You are on page 1of 12

733

Semiclassical position and


momentum information entropy
for sech2 and a family of rational
potentials
Mark W. Coffey

Abstract: The classical and semiclassical position and momentum information entropies for
the reflectionless sech2 potential and a family of rational potentials are obtained explicitly.
The sum of these entropies is of interest for the entropic uncertainty principle that is stronger
than the Heisenberg uncertainty relation. The analytic results relate the classical period of the
motion, total energy, position and momentum entropy, and dependence upon the principal
quantum number n. The logarithmic energy dependence of the entropies is presented. The
potentials considered include as special cases the attractive delta function and square well.
PACS Nos.: 03.67−a, 03.65.Sq, 03.65.Ge, 03.65.−w

Résumé : Entropies explicites de position et de moment classiques et semi-classiques pour


le potentiel sans réflexion en sech2 et pour une famille de potentiels rationnels. La somme
de ces entropies est utile pour le principe d’incertitude entropique qui est plus fort que
les relations d’indétermination de Heisenberg. Les résultats analytiques relient la période
classique du mouvement, l’énergie totale et l’entropie de position et de moment, ainsi que la
dépendance sur le nombre quantique principal n. Nous illustrons la dépendance logarithmique
en énergie des entropies. Parmi les cas spéciaux à l’étude figurent le potentiel attractif d’une
fonction delta et le potentiel d’un puits carré.
[Traduit par la Rédaction]

Introduction
The sech2 reflectionless potential has played a significant role in studies of inverse scattering [1, 2].
This potential permits an analytic solution of the Schrödinger equation, which is one reason that it has
served as a key motivating example. This attractive potential is well known to give a forward quantum
problem solvable by operator factorization, use of the Darboux transformation, or by other methods
[1, 2]. The sech2 potential has a vanishing reflection coefficient, so that some of the manipulations
in the inverse procedure are simplified. In this paper, we calculate the corresponding classical and
semiclassical entropies of both position and momentum for this potential. In addition, we calculate
explicitly the position entropy for a family of rational potentials. These new results complement very

Received 17 April 2006. Accepted 4 April 2007. Published on the NRC Research Press Web site at
http://cjp.nrc.ca/ on 18 July 2007.
M.W. Coffey. Department of Physics, Colorado School of Mines, Golden, CO 80401, USA
(e-mail: mcoffey@mines.edu).

Can. J. Phys. 85: 733–743 (2007) doi: 10.1139/P07-062 © 2007 NRC Canada
734 Can. J. Phys. Vol. 85, 2007

recent work restricted to power-law and logarithmic potentials.1 Lately information theoretic concepts
have been playing a larger role in quantum mechanics and quantum computing. Information concepts
have been employed in both fundamental discussions and in practical applications including synthesis
and analysis of electron densities in position and momentum space [3]. Indeed, the sum of quantum
position and momentum entropies has been advocated as a measure of wave-function quality [3].
Quantum entropy provides a quantitative description about the uncertainty or lack of knowledge
of an observable. For instance, the entropy is one measure of the delocalization of a wave packet. The
Shannon entropy provides an unambiguous measure that is complementary to the information content
of a system. With the recent research into quantum computing [4, 5] there has been added attention on
the fundamental physical limits of computation [6–8] and here also the quantum entropy plays a role.
We recall that Hirschman [9] anticipated a strengthened quantum uncertainty principle. Later,
Deutsch [10] and others [11] showed that the Heisenberg inequality does not properly express the
quantum uncertainty principle and is generally too weak. They introduced entropy measures such as we
use for noncommuting observables.
In this article, we establish the semiclassical position and momentum information entropy first for
the sech2 potential. This is an important instance since it applies to a potential with nonpolynomial
form. It also includes the case of an attractive delta-function potential for certain limit values of the
potential parameters of strength and width. The resulting semiclassical entropy relations have high
utility. This is because otherwise numerical computation may be required, or when closed form results
for quantum systems are available, the multiple sum and product expressions do not readily yield
physical information. We then treat a certain family of rational potentials. The current investigation
carries over to other classes of potentials, giving insight into the effect on information content of other
physical interactions [12].
A further motivation of our investigation is a recent presentation of the ground-state position entropy
of the Pöschl–Teller potential [13]. The hyperbolic form of this potential is none other than the sech2
functional dependence that we employ. The study, ref. 13 gave some numerical results for excited states,
although we note that the exact general excited state solution is expressible in terms of products of
operators Op = d/dx − p tanh αx [2]. Since asymptotic relations are now known connecting quantum
entropies to classical counterparts for both position [14] and momentum [15], we are able to derive
semiclassical expressions applicable for highly excited states.
The entropic uncertainty relation gives a lower bound to the sum of position S (x) and momentum
(p)
S entropies [11],
(x) (p)
SQ + SQ ≥ D(1 + ln π) (1)
for a D-dimensional system. This inequality stresses the reciprocity of position and momentum spaces.
For if the wave function is concentrated in coordinate space, it will necessarily be more diffuse in
momentum space, and vice versa. Not surprisingly, the bound in (1) is attained by Gaussian wave
functions [11]. Relation (1) extends to other pairs of noncommuting observables A and B, S (A) +S (B) ≥
sAB , where sAB is a positive constant. From (1) follows the Heisenberg uncertainty relation, showing
that this inequality is stronger.
The entropy sum appears in many other lower and upper bounds. Another lower bound, including
measurement device resolutions x of position and px of momentum, is [16]
 
