (x))−

Here and further L(l) (x)[P ] means the lth derivative of L (x)[P ]. jkj jkj According to [17], we can write every polynomial Q of degree less than or equal to n 1 as − s mj 1 − 1 (kj ) Q = ( Q (αj )Ljkj (x)[P ]) (1.2) kj ! Xj=1 kXj =0

This formula is of great importance, it is used in [18] to invert the confluent . In [18], the confluent Vandermonde matrix associated with the polynomial P of degree n = m + m + + ms 1 2 ··· m1 m2 ms P (x) = (x α ) (x α ) (x αs) , − 1 − 2 ··· − is defined to be the following block matrix

VG(P ) = (V1 V2 ... Vs) (1.3) where α1, α2,...,αs are distinct elements of C. The block matrix Vk is of order n mk, k =1,...,s, and defined to be the matrix × mk Vk = VG((x αk) ) − with entries i 1 i j j−1 αk− if i j (Vk)ij = ≥  0,−  otherwise , q where p denotes the binomial coefficient. For completeness, we recall the following Theorem and corollary (needed in the sequel) which provide an explicit closed-form for the inverse of the confluent Vandermonde matrices (for more details, see [18]).

m1 m2 m Theorem 1.1. Let P (x) = (x α ) (x α ) (x αs) s be a polynomial of degree n, − 1 − 2 ··· − where α1, α2,...,αs are distinct elements of C. The explicit inverse of the confluent Vander- monde matrix VG(P ) has the form

m1 L1  2m2  1 L VG− (P )= . (1.4)  .     sm  L s  for r =1, 2,...,s the block matrix rm is of order mr n and given by L r × 1 (j 1) = L − (0)[P ] rmr (j 1)! r(i 1) L  − 1 i mr,1 j n − ≤ ≤ ≤ ≤ More precisely,

(1) 1 (n 1) Lr (0)[P ] L (0)[P ] L − (0)[P ] 0 r0 ··· (n 1)! r0  (1) −1 (n 1)  − Lr1(0)[P ] Lr1 (0)[P ] (n 1)! Lr1 (0)[P ] ··· − rmr =  . . .  L  . . .   . . .  (1) ··· 1 (n 1)  −  Lrmr 1(0)[P ] Lrmr 1(0)[P ] (n 1)! Lrmr 1(0)[P ]  − − ··· − −  The following corollary treats the interesting case of the generalized Pascal matrix 4 MOHAMMED MOUC¸OUF AND SAID ZRIAA

Corollary 1.2. Let α be a complex number and n be a positive integer. Then the inverse of the Vandermonde matrix 1 0 0 ···  α 1 0 α2 2α ··· 0  ···  n  α3 3α2 0 VG((x α) )=  4 4 ···  (1.5) −  α4 α3 0  0 1 ···   . . .  . . .  n 1 n 1 n 1 n 2 ···   − α − − α − 1  0 1 ···  is   1 0 0 ···  α 1 0 −α2 2α ··· 0  − ···  1 n α3 3α2 0 −   VG ((x α) )=  4 − 4 ···  −  ( α)4 ( α)3 0  0 − 1 − ···    .  . .  . . .  n 1 n 1 n 1 n 2 ···   − ( α) − − ( α) − 1  0 − 1 − ···    2. Explicit formulas for the matrix exponential Let us consider the following system of differential equations X˙ = AX where A = (aij )1 i,j k is a constant matrix with entries in C and X(t) the vector column ≤ ≤ T defined by X(t) = (x1(t), x2(t),...,xk(t)) . Given any square matrix A, the exponential matrix function is + ∞ tn etA = An n! nX=0 It is well known that the function tA X(t)=e X0 is the theoretical solution of the equation

X˙ = AX, X(0) = X0 The last differential equation has gained much importance in linear and dynamical systems. Using the results of the previous section, we develop a purely algebraic approach, that requires only the knowledge of eigenvalues of the matrix, to derive more explicit expressions for the exponential of an arbitrary complex matrix. Let χ be a unital polynomial of degree k k k 1 k 2 χ(x)= x a x − a x − ak − 1 − 2 −···− and consider the following set of differentiable functions mapping C to the complex matrices of order k (k) (k 1) (k 2) F (χ)= f/f (t)= a f − (t)+ a f − (t)+ + akf(t) { 1 2 ··· } where f (l)(t) denote the lth derivative of f(t). It is well known that F (χ) is a C-vector space of dimension k. Let D(χ) be the vector space of all complex functions satisfying the following linear differential equation (k) (k 1) (k 2) y (t)= a y − (t)+ a y − (t)+ + aky(t) 1 2 ··· Let yi, 0 i k 1, be the elements of D(χ) with the initial conditions ≤ ≤ − (j) y (0) = δij , for 0 j k 1. i ≤ ≤ − THE EXPLICIT MATRIX EXPONENTIAL 5

It is well known that y0(t),y1(t),...,yk 1(t) is a basis of D(χ); it is called the canonical basis of D(χ). The most important{ property− of this} basis is that all the elements of F (χ) can be expressed as the linear combinations of the yi’s only in terms of their derivatives at t = 0. More precisely, each f F (χ) can be written ∈ k 1 − (i) f(t)= yi(t)f (0) (2.1) Xi=0 As a particular case of Formula (2.1), we find the following well known result.

Proposition 2.1. For any k k matrix A, the exponential matrix function is given by × k 1 tA − i e = yi(t)A Xi=0

tA Proof. Follows immediately from the fact that e F (χA), where χA is the characteristic ∈ polynomial of A and y0(t),y1(t),...,yk 1(t) is the canonical basis of D(χA).  { − } In [1] the author intents to find, in the simplest way, a method for computing the exponential of a matrix but he does not treat all cases. In what follows, we present the general results without any disadvantages. We begin with a proposition that will be useful in the sequel.

Proposition 2.2. If e (t),e (t),...,ek(t) is a basis of the vector space D(χA), then there { 1 2 } exist unique constant matrices B1,B2,...,Bk such that

tA e = e (t)B + e (t)B + + ek(t)Bk 1 1 2 2 ··· and consequently, we have

n (n) (n) (n) A = e (0)B + e (0)B + + e (0)Bk 1 1 2 2 ··· k for every n N. ∈

Proof. This result follows directly from Proposition 2.1 and the fact that e (t),e (t),...,ek(t) { 1 2 } is a basis of the vector space D(χA). 

