Introduction To Ordinary Differential Equations
Introduction To Ordinary Differential Equations
Daniele Ritelli
December 2, 2021
ii
Contents
vii
viii CONTENTS
In the same period, Euler introduced a function e(x) , given by the explicit formula (1.2) and
defined for any x > 0 , which reduces to e(n) = n! when its argument is x = n ∈ N . Euler
described his results in a letter to Goldbach2 , who had posed, together with Bernoulli, the
interpolation problem to Euler. This last problem was inspired by the fact that the additive
counterpart of the factorial has a very simple solution:
n(n + 1)
sn = 1 + 2 + . . . · · · + n = ,
2
and by the observation that the above sum function admits a continuation to C given by the
following function:
x(x + 1)
f (x) = ,.
2
Euler’s solution for the factorial is the following integral:
1 x
Z 1
e(x) = ln dt . (1.2)
0 t
Legendre3 introduced the letter Γ(x) to denote (1.2), and he modified its representation as
follows: Z ∞
Γ(x) = e−t tx−1 dt , x > 0. (1.3)
0
Observe that (1.2) and (1.3) imply the equality Γ(x + 1) = e(x) . In fact, the change of variable
t = − ln u , in the Legendre integral Γ(x + 1) , yields:
1 x 1
Z 1
ln u
Γ(x + 1) = e ln du = e(x) .
0 u u
1
James Stirling (1692–1770), Scottish mathematician.
2
Christian Goldbach (1690–1764),German mathematician.
3
Adrien Marie Legendre (1752–1833), French mathematician.
1
2 CHAPTER 1. GAMMA AND BETA FUNCTIONS
These aspects are treated with the maximal generality in the Bohr–Mollerup4 Theorems 1.1–1.2.
Note that Γ(x) appears in many formulæ of Mathematical Analysis, Physics and Mathematical
Statistics.
Theorem 1.1. For any x > 0 , the recursion relation (1.5) is true. In particular, when x = n ∈
N , then (1.4) holds.
When x > 0 , function Γ(x) is continuous and differentiable at any order. To evaluate its
derivatives, we use the differentiation of parametric integrals, obtaining:
Z ∞
0
Γ (x) = e−t tx−1 ln t dt , (1.6a)
0
Z ∞
Γ(2) (x) = e−t tx−1 (ln t)2 dt , (1.6b)
0
From its definition, Γ(x) is strictly positive. Moreover, since (1.6b) shows that Γ(2) (x) ≥ 0 , it
follows that Γ(x) is also strictly convex. We thus infer the existence for the following couple of
limits:
`0 := lim Γ(x) , `∞ = lim Γ(x) .
x→0+ x→∞
To evaluate `∞ , since we know, a priori, that such limit exists, we can restrict the focus to
natural numbers, so that we have immediately:
Observing that Γ(2) = Γ(1) and using Rolle Theorem5 , we see that there exists ξ ∈ ] 1 , 2 [
such that Γ0 (ξ) = 0 . On the other hand, since Γ(2) (x) > 0 , the first derivative Γ0 (x) is
strictly increasing, thus there is a unique ξ such that Γ0 (ξ) = 0 . Furthermore, we have that
0 < x < ξ =⇒ Γ0 (x) < 0 and x > ξ =⇒ Γ0 (x) > 0 . This means that ξ represents the absolute
minimum for Γ(x) , when x ∈ ] 0 , ∞ [ . The numerical determination of ξ and Γ(ξ) is due to
Legendre and Gauss6 :
If we take:
t x−1
f (t) = e− 2 t 2 , g(x) = f (x) ln t ,
recalling (1.3), (1.6a) and (1.6b), we find the inequality:
2
Γ0 (x) ≤ Γ(x) Γ(2) (x) .
in particular:
1 1
Γ − = −2 Γ .
2 2
When x ∈ ] − 2 , −1 [ , the evaluation is:
Γ(2 − x)
Γ(x) = ,
(x + 1) x
thus, in particular:
3 4 1
Γ − = Γ .