(x) (p) xpx
SQ + SQ ≥ 1 − ln 2 − ln (2)
h
An upper bound to the entropy sum can be prescribed in terms of the second moments in position
and momentum space [3]. We note that the entropic uncertainty relation has been extended to nonzero

1
M.W. Coffey. Unpublished work. 2005.

© 2007 NRC Canada


Coffey 735

temperatures T . With β ≡ 1/kB T , for a one-dimensional thermal coherent state of frequency ω it


follows that [17]
(p)
SQ(x) + SQ = 1 + ln π + ln{cosh[2θ (β)]}

where

cosh[θ(β)] ≡ [1 − exp(−βω)]−1/2

In turn, the Heisenberg uncertainty relation at T  = 0 is a subcase and given by

xpx ≥ (1/2) cosh[2θ (β)], =1

In the case of the sech2 potential, we recall that the classical period of motion depends upon
total energy E as T (E) ∼ |E|−1/2 . We then determine how the classical position entropy SC(x) varies
logarithmically with the energy. By invoking an asymptotic relation [14], we then know how the quantum
entropy of position SQ(x) varies for the nth energy eigenstate. Since we also know the values of the
energy levels En , we determine explicitly the dependence of the semiclassical position entropy upon
(p) (p)
principle quantum number n. We find the momentum entropy SC , and therefore SQ for highly excited
states [15], to contain ln |E|−1/2 dependence, which is not surprising given that the Hamiltonian is
quadratic in momentum p. Below, we make all these relations quantitative. The calculations require
fairly advanced integration techniques. In particular, we extend several known results employing the
Gauss hypergeometric function, with certain details of the presentation relegated to an Appendix.
The investigation of the sech2 potential makes up the majority of this paper, although we discuss
how semiclassical results are also possible for a family of rational potentials. The treatment of this set of
potentials also relies heavily upon the analytic properties of the hypergeometric function. This family
contains a parameter q, such that the limit q → ∞ yields the finite square well. For q = 2 the rational
potential is very similar to the sech2 potential about the common minimum at x = 0.