In the following, we provide the explicit expressions of the elements y0(t),y1(t),...,yk 1(t) of D(χ) in terms of the roots of χ. −

m1 m2 ms Theorem 2.3. Let χ(x) = (x α1) (x α2) (x αs) be a unital polynomial of degree k. Then the (i+1)th element of− the canonical− basis··· of D−(χ) is

s mp 1 αpt − rp e t (i) yi(t)= ( Lprp (0)[χ]), 0 i k 1 i! rp! ≤ ≤ − Xp=1 rXp=0 Proof. It is well known that

m1 1 ms 1 t − t − B = eα1t,teα1t,..., eα1t,..., eαst,teαst,..., eαst { (m 1)! (ms 1)! } 1 − − is a basis of D(χ). By expressing each member of this basis in terms of the canonical basis, we get k 1 ti − ti eαpt = ( eαpt)(j)(0)y (t) i! i! j Xj=0 6 MOHAMMED MOUC¸OUF AND SAID ZRIAA for 0 i mp 1 and p = 1, 2,...,s. The resulting matrix from B to the canonical≤ basis≤ is− the confluent Vandermonde matrix (1.3) 1 0 0 1 0 0 ··· ··· ···  α1 1 0 αs 1 0 2 ··· ··· 2 ··· α 2α1 0 α 2αs 0  1 ··· ··· s ···  VG(χ)=  α3 3α2 0 α3 3α2 0  1 1 ··· ··· s s ···   ......   ......   k 1 k 2 ··· ··· k 1 k 2 ···  α − (k 1)α − 1 α − (k 1)α − 1  1 − 1 ··· ··· s − s ···  Using the inverse given by (1.4), we obtain

s mp 1 αpt − rp e t (i) yi(t)= ( Lprp (0)[χ]), 0 i k 1 i! rp! ≤ ≤ − Xp=1 rXp=0 as desired.  As a particular case we have the following corollary.

Corollary 2.4. If χ(x) = (x α )(x α ) (x αk) has distinct roots α , α ,...,αk, then − 1 − 2 ··· − 1 2 k αj t e (i) yi(t)= L (0)[χ], 0 i k 1 i! j0 ≤ ≤ − Xj=1 In the following we state one of our main results.

m1 m2 ms Theorem 2.5. Let A be a k k matrix, and let χA(x) = (x α1) (x α2) (x αs) be its characteristic polynomial.× Then − − ··· −

k 1 s mp 1 − αpt − rp tA e t (i) i e = [ ( Lprp (0)[χA])]A i! rp! Xi=0 Xp=1 rXp=0 Proof. The result is an immediate consequence of Theorem 2.3 and Proposition 2.1.  As a particular case we have the following corollary.

Corollary 2.6. If χA(x) = (x α )(x α ) (x αk) has distinct roots α , α ,...,αk, then − 1 − 2 ··· − 1 2 k 1 k − eαj t etA = ( L(i)(0)[χ ])Ai i! j0 A Xi=0 Xj=1 Proof. Follows immediately from Theorem 2.3 and Theorem 2.5.  In the case where the matrix A has a single eigenvalue, we have the following result. k Corollary 2.7. If χA(x) = (x α) has a single root α, then − αt k 1 l+i e − ( 1) l i l yi(t)= ( − α − t ), 0 i k 1 i! (l i)! ≤ ≤ − Xl=i − and k 1 k 1 l+i tA − 1 − ( 1) l i l i αt e = ( − α − t )A e i! (l i)! Xi=0 Xl=i − Proof. The result is an immediate consequence of Theorem 2.3 and Theorem 2.5.  For illustration purposes, we consider the case of a square matrix of order 3. Example 1. Let A be a square matrix of order 3 with characteristic polynomial 3 2 χA(x)= x a x a x a − 1 − 2 − 3 THE EXPLICIT MATRIX EXPONENTIAL 7

1. If α is a root of multiplicity 3 of χA(x) then, using Theorem 2.7, we have

2 l 2 l+1 αt 2 l tA αt ( 1) l l αt ( 1) l 1 l e ( 1) l 2 l 2 e = (e − α t )I + (e − α − t )A + ( − α − t )A l! (l 1)! 2! (l 2)! Xl=0 Xl=1 − Xl=2 − Consequently, we obtain

1 1 etA = (1 αt + α2t2)eαtI + (t αt2)eαtA + t2eαtA2 − 2 − 2

2. If α1 is a root of multiplicity 2, and α2 is a simple root, then Theorem 2.5 gives

etA =

α1t α2t (e (L10(0)[χA]) + tL11(0)[χA]) + e L20(0)[χA])I+

α1t (1) (1) α2t (1) (e (L10 (0)[χA]) + tL11 (0)[χA]) + e L20 (0)[χA])A+ 1 (eα1t(L(2)(0)[χ ]) + tL(2)(0)[χ ]) + eα2tL(2)(0)[χ ])A2 2! 10 A 11 A 20 A

Formula (1.1) yields

x2 +2α x 2α α + α2 L (x)[χ ] = − 1 − 1 2 2  10 A (α α )2 1 − 2  2  x (α1 + α2)x + α1α2  L (x)[χ ] = −  11 A (α α )  1 − 2 2 2  x 2α1x + α1  L20(x)[χA] = − 2  (α2 α1)  −  A simple calculation gives

1 etA = ((α2 2α α )eα1t + (α2α α α2)teα1t + α2eα2t)I+ (α α )2 2 − 1 2 1 2 − 1 2 1 1 − 2  ((α2 α2)teα1t +2α (eα1t eα2t))A + (eα2t eα1t + (α α )teα1t)A2 1 − 2 1 − − 1 − 2 

3. If α1, α2, α3 are simple roots of χA(x) then, using Corollary 2.6, we have

3 3 3 eαj t etA = ( eαj tL (0)[χ ]))I + ( eαj tL(1)(0)[χ ]))A + ( L(2)(0)[χ ]))A2 j0 A j0 A 2! j0 A Xj=1 Xj=1 Xj=1