2 3 2
In other words, Γ(x) is defined on the real line, except the singular points x = 0 , −1 , −2 , · · · ,
and so on, as shown in Figure 1.1.
y
x
-4 -3 -2 -1 1 Ξ 2 3 4 5
Notice that the change of variable t = 1 − s provides, immediately, the symmetry relation
B(x , y) = B(y , x) in (1.10), which yields:
Z π/2
B(x , y) = 2 cos2 x−1 ϑ sin2 y−1 ϑ dϑ . (1.10a)
0
The main property of the Beta function is its relationship with the Gamma function, as expressed
by Theorem 1.4 below.
Γ(x) Γ(y)
B(x , y) = . (1.11)
Γ(x + y)
1.2. BETA FUNCTION 5
Proof. From the usual definition (1.3) for the Gamma function, after the change of variable
t = u2 , we have: Z +∞
2
Γ(x) = 2 u2 x−1 e−u du .
0
In the same way: Z +∞
2
Γ(y) = 2 v 2 y−1 e−v dv .
0
Now, we form the product of the last two integrals above, and we use Fubini Theorem, obtaining:
ZZ
2 2
Γ(x) Γ(y) = 4 u2 x−1 v 2 y−1 e−(u +v ) du dv .
[0 ,+∞)×[0 ,+∞)
At this point, we change variable in the double integral, using polar coordinates:
(
u = ρ cos ϑ ,
v = ρ sin ϑ ,
Remark 1.5. Theorem 1.4 can be shown also using, again, Fubini Theorem, starting from:
Z ∞Z ∞
Γ(x) Γ(y) = e−(t+s) tx−1 sy−1 dt ds ,
0 0
1
1.2.1 Γ 2
and the probability integral
1
Using (1.11), we can evaluate Γ and, then, the probability integral. In fact, by taking x = z
2
and y = 1 − z , where 0 < z < 1 , we obtain:
Z 1 Z 1
z−1 −z t z 1
Γ(z) Γ(1 − z) = B(z , 1 − z) = t (1 − t) dt = dt .
0 0 1−t t
The following representation Theorem 1.6 decribes the family of integrals related to the Beta
function.
Observe that, with the change of variable 2 t = u in J , it follows that I = J . Observe, further,
that:
1 1 1
I = B x+ , ,
2 2 2
Z π
2
2x 2x−1 1 1
J= (2 sin t cos t) dt = 2 B x + ,x + .
0 2 2
Hence:
1 1 1 2 x−1 1 1
B x+ , =2 B x + ,x + . (1.17)
2 2 2 2 2
Recalling (1.11), equality (1.17) implies (1.16).
Formula (1.16) is generalised by Gauss multiplication Theorem 1.8, which we present without
proof.
From the reflexion formula (1.19), it follows immediately the computation of integral (1.20).
The same argument followed to demonstrate Theorem 1.11 can be applied to prove the following
Theorem 1.12, thus, we leave it as an exercise.
Theorem 1.12. If 2 a < n , then:
Z 1
xa−1 a Z ∞ z a−1
√ dx = cos π √ dz . (1.23)
0 1 − xn n 0 1 + zn
1
Proof. The change of variable 1 + xn = is employed in the left hand–side integral of (1.25),
t
1 1 1
and therefore dx = − t− n −1 (1 − t) n −1 dt :
n
Z ∞
1 1 −1 1 1 (1− 1 )−1
Z Z
dx 1
−1 1
n
= t n (1 − t) n dt = t n (1 − t) n −1 dt
0 1+x n 0 n 0
1 1 1 1 1 1
= B 1− , = Γ 1− Γ .
n n n n n n
Thesis (1.25) follows from reflexion formula (1.19). Formula (1.26) also follows, using an analo-
gous argument.