Entropies of position and momentum for the sech2 potential


In this section, we assume a potential function V (x) = −U sech2 αx, where U > 0 gives the depth
and α determines the width of the potential. The limit of U → ∞ and α → ∞ such that U/α →
constant yields the attractive delta-function potential V (x) → −2U δ(αx) = −2(U/|α|)δ(x). (The
area under the potential, −2U/α, is preserved in this limit process.)
We first recall the classical period of motion T (E) for this potential, as it is needed in detail for the
position density and its derivation is of physical interest. The period is given by
 
2m x+ dx
T (E) = (3)
E x− [1 − V (x)/E]1/2

where m is the reduced mass and x± the classical turning points. By introducing v(x) ≡ cosh2 αx and
employing symmetry, the period can be written as
 
2m 1 −U/E dv
T (E) = √ (4)
E α 1 (v + U/E)(v − 1)
With the further change of variable y(v) = (1 − v)/(1 + U/E), a standard Beta function integral
B(1/2, 1/2) is obtained, giving

2m π
T (E) = (5)
|E| α
© 2007 NRC Canada
736 Can. J. Phys. Vol. 85, 2007

In the quantum case, the well-known energy levels are given by


   2
2 α 2 1 8mU 1
En = − +1− n+ , n = 0, 1, 2, . . . nmax (6)
2m 2 2 α 2 2

These relations illustrate the semiclassical connection T (E) = h∂n/∂E in the limit of large quantum
number n.
The classical probability distribution in space (spatial density) is given by [18]

1 2m
PC (x) = √ , x− ≤ x ≤ x+ (7)
T (E) E − V (x)

and the classical position entropy is given by


 x+
(x)
SC = − PC (x) ln PC (x) dx (8)
x−

By substituting (7) into (8) and noting the normalization of the spatial density, we immediately have
 √  √  x+
(x) T (E) E 2m ln[1 + U/E cosh2 αx]
SC = ln √ + √  dx (9)
2m T (E) E 0 1 + U/E cosh2 αx

By performing two successive changes of variable, as was done in finding the period, we may rewrite
the integral in this equation as a sum of three integrals
 x+   
ln[1 + U/E cosh2 αx] i 1 dy U
 dx = √ − ln 1 − 1 + y
0 1 + U/E cosh αx
2 2α 0 y(1 − y) E
 
U
+ ln 1 + + ln(1 − y) (10)
E

The second integral on the right of (10) is elementary and the third is given by
 1
y −1/2 (1 − y)−1/2 ln(1 − y) dy = B(1/2, 1/2)[ψ(1/2) − ψ(1)] = −2π ln 2 (11)
0

In finding (11), we have used the formula


 b
(x − a)µ−1 (b − x)ν−1 ln(x − a) dx = (b − a)µ+ν−1 B(µ, ν)[ln(b − a) + ψ(µ) − ψ(µ + ν)]
a
b > a, Reµ > 0, Reν > 0 (12)

which is easily derived from a known result in ref. 19, where ψ ≡ / is the digamma function, such
that ψ(1/2) = −γ − 2 ln 2, ψ(1) = −γ , and γ 0.577 215 is the Euler constant. The first integral
on the right of (10) may be evaluated by means of certain manipulations of the integral representation
of the Gauss hypergeometric function, some of the details of which are relegated to the Appendix. The
result takes a simple form
 1      
dy U U
√ ln 1 − 1 + y = 2π ln 1 + − ln 2 (13)
0 y(1 − y) E |E|

© 2007 NRC Canada


Coffey 737

By combining (9)–(13) we find for the classical entropy of position


    
π  1  U  U
(x)
SC (E) = ln + ln 1 +  − 2 ln 1 + (14)
α 2 E |E|

where (5) has been used for the period. The first term on the right of (14) is independent of the energy
E, while for a very deep potential the second and third terms cancel. On the other hand, for |U/E| 1,
it is possible to expand the entropy to write the approximation
  
π  1 U U
(x)
SC ≈ ln + −2
α 2 |E| |E|

For a broad potential, corresponding to small values of α, we physically expect a larger position
entropy since the wave function in space is less localized. On the other hand, for a deeper potential,
there is better confinement and we expect less position entropy. Equation (14) exhibits these features.
By invoking the asymptotic relation SQ(x) ∼ SC(x) − 1 + ln 2 [14] for the quantum entropy of position,
we have also demonstrated the dependence
    
π  1  U  U
(x) 
SQ (E) ∼ ln + ln 1 +  − 2 ln 1 + + constant (15)
α 2 E |E|

where the energies En ∼ (n + 1/2)2 for large n are given in (6). A benefit of our approach in obtaining
(15) and related results is that we avoid solving for the wave function ψ, forming the quantum probability
|ψ|2 , inserting into the expression for the quantum position entropy

(x)
SQ = − |ψ(x)|2 ln |ψ(x)|2 dx (16)

and then performing asymptotic evaluations.