Formula (1.1) gives

(x α )(x α ) x2 (α + α )x + α α L (x)[χ ]) = − 2 − 3 = − 2 3 2 3  10 A (α α )(α α ) (α α )(α α ) 1 − 2 1 − 3 1 − 2 1 − 3  2  (x α1)(x α3) x (α1 + α3)x + α1α3  L (x)[χ ]) = − − = −  20 A (α α )(α α ) (α α )(α α )  2 − 1 2 − 3 2 − 1 2 − 3 2  (x α1)(x α2) x (α1 + α2)x + α1α2  L30(x)[χA]) = − − = −  (α3 α1)(α3 α2) (α3 α1)(α3 α2)  − − − −  8 MOHAMMED MOUC¸OUF AND SAID ZRIAA

Consequently, the matrix exponential in this case is given by

etA = α α α α α α 2 3 eα1t + 1 3 eα2t + 1 2 eα3t I (α α )(α α ) (α α )(α α ) (α α )(α α )  1 − 2 1 − 3 2 − 1 2 − 3 3 − 1 3 − 2 α + α α + α α + α + 2 3 eα1t + 1 3 eα2t + 1 2 eα3t A (α α )(α α ) (α α )(α α ) (α α )(α α )  3 − 1 1 − 2 3 − 2 2 − 1 1 − 3 3 − 2 eα1t eα2t eα3t + + + A2 (α α )(α α ) (α α )(α α ) (α α )(α α )  1 − 2 1 − 3 2 − 1 2 − 3 3 − 1 3 − 2

In the present approach, one gains a basic construction of the so-called spectral decomposition of A. The approach makes the determination of the spectral decomposition of any square matrix more practical than the usual method of partial fraction decomposition. Using Lagrange polynomials W. A. Harris et al. [3] have derived the spectral decomposition of a matrix with simple eigenvalues. Here, we generalize this result to any matrix using a generalization of Hermite’s formula given by A. Spitzbart [17].

m1 m2 ms Theorem 2.8. Let A be a k k matrix, and let χA(x) = (x α1) (x α2) (x αs) be its characteristic polynomial.× Then − − ··· −

s mj 1 − kj tA t αj t e = e Bjkj (2.2) kj ! Xj=1 kXj =0

where Bjk = Ljk (A)[χA]. Moreover, for i = j we have j j 6

Biki Bjkj = Bjkj Biki = 0 (2.3)

Proof. Using Proposition 2.2 with the fact that

m1 1 ms 1 t − t − eα1t,teα1t,..., eα1t,..., eαst,teαst,..., eαst { (m 1)! (ms 1)! } 1 − −

is a basis of D(χA), we can find unique constant matrices Bjk , j = 1, 2,...,s and 0 kj j ≤ ≤ mj 1, such that −

m1 1 m2 1 ms 1 − j − j − j tA t α1t t α2t t αst e = e B j + e B j + + e Bsj j! 1 j! 2 ··· j! Xj=0 Xj=0 Xj=0 THE EXPLICIT MATRIX EXPONENTIAL 9

Applying Proposition 2.2 to this equality yields

1 0 0 1 0 0 ··· ··· ···  α1 1 0 αs 1 0 2 ··· ··· 2 ··· α1 2α1 0 αs 2αs 0  3 2 ··· ··· 3 2 ···   α1 3α1 0 αs 3αs 0  ··· ··· ···   ......   ......   ··· ··· ···   ......   ......   ··· ··· ···   ......   ......   k 1 k 2 ··· ··· k 1 k 2 ···  α − (k 1)α − 1 α − (k 1)α − 1  1 − 1 ··· ··· s − s ···  I B10 A  B11    . .  .   .   m1 1   A −  B1m1 1  −   .   .  =  .  (2.4)  .   .     .   B   .   s0   .   .     .   .     .     k 1  Bsms 1  A   −   − 

On the other hand, utilizing Formula (1.2) for the canonical basis of Ck 1[x], we obtain the following system −

1 0 0 1 0 0 ··· ··· ···  α1 1 0 αs 1 0 2 ··· ··· 2 ··· α1 2α1 0 αs 2αs 0  3 2 ··· ··· 3 2 ···   α1 3α1 0 αs 3αs 0  ··· ··· ···   ......   ......   ··· ··· ···   ......   ......   ··· ··· ···   ......   ......   k 1 k 2 ··· ··· k 1 k 2 ···  α − (k 1)α − 1 α − (k 1)α − 1  1 − 1 ··· ··· s − s ···  1 L10(x)[χA] x  L11(x)[χA]    . .  .   .   m1 1   x −  L1m1 1(x)[χA]  −   .   .  =  .   .   .     .   L (x)[χ ]   .   s0 A   .   .     .   .     .     k 1  Lsms 1(x)[χA]  x   −   −  10 MOHAMMED MOUC¸OUF AND SAID ZRIAA

Replacing in this matrix equation x with A yields 1 0 0 1 0 0 ··· ··· ···  α1 1 0 αs 1 0 2 ··· ··· 2 ··· α1 2α1 0 αs 2αs 0  3 2 ··· ··· 3 2 ···   α1 3α1 0 αs 3αs 0  ··· ··· ···   ......   ......   ··· ··· ···   ......   ......   ··· ··· ···   ......   ......   k 1 k 2 ··· ··· k 1 k 2 ···  α − (k 1)α − 1 α − (k 1)α − 1  1 − 1 ··· ··· s − s ···  I L10(A)[χA] A  L11(A)[χA]    . . .  .   .   m1 1   A − L1m1 1(A)[χA]    −   .   .  =  .   .   .     .   L (A)[χ ]   .   s0 A   .   .     .   .     .   k 1  Lsms 1(A)[χA]  A   −   −  Since the confluent Vandermonde matrix is invertible, we have

Bjk = Ljk (A)[χA] for all 1 j s, 0 kj mj 1 j j ≤ ≤ ≤ ≤ − On the other hand, by using (1.1), we can show that χA(x) divides