1 1 3 1
The change of variable 1 + x2 = , that is, dx = − t− 2 (1 − t)− 2 dt , leads to:
t 2
Z ∞ √
dx 1 1 π 1
= B n− , = Γ n− .
−∞ (1 + x2 )n 2 2 (n − 1)! 2
The assumption we made for the parameter b ensures summability. Hence, exploiting Fubini
Theorem, the integration can be performed regardless of the order. Let us integrate, first, with
respect to x : Z +∞
Z +∞
I(b , p) = y p−1 sin(b x) e−x y dx dy .
0 0
In this way, the integral above turns out to be an elementary one, since:
Z +∞
b
sin(b x) e−x y dx = 2 ,
0 b + y2
and then:
+∞
y p−1
Z
I(b , p) = b dy ,
0 b2 + y 2
y
for which, employing the change of variable t = b , we obtain:
∞
tp−1
Z
p−1
I(b , p) = b dt .
0 1 + t2
The latest formula allows to use identity (1.26) and complete the first computation:
π b p−1
I(b , p) = . (1.31)
2 sin p π2
The inner integral is immediately evaluated, in terms of the Gamma function, setting u = x y :
Z +∞ Z ∞
1 1
y p−1 e−x y dy = p u p−1 e−u du = p Γ(p) . (1.32)
0 x 0 x
Equating (1.32) and (1.31) leads to:
∞
b p−1 π
Z
sin(b x)
Γ(p) dx = ,
0 xp 2 sin p π2
The right hand–side integral, above, has the form (1.30), with b = 1 and p = 2 − 1q , therefore:
∞
sin xq
Z
π
dx = .
xq
0 1 1 π
2 q Γ 2− q sin 2− q 2
we arrive at:
1
Z ∞
sin xq Γ q sin πq
dx = .
0 xq 2 (q − 1) sin π
2q
We now show, employing again reversal integration, a cosine relation similar to (1.30), namely:
∞
π b p−1
Z
cos(b x)
dx = , (1.35)
0 xp 2 Γ(p) cos p π2
where we must assume that 0 < p < 1 , to ensure convergence of the integral, due to the
singularity in the origin. To prove (1.35), we consider the double integral:
Z ∞Z ∞
cos(b x) y p−1 e−x y dx dy ,
0 0
12 CHAPTER 1. GAMMA AND BETA FUNCTIONS
from which we show that (1.35) can be reached, via the Fubini Theorem regardless of the order
of integration. The starting point is, then, the equality:
Z ∞ Z ∞ Z ∞ Z ∞
p−1 −x y p−1 −x y
cos(b x) y e dy dx = y cos(b x) e dx dy . (1.36)
0 0 0 0
The last integral above is in the form (1.26). Therefore, the right hand–side integral of (1.36)
turns out to be: Z ∞
yp π b p−1 π b p−1
dy = = . (1.37)
b2 + y 2 2 cos p2 π
0 2 sin p+1 π 2
The inner integral in the left hand–side of (1.36) is given by (1.32). Hence, the left hand–side
integral of (1.36) is: Z ∞
cos(b x)
Γ(p) dx . (1.38)
0 xp
Equating (1.37) and (1.38) leads to (1.35).
There is a further, very interesting consequence of equation (1.30) leading to the evaluation of
the Fresnel8 integrals (1.39), that hold for b > 0 and k > 1 :
Z ∞
k 1 1 π
sin(b x ) dx = 1 Γ sin . (1.39)
0 k bk k 2k
To prove (1.39), we start from considering its left hand–side integral, inserting in it the change
xk
of variable u = xk , i.e., du = k xk−1 dx = k dx , thus:
x
Z ∞ Z ∞ 1
1 ∞ sin(b u)
Z
k 1 uk
sin(b x ) dx = sin(b u) du = 1 du .