In finding the classical momentum entropy
 p+
(p)
SC = − PC (p) ln PC (p) dp (17)
p−

where p± are the maximum and minimum values of momentum, we require the momentum distribution
[18]
2 1
PC (p) = (18)
T |F (x)|
where the force F (x) = −dV (x)/dx
√ must be√re-expressed in terms of the momentum p. For the sech2
potential, we have 1/|F (x)| = U /2αV (x) V (x) + U , so that

(2m)3/2 U 1
PC (p) =  (19)
αT (E) (p0 − p 2 ) p 2 + 2mU − p 2
2
0

where p02 ≡ 2mE. The substitution of (19) into (17) gives

(p) T (E)α
SC (E) = ln √
(2m)3/2 U
√  √2m(E+U )   2   
2 U (2m)3/2 ln p0 − p 2 + (1/2) ln 2m(E + U ) − p 2
+  2  dp (20)
αT (E) 0 p − p 2 2m(E + U ) − p 2
0

© 2007 NRC Canada


738 Can. J. Phys. Vol. 85, 2007

The change of variable p 2 = 2m(E + U )u brings the latter integral into the form
 √
2m(E+U )
  2   
ln p0 − p 2 + (1/2) ln 2m(E + U ) − p 2
 2  dp
0 p0 − p 2 2m(E + U ) − p 2
 1  2 
1 ln p0 − 2m(E + U )u + (1/2) ln {[2m(E + U )] (1 − u)}
=  √ du (21)
2p02 0 1 − 2m(E + U )u/p02 u(1 − u)
with one term given simply by
 1 
du |E|
  √ = π (22)
0 1 − 2m(E + U )u/p 2
0 u(1 − u) U
 
One other integral in (21) can be carried out by differentiating the hypergeometric function F α, 21 ; 1; z
with respect to the numerator parameter α and putting α = 1, similar to an example described in the
Appendix:
 1   
(1 − zu)−1 ∂  1
√ ln(1 − zu) du = −π F α, ; 1; z (23)
0 u(1 − u) ∂α α=1 2
The remaining integral in (21) requires differentiating F with respect to the denominator parameter:
 1
t −1/2 (1 − t)γ −3/2 (1 − tz)−1 ln(1 − t) dt
0
       
1 1 1 1 ∂ 1
=B ,γ − ψ γ− − ψ(γ ) F 1, ; γ ; z + F 1, ; γ ; z (24)
2 2 2 2 ∂γ 2 γ =1

where
   ∞
 k
∂  1 (1/2)k k  1
F 1, ; γ ; z = − z (25)
∂γ γ =1 2 k! j
k=1 j =1

(β)k is the Pochhammer symbol, and F (1, 1/2; 1; x) = 1/ 1 − x. In writing the final result for the
classical entropy of momentum, we combine (20)–(25), and put

  k 
k
(1/2)k U 1
L= 1+ (26)
k! E j
k=1 j =1

yielding
√    
(p) T (E)α π 2mU |E| mE
SC (E) = ln √ + ln + (1/2) ln [2m(E + U )] − 2L (27)
(2m)3/2 U αT (E)E U 2

The first term on the right of (27) has the expected dependence of ln |E|−1/2 , as the Hamiltonian is
simply quadratic in the momentum. Since the product αT (E) is independent of α, it appears that the
classical momentum entropy is independent of the width of the potential. It is known for strictly confining
(p) (p)
potentials [15] that SQ ∼ SC + constant. However, since the sech2 potential is asymptotically flat and
so has both discrete and continuous spectra, there may be another contribution to this asymptotic relation.
The resolution of the asymptotic relation between classical and quantum entropies of momentum for
general potentials remains an open problem.
© 2007 NRC Canada
Coffey 739