Ljkj (x)[χA]Liki (x)[χA] for i = j. As a consequence, we get 6

Bjkj Biki = Biki Bjkj =0 

The following example is used to illustrate Theorem 2.5 and Theorem 2.8. Example 2. Let us consider Example 1. of [4] 2 0 1 A = 0 2 0 0 0 3   2 The characteristic polynomial of A is χA(x) = (x 2) (x 3). In this case, by using For- mula (1.1), we have − − 2 L (x)[χA] = (x 3)(1 x)= x +4x 3 10 − − − − 2 L (x)[χA] = (x 3)(2 x)= x +5x 6 11 − − − − 2 2 L (x)[χA] = (x 2) = x 4x +4 20 − −

We deduce that ′ ′′ L10(0)[χA] = 3,L10(0)[χA]= 4,L10(0)[χA] = 2 − ′ ′′ − L11(x)[χA] = 6,L11(0)[χA]= 5,L11(0)[χA] = 2 − ′ ′′ − L (x)[χA]= 4,L (0)[χA] = 4,L (0)[χA]= 2 20 20 − 20 Using the formula of Theorem 2.5, we obtain 1 etA = (e2t( 3 6t)+4e3t)I + (e2t(4+5t) 4e3t)A + (e2t( 2 2t)+2e3t)A2 − − − 2 − − THE EXPLICIT MATRIX EXPONENTIAL 11

Therefore e2t 0 e3t e2t tA − e =  0 e2t 0  00 e3t   Now we find the matrix exponential of A, but using this time Theorem 2.8. From Formula (2.2), we get tA 2t 2t 3t e =e B10 + te B11 +e B20 where 2 B = L (A)[χA]= A +4A 3I 10 10 − − 2 B = L (A)[χA]= A +5A 6I 11 11 − − 2 B = L (A)[χA]= A 4A +4I 20 20 −

More explicitly, 1 0 1 0 0 0 0 0 1 − B10 = 01 0  ,B11 = 0 0 0 ,B20 = 0 0 0 00 0 0 0 0 0 0 1       Thus tA 2t 2t 3t e =e B10 + te B11 +e B20 That is e2t 0 e3t e2t tA − e =  0 e2t 0  00 e3t   Next, we derive some corollaries of Theorem 2.8.

m1 m2 Corollary 2.9. If χA(x) = (x α1) (x α2) , α1 and α2 are two distinct complex numbers, then − − m1 1 m2 1 − tj − tj etA = eα1tB + eα2tB j! 1j j! 2j Xj=0 Xj=0

m1 j 1 i m2+i 1 − − ( 1) − j m2 m2 1 i B1j = (A α1I) (A α2I) − − (A α1I)  − − (α α )m2+i −  Xi=0 1 − 2  m2 j 1 i m1+i 1  − − ( 1) −  m1 j m1 1 i B2j = (A α1I) (A α2I) − − (A α2I) − − (α α )m1+i −  Xi=0 2 1  −  Proof. This follows directly from (2.2) and the fact that

m1 j 1 i m2+i 1 − − ( 1) − m2 1 i+j m2 L1j(x)[χA] = − − (x α1) (x α2)  (α α )m2+i − −  Xi=0 1 − 2  m2 j 1 i m1+i 1  − − ( 1) −  m1 1 m1 i+j L2j(x)[χA] = − − (x α1) (x α2) (α α )m1+i − −  Xi=0 2 1  −  

m1 Corollary 2.10. If χA(x) = (x α1) (x α2), α1 and α2 are distinct two complex numbers, then − − m1 1 − tj etA = eα1tB +eα2tB j! 1j 20 Xj=0 12 MOHAMMED MOUC¸OUF AND SAID ZRIAA where m1 j 1 − − i j ( 1) i B j = (A α I) (A α I) − (A α I)  1 − 1 − 2 (α α )i+1 − 1  Xi=0 1 2  −  1 m1 B20 = (A α1I) (α α )m1 −  2 1  − More generally, we have the following result. s m1 Corollary 2.11. If χA(x) = (x α ) (x αj ), α ,...,αs are distinct complex numbers, − 1 − 1 jQ=2 then m1 1 s − j tA t α1t αj t 1 e = e B1j + e Pj (A) (2.5) j! Pj (αj ) Xj=0 Xj=2 where P = χA,

s m1 j 1 s − − i j ( 1) al i B1j = (A α1I) (A αlI) − i+1 (A α1I) − − (α1 αl) − Yl=2 Xi=0 Xl=2 − and 1 s if s 3  ≥ al = (αl αp)  p=2,p=l −  Q 6 1 if s =2  Proof. In this case, Formula (2.2) becomes

m1 1 − j tA t α1t α2t αst e = e B j +e B + +e Bs j! 1 20 ··· 0 Xj=0 where, for j =2,...,s, 1 Bj0 = Lj0(A)[P ]= Pj (A)gj (αj )= Pj (A) Pj (αj )

On the other hand, for j =0,...,m 1, we have B j = L j (A)[P ] where 1 − 1 1 m1 1 j − − j 1 (i) i L j(x)[P ]= P (x)(x α ) g (α )(x α ) , 1 1 − 1 i! 1 1 − 1 Xi=0 s P (x)= (x αl) 1 − Yl=2 1 g1(x)= s (x αl) − lQ=2 But since s al g1(x)= x αl Xl=2 − we have s i (i) ( 1) i!al g1 (x)= − i+1 . (x αl) Xl=2 − Then s m1 1 j s − − i j ( 1) al i L1j (x)[P ] = (x α1) (x αl) − i+1 (x α1) . − − (α1 αl) − Yl=2 Xi=0 Xl=2 − THE EXPLICIT MATRIX EXPONENTIAL 13

Therefore

s m1 j 1 s − − i j ( 1) al i B1j = (A α1I) (A αlI) − i+1 (A α1I) . − − (α1 αl) − Yl=2 Xi=0 Xl=2 − Thus, the proof is completed. 