0 0 k u k 0 u1− k
1
The latest integral above is in the form (1.30), with p = 1 − k . Hence:
1
∞
π b− k
Z
sin(b u)
1 du = , (1.40)
u1− k 2 Γ 1 − k1 sin 1 − k1 π
0
2
and, thus: Z ∞
π
sin(b xk ) dx = 1
1 π π
0 2 k b Γ 1−
k
k sin 2 − 2k
we obtain (1.39).
In (1.39), the particular choices b = 1 and k = 2 correspond to the sine Fresnel integral:
Z ∞ π rπ
1 1
sin(x2 ) dx = Γ sin = . (1.41)
0 2 2 4 8
Exploiting the same technique that produced (1.39) from (1.30), it is possible to derive the
cosinus analogue (1.42) of Fresnel integrals, that hold for b > 0 and k > 1 :
Z ∞
1 1 π
cos(b xk ) dx = 1 Γ cos . (1.42)
0 k bk k 2k
To prove (1.42), the starting point is (1.35). Then, as in the sine case, we introduce the change
of variable u = xk , we choose p = 1 − k1 , and, via calculations similar to those performed in
the sine case, we arrive at:
Z ∞
π
cos(b xk ) dx = 1 .
2 k b k Γ 1 − k1 cos π2 − 2πk
0
Definition 2.1. The real or complex number ` is called the limit of the infinite sequence (un )
if, for any positive number ε , there exists a positive number n(ε) , depending on ε , such that
|un − `| < ε for all integers n > n(ε) . In such a case, we denote:
lim un = ` .
n→∞
∞
X
Given a sequence (un ) , we say that its associated infinite series un :
n=1
n
X
(ii) diverges, when the limit of the partial sums uk does not exist.
k=1
Let us analyse what happens when n → ∞ . It is easy to realise that a sequence of continuous
functions may converge to a non–continuous function. Indeed, for the sequence of functions in
Example 2.3, it holds:
(
1 if x = 1 ,
lim fn (x) = lim xn =
n→∞ n→∞ 0 if 0 ≤ x < 1 .
15
16 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Thus, even if every function of the sequence fn (x) = xn is continuous, the limit function f (x) ,
defined below, may not be continuos:
The convergence of a sequence of functions, like that of Example 2.3, is called simple convergence.
We now provide its rigorous definition.
Definition 2.4. If (fn ) is a sequence of functions in I ⊆ [a , b] and f is a real function on I ,
then fn pointwise converges to f if, for any x ∈ I , there exists the limit of the real sequence
(fn (x)) and its value is f (x) :
lim fn (x) = f (x) .
n→∞
Pointwise convergence is denoted as follows:
I
fn −
→f .
I
Remark 2.5. Definition 2.4 can be reformulated as follows: it holds that fn − → f if, for any
ε > 0 and for any x ∈ I , there exists n(ε, x) ∈ N , depending on ε and x , such that:
We do not have the tools, yet, to evaluate the integral in the left hand–side of the above equality
(but we will soon), but it is clear that it is a positive real number, so we have:
Z ∞ Z ∞ Z ∞
2
lim fn (x)dx = e−y dy = α > 0 6= lim fn (x)dx = 0 .
n→∞ 0 0 0 n→∞
To establish a ‘good’ notion of convergence, that allows the passage to the limit, when we take the
integral of the considered sequence, and that preserves continuity, we introduce the fundamental
notion of uniform convergence.
Definition 2.6. If (fn ) is a sequence of functions defined on the interval I , then fn converges
uniformly to the function f if, for any ε > 0 , there exists nε ∈ N such that, for n ∈ N , n > nε ,
it holds:
sup |fn (x) − f (x)| < ε . (2.1)
x∈I
Remark 2.7. Definition 2.6 is equivalent to requesting that, for any ε > 0 , there exists nε ∈ N
such that, for n ∈ N , n > nε , it holds:
and this implies (2.2). Viceversa, if (2.2) holds then, for any ε > 0 , there esists nε ∈ N such
that:
sup |fn (x) − f (x)| < ε , for any n ∈ N , n > nε ,
x∈I
I
that is to say, fn ⇒ f .