Fig. 1. The sequence of potentials V (x)/V0 = −1/[1 + (|x|/a)q ] obtained for varying values of q is plotted
versus x/a. For q → ∞ the square-well potential is obtained.
0

-0.1

-0.2

-0.3

-0.4
V(x)/V0

-0.5

-0.6

-0.7

-0.8

-0.9

-1.0
-5 -4 -3 -2 -1 0 1 2 3 4 5
x/a

Classical and semiclassical entropy of position for a family of rational


potentials
In this section, we consider potentials of the form V (x) = −V0 /[1 + (|x|/a)q ], where V0 > 0.
This family contains useful special cases. When the exponent q = 2, we expect similarity with the
sech2 potential just treated. In the limit of q → ∞, we recover the square-well potential, V = −V0 for
|x| < a and V = 0 otherwise. A sequence of such potentials is shown in Fig. 1, where the increasing
flatness of the potential near the minimum at x = 0 with increasing values of q is evident.
The period for the rational family of potentials may be calculated from (3), with the result employing
the hypergeometric function:
      
2a 2m V0 1/q−1/2 1 1 1 1 1 1 V0
T (E) = 1+ B , 2 F1 − , ; + ; 1 + (28)
q |E| E q 2 2 q 2 q E

When q = 2, the period can be rewritten in terms of a complete elliptic integral. Alternatively, the WKB
quantization condition [20] for the energies can be expressed in implicit form:
   
 a V0 1/q+1/2
1 3 1 1 1 3 V0 h 1
2 2m|E| − 1+ B , 2 F1 , ; + ;1 + = n(E) + (29)
q E q 2 2 q q 2 E 2 2

When employing the relation T (E) = h∂n/∂E, we have


   
∂ 1 1 1 3 V0 1 V0 3 1 1 5 V0
2 F1 , ; + ;1 + =− 2 F1 , + 1; + ; 1 + (30)
∂E 2 q q 2 E (2 + 3q) E 2 2 q q 2 E

Results such as (28)–(30) may be rewritten by applying transformation formulas (see, ref. 19, p. 1043)
to the 2 F1 function. These other forms correspond to changes of variable in the integral expressions.
© 2007 NRC Canada
740 Can. J. Phys. Vol. 85, 2007

The expansion B(1/2, 1/q) = q − γ + O(1/q) holds as q → ∞. Since √ the function 2 F1 reduces to
unity whenever a numerator parameter is zero, the period reduces to 2a 2m/(E + V0 ) for the square
well potential.
The classical entropy of position is given by (8). In carrying out this procedure, we may write

   
x+ V (x) 2m 1 x+ 1 y + (x+ /a)−q (1/q)−1
PC (x) ln 1 − dx = y
0 E E T (E) q 0 y−1
 
× ln(y − 1) − ln[y + (x+ /a)−q ] dy (31)

where the change of variable y = (x/x+ )q has been used and x+ is the right turning point. The
logarithmic integrals in (31) may be performed according to the techniques presented in the previous
section, taking appropriate derivatives of the hypergeometric function. The final result can be written as
      
(x) 2m 1 2m 1 a V0 (1/q)−(1/2) 1 1
SC (E) = ln + − 1+ B ,
E T (E) |E| T (E) q E q 2
       
V0 1 1 1 1 1 1 1 V0
× −iπ + ln − 1 + +ψ −ψ + 2 F1 − , ; + ;1 +
E 2 q 2 2 q q 2 E
  
∂  1 1 V0
+  2 F1 − , ; γ ; 1 +
∂γ γ =(1/q)+(1/2) 2 q E
  
∂  1 1 1 V0
+ 2 F1 α, ; + ;1 + (32)
∂α α=−1/2 q q 2 E

Once again the asymptotic relation SQ(x) ∼ SC(x) − 1 + ln 2 [14] yields the semiclassical position entropy.
It is possible to calculate the classical momentum entropy for the family of rational potentials,
although the expressions are lengthy. The momentum density (18) can be written as

2a  
PC (p) = V q−1 (V + V0 )1−q + 2(−1)q V −1 (V + V0 ) + V −(q+1) (V + V0 )1−q (33)
qV0 T (E)

where V (p) = (p02 − p 2 )/2m.