To illustrate Corollary 2.11, consider the following example. Example 3. Let 11 0 0 11 0 0  A = 2 3 1 1  −  11 1 1  −  2 The characteristic polynomial of A is χA(x) = x (x + 2)(x 2). Let us choose, for example, − α1 = 0, α2 = 2 and α3 = 2. Then, in light of Formula (2.5), the exponential matrix of A is given by − tA 2t 2t e = B10 + tB11 +e− B20 +e B30 (2.6) where a2 a3 a2 a3 B10 = (A +2I)(A 2I) ( + )I ( 2 + 2 )A  − h α1 α2 α1 α3 − (α1 α2) (α1 α3) i  − − − −  a2 a3 B11 = A(A +2I)(A 2I) ( + ) − h α1 α2 α1 α3 i  − − and  1 1 2 1 3 1 2 B20 = P2(A)= − A (A 2I)= − A + A  P2(α2) 16 − 16 8  1 1 1 1  B = P (A)= A2(A +2I)= A3 + A2 30 P (α ) 3 16 16 8  3 3  1 1 1 1 In this case a = = − and a = = . Consequently, we have 2 α α 4 3 α α 4 2 − 3 3 − 2 1 2 B10 = − A + I  4 1  B = A3 + A  11 −  4  1 1  B = − A3 + A2 20 16 8  1 3 1 2  B30 = A + A  16 8  Hence Formula (2.6) becomes  2t 2t 2t 2t e e− t e e− 1 etA = ( − )A3 + ( + )A2 + tA + I 16 − 16 − 4 8 8 − 4 Since 22 0 0 22 0 0  A2 = 43 2 2  −  3 4 2 2   −  and 44 0 0 44 0 0 A3 =   9 11 4 4  −  5 3 4 4  −  14 MOHAMMED MOUC¸OUF AND SAID ZRIAA we obtain e2t +1 e2t 1 − 0 0  2 2  2t 2t  e 1 e +1   − 0 0  etA =  2 2   2t 2t 2t 2t 2t 2t  17e e− 4t 16 17e 5e− +4t 12 e− +1 e− +1  − − − − − −     16 16 2 2   2t 2t 2t 2t 2t 2t  11e +e− 4t 12 11e + 5e− +4t 16 e− +1 e− +1   − − − −   16 16 2 2  The following result is due to W. A. Harris et al. [3]

Corollary 2.12. If χA(x) = (x α )(x α ) (x αk) has distinct roots α , α ,...,αk, then − 1 − 2 ··· − 1 2 tA α1t α2t αkt e =e B +e B + +e Bk 1 2 ··· k x αj 2 where Bi = j=1,j=i − . Moreover, BiBj = Bj Bi =0 if i = j, and Bi = Bi. 6 αi αj 6 Q − Proof. Is a particular case of Theorem 2.8. 

When A is a matrix with only one eigenvalue, we have the following known result shown by Apostol [1]

k Corollary 2.13. If χA(x) = (x α) has a single root α, then − k 1 tA αt αt t − αt e =e B + te B + + e Bk 1 2 ··· (k 1)! − i 1 where Bi = (A αI) − , 1 i k. − ≤ ≤ Proof. Follows immediately from Theorem 2.8. 

3. Closed-form formula for the nth power of a matrix and Drazin inverses In this section, we present some applications of the previous results. We begin by proposing a new closed-form formula for the nth power of an arbitrary complex matrix A. We then deduce another closed-form formula for the nth power of the Drazin inverse AD of A. Also, we will be interested in determining explicit formulas for the Chevalley–Jordan decomposition and the spectral projections of A. Our contribution does not consist only in giving these formulas, but also in making their determination much practical and expressing them in an elegant representations.

m1 m2 ms Theorem 3.1. Let A be a k k matrix, and let χA(x) = (x α1) (x α2) (x αs) ,α1 = 0, be its characteristic polynomial× (possibly m =0). Then− for every−n N,··· we have− 1 ∈ m1 1 m2 1 ms 1 n − − n n j − n n j A = δnj B j + α − B j + + α − Bsj (3.1) 1 j 2 2 ··· j s Xj=0 Xj=0 Xj=0 where Bjkj = Ljkj (A)[χA] and δnj denotes the Kronecker symbol. Proof. Using Theorem 2.8, we can easily prove this result. 

Remark 3.2. It is easy to verify the following

Bjkj Bjmj kj =0 − Using the previous theorem we find a new method to calculate the Chevalley–Jordan decom- position and the spectral projections of A at the same time. THE EXPLICIT MATRIX EXPONENTIAL 15

m1 m2 Theorem 3.3. Let A be a k k matrix, and let χA(x) = (x α1) (x α2) (x ms × − − ··· − αs) , α1 =0, be its characteristic polynomial (possibly m1 =0). Then the Chevalley–Jordan decomposition of A is A = D + N (3.2) where N = B + B + + Bs (3.3) 11 21 ··· 1 and D = α B + α B + + αsBs (3.4) 2 20 3 30 ··· 0 where Bjr = Ljr(A)[χA] for j = 1, 2,...,s and r = 0, 1. Moreover, B10,B20,...,Bs0 are the spectral projections of A at α1, α2,...,αs, respectively. Proof. To obtain A = D + N and I = B + B + + Bs , (3.5) 10 20 ··· 0 it suffices to take n = 1 and n = 0 in Formula (3.1), respectively. Since Bj0,Bj1, j =1, 2,...,s, are polynomials of the matrix A, we have DDNN = NNDD. On the other hand, it is clear that Bj0Bi0 = Bi0Bj0 = 0,i = j, is a particular case of For- 26 mula (2.3). Multiplying both sides of (3.5) by Bj0, we get Bj0 = Bj0, j =1, 2,...,s. Furthermore, the matrix Bj0 is diagonalizable, then so is αj Bj0. The fact that αj Bj0 and αiBi0 commute assures then that D is diagonalizable. To complete the proof, it remains to show that the matrix N is nilpotent. To see this, it suffices to verify, using Formula (1.1), that each Bj1 is nilpotent.  Remarks 3.4.