Remark 2.8. Uniform convergence implies pointwise convergence. The converse does not hold,
as Example 2.3 shows.
In the next Theorem 2.9, we state the so–called Cauchy uniform convergence criterion.
Theorem 2.9. Given a sequence of functions (fn ) in [a , b] , the following statements are equiv-
alent:
(ii) for any ε > 0 , there exists nε ∈ N such that, for n , m ∈ N , with n , m > nε , it holds:
Proof. We show that (i) =⇒ (ii). Assume that (fn ) converges uniformly, i.e., for a fixed ε > 0 ,
ε
there exists nε > 0 such that, for any n ∈ N , n > nε , inequality |fn (x) − f (x)| < , holds for
2
any x ∈ [a , b] . Using the triangle inequality, we have:
ε ε
|fn (x) − fm (x)| ≤ |fn (x) − f (x)| + |f (x) − fm (x)| < + =ε
2 2
for n , m > nε .
To show that (ii) =⇒ (i), let us first observe that, for a fixed x ∈ [a , b] , the numerical
sequence (fn (x)) is indeed a Cauchy sequence, thus, it converges to a real number f (x) . We
prove that such a convergence is uniform. Let us fix ε > 0 and choose nε ∈ N such that, for
n , m ∈ N , n , m > nε , it holds:
|fn (x) − fm (x)| < ε
for any x ∈ [a , b] . Now, taking the limit for m → +∞ , we get:
Example 2.10. The sequence of functions fn (x) = x (1+n x)−1 converges uniformly to f (x) =
0 in the interval [0 , 1] . Since fn (x) ≥ 0 for n ∈ N and for x ∈ [0 , 1] , we have:
x 1
sup = →0 as n→∞.
x∈[0,1] 1 + n x 1+n
18 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Example 2.11. The sequence of functions fn (x) = (1 + n x)−1 does not converge uniformly to
f (x) = 0 in the interval [0 , 1] . In spite of the pointwise limit of fn for x ∈]0 , 1] , we have in
fact:
1
sup = 1.
x∈[0,1] 1 + nx
ln fn (x) = ln nα x e−αx = α ln n − n x .
It follows that lim ln fn (x) = −∞ and, then, lim fn (x) = 0 . Pointwise convergence is proved.
n→∞ n→∞
For uniform convergence, we show that, for any n ∈ N , the associate function fn reaches its
absolute maximum in R+ . By differentiating with respect to x , we obtain, in fact:
1
from which we see that function fn assumes its maximum value in xn = ; such a maximum
n
is absolute, since fn (0) = 0 and lim fn (x) = 0 . We have thus shown that:
x→+∞
1
sup fn (x) = fn = e−α nα−1 .
x∈R+ n
Now, lim sup fn (x) = lim e−α nα−1 = 0 , when α < 1 . Hence, in this case, convergence is
n→∞ x∈R+ n→∞
indeed uniform.
2.3. UNIFORM CONVERGENCE 19
In the following example, we compare two sequences of functions, apparently very similar, but
the first one is pointwise convergent, while the second one is uniformly convergent.
Example 2.13. Consider the sequences of functions (fn ) and (gn ) , both defined on [0 , 1] :
1
n2 x (1 − n x)
if 0 ≤ x < ,
fn (x) = n (2.3)
1
0 if ≤x≤1,
n
and
1
n x2 (1 − n x)
if 0≤x< ,
gn (x) = n (2.4)
1
0 ≤x≤1.
if
n
Sequence (fn ) converges pointwise to f (x) = 0 for x ∈ [0 , 1] ; in fact, it is fn (0) = 0 and
1
fn (1) = 0 for any n ∈ N . When x ∈ (0 , 1) , since n0 ∈ N exists such that < x , it follows
n0
that fn (x) = 0 for any n ≥ n0 .