Summary
Information theoretic concepts are becoming of increasing interest in quantum mechanics and quan-
tum computing. The sum of quantum entropies of position and momentum is central to the entropic
uncertainty principle. This principle subsumes the Heisenberg inequality, including at nonzero temper-
atures. The semiclassical entropies derived in this paper are directly relevant to highly excited states and
supplement some existing numerical results [13]. In this paper, we have presented general expressions
for the classical and semiclassical entropies of position and momentum for the reflectionless sech2 po-
tential and a family of rational potentials parameterized by exponent q. These are important models that
also include as special cases the attractive delta function and square-well potentials. In particular, the
sech2 potential has played a key role in examples in inverse scattering theory [1, 2]. The sech2 potential
becomes asymptotically flat and does so sufficiently fast (exponentially) that there is a finite number
of bound states. By making use of an asymptotic relation between the classical and quantum position
entropies [14] we obtained the quantum entropy of position for energy eigenstates in the semiclassical
approximation. In the process of developing the entropy results, we may have obtained new algebraic
and logarithmic integrals, including those presented in the Appendix.

© 2007 NRC Canada


Coffey 741

The resulting semiclassical entropy relations have high utility. This is because otherwise numerical
computation may be required, or when closed form results for quantum systems are available, the mul-
tiple sum and product expressions do not readily yield physical information. The current investigation
carries over to other classes of potentials, giving insight into the effect on information content of other
physical interactions [12].
For the sech2 and rational potentials, we have been able to explicitly relate the corresponding system
period, energy, and entropy. We found how the classical and semiclassical entropies vary logarithmically
with total energy E. By knowing the values of the energy levels En in the quantum case, we have
determined explicitly the dependence of the semiclassical entropies upon the principle quantum number
n. Our semiclassical expressions
 x+ also permit, in principle, the expected values < x k > and < p k > to
p
k
be computed, as < x >≈ x− x PC (x)dx and < pk >≈ −p0 0 p k PC (p)dp. As an example, for the
k

family of rational potentials we find for even values of k


    
a k+1 2m 1 k+1 1 1 k+1 1 k+1 V0
< x k >= 2 B , 2 F1 − , ; + ;1 + (34)
q |E| T (E) q 2 2 q 2 q E

By symmetry, these moments are zero for odd values of k. In particular, the cases of k = 1 and 2 can
then be used to form the Heisenberg uncertainty product xpx .

Acknowledgement
This work was partially supported by Air Force Contract No. FA8750-06-1-001.

References
1. R.K. Dodd, J.C. Eilbeck, J.D. Gibbon, and H.C. Morris. Solitons and nonlinear waves. Academic Press,
New York. 1982.
2. G.L. Lamb, Jr. Elements of soliton theory. John Wiley, New York. 1980.
3. S.R. Gadre, S.B. Sears, S.J. Chakravorty, and R.D. Bendale. Phys. Rev. A, 32, 2602 (1985); S.R. Gadre
and R.D. Bendale. Phys. Rev. A, 36, 1932 (1987); S.R. Gadre. Phys. Rev. A, 30, 620 (1984).
4. M.A. Nielsen and I.L. Chuang. Quantum computation and quantum information. Cambridge University
Press, cambridge, UK. 2000.
5. A.Yu. Kitaev, A. Shen, and M.N. Vyalyi. Classical and quantum computation. American Mathematical
Society, Providence, Rhode Island. 2002.
6. S. Lloyd. Nature, 406, 1047 (2000); Phys. Rev. Lett. 88, 237901 (2002).
7. M.W. Coffey. Phys. Lett. 304A, 8 (2002).
8. E.T. Jaynes. Phys. Rev. 106, 620 (1957); Phys. Rev. 108, 171 (1957).
9. I.I. Hirschman. Am. J. Math. 79, 152 (1957).
10. D. Deutsch. Phys. Rev. Lett. 50, 631 (1983).
11. I. Bialynicki-Birula and J. Mycielski. Commun. Math. Phys. 44, 129 (1975).
12. M.W. Coffey. J. Phys. A, 36, 7441 (2003).
13. R. Atre, A. Kumar, N. Kumar, and P.K. Panigrahi. Phys. Rev. A, 69, 052107 (2004).
14. J. Sánchez-Ruiz. Phys. Lett. 226A, 7 (1997).
15. M.W. Coffey. Phys. Lett. 324A, 446 (2004).
16. I. Bialynicki-Birula. Phys. Lett. 103A, 253 (1984); M.H. Portovi. Phys. Rev. Lett. 50, 1883 (1983).
17. S. Abe and N. Suzuki. Phys. Rev. A, 41, 4608 (1990).
18. R.W. Robinett. Am J. Phys. 63, 823 (1995).
19. I.S. Gradshetyn and I.M. Rhyzik. Table of integrals, series, and products. Academic Press, San Diego.
1980.
20. L.I. Schiff. Quantum mechanics. McGraw-Hill, New York. 1968.