1. We can immediately find that Bj0, j =1, 2,...,s, are the spectral projections of A by using a result of [19]. 2. Formulas (3.2),(3.3), and (3.4) can be also shown using the matrix equation (2.4). The following example is an illustration of Theorem 3.3. Example 4. Let us consider the matrix used in Example 2 2 0 1 A = 0 2 0 0 0 3   The Chevalley–Jordan decomposition of this matrix is A = D + N where

D =2B10 +3B20 and N = B11 A trivial calculation yields 1 0 1 − B10 = L10(A)[χA]= 01 0  00 0   0 0 0 B11 = L11(A)[χA]= 0 0 0 0 0 0   0 0 1 B20 = L20(A)[χA]= 0 0 0 0 0 1   16 MOHAMMED MOUC¸OUF AND SAID ZRIAA

The following Theorem shows that knowing only the associated matrices Bij of A, we can simply provide the minimal polynomial of the matrix A and the positive powers of its Drazin inverse.

m1 m2 m Theorem 3.5. Let A be a matrix and let χA(x) = (x α ) (x α ) (x αs) s be its − 1 − 2 ··· − characteristic polynomial with α1 =0 (possibly m1 =0). Then

1. The index of αi is the greatest integer j such that Bij 1 =0. − 6 2. The index of αi is 1 if and only if Bi1 =0. 3. A is diagonalizable if and only if Bi1 =0,i =1, 2,...,s. 4. For all positive integer n, we have

m2 1 ms 1 n − n n j − n n j A = − α− − B j + + − α− − Bsj D  j  2 2 ···  j  s Xj=0 Xj=0

where AD denotes the Drazin inverse of A. Proof. Clearly formula (3.1) is a -canonical form of A (see [19]) and the result follows then from Corollary 3.6. and TheoremP 3.9. of [19] and Theorem 3.1 above. 

Remark 3.6. Let rj be the index of αj . In view of Property 1. of Theorem 3.5, Formula (2.2) becomes s rj 1 − kj tA t αj t e = e Bjkj kj ! Xj=1 kXj =0 To constitute an illustration of Theorem 3.5, let us consider the same example given in [20] Example 5. Let us determine the Drazin inverse and its powers of the following matrix 2 0 0 A =  1 1 1  −1 1 1 − − −  2 The characteristic polynomial of A is χA(x)= x (x 2). Using the formula of Theorem 3.5, we obtain for all−n 1 ≥ n n n A = − 2− B D  0  20 On the other hand, we have 1 B = A2 20 4 Simple calculation gives 1 00 B20 =  1 0 0 −0 00   Therefore n 2− 0 0 n n AD =  2− 0 0 , n 1 − 0 00 ≥   for n =1 the Drazin inverse of the matrix A is

1 2 0 0   A = 1 D 2 0 0 −   0 00   which is in agreement with the result in [20]. The following example appears in [21] it is used here to illustrate our result on the Drazin inverse THE EXPLICIT MATRIX EXPONENTIAL 17

Example 6. Let 1 1 1 1 − 0 1 1 1 A = 1 1− 1 2  −  1 1 1 1  −  2 be a matrix with characteristic polynomial χA(x)= x (x (2 + √2))(x (2 √2)). Using the formula of Theorem 3.5, we obtain for all n −1 − − ≥ n n n n n A = − (2 + √2)− B + − (2 √2)− B D  0  20  0  − 30 On the other hand, we have

3√2 4 2 √ B20 = 8− A (A (2 2)I)  − −  3√2+4 2 √ B30 = 8 A (A (2 + 2)I) − − Simple calculation gives 

1 √2 √2 1 −  4 4 4 4  4 3√2 3 2√2 2√2 3 4 3√2 −8 −4 4− −8 B20 =    4 √2 1 2√2 2√2 1 4 √2   − − − −   8 4 4 8   1 √2 √2 1   −   4 4 4 4  and 1 √2 √2 1 −  4 4 4 4  3√2+4 3+2√2 3 2√2 3√2+4 8 4 − −4 8 B30 =    4+√2 1+2√2 2√2 1 4+√2   − −   8 4 4 8   1 √2 √2 1   −   4 4 4 4  It follows that n n n A = (2+ √2)− B + (2 √2)− B , n 1 D 20 − 30 ≥ for n =1 the Drazin inverse of the matrix A is

2 2 2 2 − 1 7 10 10 7 AD = − 4 5 6 6 5  −  2 2 2 2  − 

4. Explicit formulas for logarithms of matrices In this section, we derive some explicit and elegant formulas for logarithms of matrices with the aid of the results of Section 3 and Theorem 4.2 of [16].

m1 Theorem 4.1. Let A be a k k nonsingular matrix, and let χA(x) = (x α1) (x m2 ms × β1 − βs − α2) (x αs) be its characteristic polynomial where α1 = e ,...,αs = e . Then the matrix··· − m s s p 1 j 1 − ( 1) − βpBp0 + − j Bpj , (4.1) Xp=1 Xp=1 Xj=1  jαp  where Bjkj = Ljkj (A)[χA], is a logarithm of A. 18 MOHAMMED MOUC¸OUF AND SAID ZRIAA

Proof. By Theorem 3.1, we have for all n N ∈ m1 1 ms 1 − − n n β1(n j) n βs(n j) A = e − B j + + e − Bsj j 1 ··· j  Xj=0 Xj=0 If we plug t into this formula for n, we obtain the following smooth matrix function

m1 1 ms 1 − − t β1(t j) t βs(t j) A(t)= e − B j + + e − Bsj . j 1 ··· j Xj=0 Xj=0 Since t t(t 1) (t j + 1) = − ··· − , j j! t for all positive integer j, the derivative at 0 of the function j is j 1  t ′ ( 1) − (0) = − . j j t Consequently, using the convention that 0 = 1, we obtain  s s mp 1 j 1 − ( 1) − A′(0) = βpBp0 + − j Bpj . Xp=1 Xp=1 Xj=1 jαp Then the conclusion follows from Theorem 4.2 of [16].  As an illustration of Theorem 4.1, consider the following example. Example 7. Let 3 0 0 A = 2 3 0 1 4 2  2 The characteristic polynomial of A is χA(x) = (x 3) (x 2). Applying Formula (4.1), we obtain that − − 1 C = ln(3)B + ln(2)B + B 10 20 3 11 is the principal logarithm of A, where 1 j − j i i B j = (A 3I) (A 2I) ( 1) (A 3I) , j =0, 1  1 − − − −  Xi=0 B = (A 3I)2 20 − Simple calculation gives 2 B10 = A +6A 8I −2 −  B11 = A 5A +6I  B = A2 − 6A +9I 20 − Finally, we obtain  ln(3) 0 0  2  C = 3 ln(3) 0 7 ln( 2 )+ 8 4 ln( 3 ) ln(2)  3 3 2  Corollary 4.2. If A is a nonsingular matrix and χA(x) = (x α1)(x α2) (x αk) its β1 −βk − ··· − characteristic polynomial with distinct roots α1 =e ,...,αk =e , then the matrix k βj Bj , Xj=1 k 1 where Bj = (A αiI), is a logarithm of A. i=1,i=j αj αi − Q6 − THE EXPLICIT MATRIX EXPONENTIAL 19