© 2007 NRC Canada


742 Can. J. Phys. Vol. 85, 2007

Appendix A.
The derivation of the logarithmic integral (13) is herein discussed. By way of the integral represen-
tation of a special case of the hypergeometric function F ,
  
1 1 1 (1 − zt)−α
F α, ; 1; z = √ dt (A.1)
2 π 0 t (1 − t)
we have
  
∂ 1 1 1 (1 − zt)−α
− F α, ; 1; z = √ ln(1 − zt)dt (A.2)
∂α 2 π 0 t (1 − t)

with the integral of interest (13) corresponding to the case α = 0. We have2


  ∞ k−1
∂ 1 (α)k (1/2)k zk  1
F α, ; 1; z = (A.3)
∂α 2 (1)k k! (α + j )
k=1 j =0

where (β)k is the Pochhammer symbol, (a)j = (a + j )/ (a), with derivative


j −1

d 1
(a)j = (a)j [ψ(a + j ) − ψ(a)] = (a)j (A.4)
da (a + )
=0

and j is an integer. Furthermore, we have the property



∂ 
(α)k = (k − 1)!, k≥1 (A.5)
∂α α=0
and the series
∞   k
1 1 z √
= 2[ln 2 − ln(1 + 1 − z)], |z| ≤ 1 (A.6)
k 2 k k!
k=1

This series [19] follows since


 
1 1 · 3 · 5 · · · (2k − 1) (2k − 1)! 2(2k − 1)!
= = k = 2k (A.7)
2 k 2k 2 2 · 4 · 6 · 8 · · · (2k − 2) 2 (k − 1)!
From (A.2) and (A.6), with z = 1 + U/E, U > 0, E < 0, (13) follows.
More generally, following steps similar to above, we have the logarithmic integral
 1 
∂ 
t β−1 (1−t)γ −β−1 ln(1−zt) dt = −B(β, γ −β) F (α, β; γ ; z), Re γ > Re β > 0 (A.8)
0 ∂α α=0

where B is the Beta function. By using the property (β)k+1 = β(β + 1)k , we have
 1 β
t β−1 (1 − t)γ −β−1 ln(1 − zt) dt = − B(β, γ − β)z 3 F2 (1, 1, 1 + β; 2, 1 + γ ; z),
0 γ
Re β > −1, Re(γ − β) > 0 (A.9)

2
M.W. Coffey. Unpublished work. 2005.

© 2007 NRC Canada


Coffey 743

where p Fq is the generalized hypergeometric function [A.1.] and

β β (β) (γ − β) (β + 1)
B(β, γ − β) = = (γ − β) = B(β + 1, γ − β) (A.10)
γ γ (γ ) (γ + 1)
References

A.1. I.S. Gradshetyn and I.M. Rhyzik. Table of integrals, series, and products. Academic Press, San
Diego. 1980.

© 2007 NRC Canada

You might also like