Proof. Is a direct consequence of Theorem 4.1 and the fact that in the present case,

k 1 Bj0 = (A αiI). αj αi − i=1Y,i=j 6 − 

Example 8. To illustrate the previous corollary, we consider the same matrix that in [14]

5 1 0 16 −32   A = 1 1 0 −   0 0 1    1 1 1 1 The characteristic polynomial of A is χA(x) = (x 1)(x ( + i))(x ( i)). − − 2 4 − 2 − 4 Let α = 1, α = 1 + 1 i and α = 1 1 i. It is clear that the eigenvalues of A are all not 1 2 2 4 3 2 − 4 in R−, then the principal logarithm of A exists. To obtain this logarithm we must take the √5 1 √5 1 principal logarithm of α1, α2 and α3 which are 0,ln( 4 )+ i arctan( 2 ) and ln( 4 ) i arctan( 2 ) respectively. Applying the previous result, we obtain that −

B = β2B2 + β3B3 (4.2) is the principal logarithm of A, where 1 B2 = (A α1I)(A α3I)  (α2 α1)(α2 α3) − −  − −  1  B = (A α I)(A α I)  3 (α α )(α α ) − 1 − 2  3 1 3 2  − − √5 1 β2 = ln( 4 )+ i arctan( 2 )   √5 1  β3 = ln( ) i arctan( )  4 − 2  Simple calculation gives

1 5 i 2 + i −8 i 16    B = 1 1 i  2 2i 2 i −20 + 10   −   0 0 0         1 i 5 i i 2 − 8 − 16  1 1 i   B3 = 2i + i − + −  − 2 20 10      0 0 0      Substituting B2,B3,β2 and β3 by their values in (4.2), we obtain

5 1 10 0 2 − √ − 4 8 5  1  1  1  B = ln 0 1 − + arctan 4 2 − 4 10 2 − 5       00 0   0 0 0      Remark 4.3. Let A be a nonsingular matrix of order 2 and α1, α2 its eigenvalues. log(α ) log(α ) α log(α ) α log(α ) i) If α = α then log(A)= 1 − 2 A + 1 2 − 2 1 I. 1 6 2 α α α α 1 − 2 1 − 2 1 ii) If α = α = α then log(A)= A + (log(α) 1)I. 1 2 α − 20 MOHAMMED MOUC¸OUF AND SAID ZRIAA

k Corollary 4.4. If A is a nonsingular matrix with characteristic polynomial χA(x) = (x α) , α =eβ, then the matrix − k 1 j 1 − ( 1) − βB + − B , 0 jαj j Xj=1 j where Bj = (A αI) , is a logarithm of A. − Proof. Follows immediately from Theorem 4.1.  Example 9. For any complex number a, let us calculate the logarithm of the following matrix 1 0 0 ···  a 1 0 a2 2a ··· 0  3 2 ···   a 3a 0 A(a)=  4 4 ···   a4 a3 0  0 1 ···   . . .  . . .  k 1 k 1 k 1 k 2 ···   − a − − a − 1  0 1 ···  The characteristic polynomial of A(a) isχ (x) = (x 1)k. A(a) − In this example we have α =e0. Applying the previous result, we obtain that

k 1 j 1 − ( 1) − B(a)= − (A(a) I)j j − Xj=1 is the principal logarithm of A(a). We have j j j j i i (A(a) I) = ( 1) − A(a) − i − Xi=0 It is easily seen that A(a)i = A(ia) Thus j j j j i (A(a) I) = ( 1) − A(ia) − i − Xi=0 The (l, q)-th entry of the matrix A(ia) is l 1 l q q−1 (ia) − if l q (A(ia))lq = ≥  0,−  otherwise , Therefore l 1 l q j j j i l q j q−1 a − i=0 i ( 1) − i − if l q (A(a) I) )lq = − ≥ −  0,−  P  otherwise , It is easy to show that if l q, then ≥ j j j i l q qa if l = q +1 ( 1) − i − = i −  0, otherwise , Xi=0 so the principal logarithm of A(a) is 00 0 0 0 ··· a 0 0 0 0 0 2a 0 0 ··· 0  ···  B(a)= 0 0 3a 0 0  ···  ......  ......    0 0 0 (k 1)a 0  ··· −  The following interesting corollaries are consequence of Theorem 4.1. THE EXPLICIT MATRIX EXPONENTIAL 21

m1 m2 β1 β2 Corollary 4.5. If χA(x) = (x α1) (x α2) , α1 = e and α2 = e are nonzero two distinct complex numbers, then the− matrix −

m1 1 j 1 m2 1 j 1 − ( 1) − − ( 1) − β1B10 + β2B20 + − j B1j + − j B2j Xj=1 jα1 Xj=1 jα2 is a logarithm of A, where

m1 j 1 i m2+i 1 − − ( 1) − j m2 m2 1 i B1j = (A α1I) (A α2I) − − (A α1I)  − − (α α )m2+i −  Xi=0 1 − 2  m2 j 1 i m1+i 1  − − ( 1) −  m1 j m1 1 i B j = (A α I) (A α I) − − (A α I) 2 1 2 m1+i 2 − − (α2 α1)  −  Xi=0 −   m1 β1 β2 Corollary 4.6. If χA(x) = (x α1) (x α2), α1 =e and α2 =e are nonzero two distinct complex numbers, then the matrix− −

m1 1 j 1 − ( 1) − β1B10 + β2B20 + − j B1j Xj=1 jα1 is a logarithm of A, where

m1 j 1 − − i j ( 1) i B j = (A α I) (A α I) − (A α I)  1 − 1 − 2 (α α )i+1 − 1  Xi=0 1 2  −  1 m1 B20 = (A α1I) (α α )m1 −  2 1  − More generally, we have the following result. s m1 β1 βs Corollary 4.7. If χA(x) = (x α ) (x αj ), α =e ,...,αs =e are nonzero distinct − 1 − 1 jQ=2 complex numbers, then the matrix

m1 1 j 1 s − ( 1) − βj β1B10 + − j B1j + Pj (A) (4.3) Pj (αj ) Xj=1 jα1 Xj=2 is a logarithm of A, where P = χA,

s m1 j 1 s − − i j ( 1) al i B1j = (A α1I) (A αlI) − i+1 (A α1I) − − (α1 αl) − Yl=2 Xi=0 Xl=2 − and 1 s if s 3 Q (αl αp) ≥ a =  p=2,p=l − l  6 1 if s =2  Example 10. We now illustrate this case by computing the principal logarithm of the following matrix 12410 01200 A = 00210   00021   00023   3 The characteristic polynomial of A is χA(x) = (x 1) (x 2)(x 4). Using the result of the previous corollary, we obtain− that th−e matrix− 1 ln(2) ln(4) C = B11 B12 + P2(A)+ P3(A) − 2 P2(2) P3(4) 22 MOHAMMED MOUC¸OUF AND SAID ZRIAA

is the principal logarithm of A. On the other hand, we can easily check that B = (A I)(A 2I)(A 4I)( 4 A 1 I) 11 − − − 9 − 9  B = 1 (A I)2(A 2I)(A 4I)  12 3 − − −   P = (A I)3(A 4I) 2 − − 3  P3 = (A I) (A 2I)  − − Thus the principal logarithm of A is

0 2 8ln(2) 4 130 ln(2) 26 19 86 ln(2)  − 27 − 9 9 − 27    00 2ln(2) 11 ln(2) 4 2 7 ln(2)   9 − 3 3 − 9      C =  5 1  0 0 ln(2) 6 ln(2) 6 ln(2)   −     2 2  00 0 ln(2) ln(2)   3 3      00 0 4 ln(2) 4 ln(2)   3 3    5. Conclusion We have presented a new and elegant method to facilitate the computation of the matrix exponential function. In addition, a closed-form formula for the nth power of an arbitrary complex matrix A is provided. This formula allows us to deduce the Chevalley–Jordan de- composition and the spectral projections of A. We also deduce from this formula explicit and elegant formulas for the computation of the logarithm and the Drazin inverse of matrices.

References [1] T. M. Apostol, Some explicit formulas for the exponential matrix etA. The American Mathematical Monthly 76 (1969), 284–292. [2] R. Bellman, On the calculation of matrix exponential. Linear and Multilinear Algebra. 1983;13(1):73–79. [3] W. A. Harris, J. P. Fillmore, and D. R. Smith, Matrix Exponentials-Another Approach. SIAM review 43 (2001), 694–706. [4] E. I. Leonard, The matrix exponential. SIAM review 38 (1996), 507–512. [5] E. Liz, Classroom note: A note on the matrix exponential. SIAM review 40 (1998), 700–702. [6] C. Moler, and C. V. Loan, Nineteen dubious ways to compute the exponential of a matrix. SIAM review 20 (1978), 801–836. [7] J. L. Howland, Further ways to approximate the exponential of a matrix. Linear and Multilinear Algebra. 1983;14(2):121–129. [8] E. J. Putzer, Avoiding the Jordan canonical form in the discussion of linear systems with constant coeffi- cients. The American Mathematical Monthly 73 (1966), 2–7. [9] S. L. Campbell, The Drazin inverse and systems of second order linear differential equations. Linear and Multilinear Algebra. 1983;14(2):195–198. [10] C. Bu, K. Zhang, & J. Zhao, Representations of the Drazin inverse on solution of a class singular differential equations. Linear and Multilinear Algebra. 2011;59(8):863–877. [11] Y. Wei, Expressions for the Drazin inverse of a 2×2 block matrix. Linear and Multilinear Algebra. 1998;45(2-3):131–146. [12] N. J. Higham, Functions of matrices: Theory and applications. SIAM, Philadelphia, 2008. [13] C. D. Ahlbrandt, J. Ridenhour, Floquet theory for time scales and Putzer representations of matrix loga- rithms. Journal of Difference Equations and Applications. 2003;9(1):77–92. [14] J. A. Marrero, R. B. Taher, & M. Rachidi, On explicit formulas for the principal matrix logarithm. Applied Mathematics and Computation. 2013;220:142–148. [15] J. R. Cardoso, An explicit formula for the matrix logarithm. S Afr Optom. 2005;64(3):80–83. [16] M. Mou¸couf, An explicit expression for the minimal polynomial of the Kronecker product of matrices. Explicit formulas for matrix logarithm and matrix exponential. arXiv: 2010.11873v3 [math.RA](2021). [17] A. Spitzbart, A generalization of Hermite’s formula. The American Mathematical Monthly 67 (1960), 42–46. THE EXPLICIT MATRIX EXPONENTIAL 23

[18] M. Mou¸couf, and S. Zriaa, A new approach for computing the inverse of confluent Vandermonde matrices via Taylor’s expansion. Linear and Multilinear Algebra. 2021. DOI:10.1080/03081087.2021.1940807. [19] M. Mou¸couf, P-canonical forms and Drazin inverses. arXiv: 2007.10199v4 [math.RA](2021). [20] L. Zhang, A characterization of the Drazin inverse. and its Applications. 2001;335:183–188. [21] R. E. Hartwig, A method for calculating Ad. Math Japonica. 1981;26(1):37–43.

Mohammed Mouc¸ouf and Said Zriaa, University Chouaib Doukkali. Department of Mathematics, Faculty of science Eljadida, Morocco Email address: [email protected] Email address: [email protected]