1
The convergence of (fn ) is not uniform; to show this, observe that ξn = maximises fn ,
2n
since:
1
n2 (1 − 2 n x)
if 0 ≤ x < ,
fn0 (x) = n
1
0 if ≤x≤1.
n
It then follows:
n
sup |fn (x) − f (x)| = sup fn (x) = fn (ξn ) =
x∈[0,1] x∈[0,1] 4
which prevents uniform convergence. With similar considerations, we can prove that (gn ) con-
verges pointwise to g(x) = 0 , and that the convergence is also uniform, since:
1
n x (2 − 3 n x)
if 0 ≤ x < ,
0
gn (x) = n
1
0 if ≤x≤1,
n
2
implying that ηn = maximises gn and that:
3n
4
sup |gn (x) − g(x)| = sup gn (x) = gn (ηn ) = ,
x∈[0,1] x∈[0,1] 27 n
Using the continuity of fn , we can see that there exists δ > 0 such that:
ε
|fn (x) − fn (x0 )| < (2.6)
3
20 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Remark 2.17. Consider again the sequences of functions (2.3) and (2.4), defined on [0 , 1] , with
fn → 0 and gn ⇒ 0 . Observing that:
Z 1 Z 1
n 1
fn (x)dx = n x2 (1 − n x)dx =
0 0 6
and 1
Z 1 Z
n 1
gn (x)dx = n2 x(1 − n x)dx = ,
0 0 12 n2
it follows: Z 1 Z 1
1
lim fn (x)dx = =6 f (x)dx = 0 ,
n→∞ 0 6 0
while: Z 1 Z 1
1
lim gn (x)dx = lim =0= g(x)dx = 0 .
n→∞ 0 n→∞ 12 n2 0
In other words, the pointwise convergence of (fn ) does not permit the passage to the limit, while
the uniform convergence of (gn ) does.
We provide a second example to illustrate, again, that pointwise convergence, alone, does not
allow the passage to the limit.
Example 2.18. Consider the sequence of functions (fn ) on [0 , 1] defined by:
1
n2 x if 0 ≤ x ≤ ,
n
1 2
fn (x) = 2 n − n2 x if < x ≤ ,
n n
2
0 if < x ≤ 1 .
n
Observe that each fn is a continuous function. Plots of fn are shown in Figure 2.3, for some
values of n ; it is clear that, pointwise, fn (x) → 0 for n → ∞ .
y
Figure 2.3: Plot of functions fn (x) , n = 3 , . . . , 6 , in Example 2.18. Solid lines are used for even
values of n ; dotted lines are employed for odd n .
By construction, though, each triangle in Figure 2.3 has area equal to 1 , thus, for any n ∈ N :
Z 1
fn (x)dx = 1 .
0
In conclusion: Z 1 Z 1
1 = lim fn (x)dx 6= lim fn (x)dx = 0
n→∞ 0 0 n→∞
22 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
In presence of pointwise convergence alone, therefore, swapping between integral and limit is
not possible.
The hypotheses of Theorem 2.19 are essential, as it is shown by the following example.
Example 2.20. Consider the sequence (fn ) on the open interval ] − 1 , 1[ :
2x
fn (x) = , n ∈ N.
1 + n2 x2
Observe that fn converges to 0 uniformly in ] − 1 , 1[ , since:
1
sup |fn (x)| = −→ 0 .
x∈]−1,1[ n n→∞
Function fn is differentiable for any n ∈ N and, for any x ∈ ] − 1 , 1[ and any n ∈ N , the
derivative of fn , with respect to x , is:
2(1 − n2 x2 )
fn0 (x) = .
(1 + n2 x2 )2
Now, consider function g : ] − 1 , 1[ → R :
(
0 if x 6= 0 ,
g(x) =
2 if x = 0 .
]−1,1[
Clearly, fn0 −−−→ g ; such a convergence holds pointwise, but not uniformly; by Theorem 2.14,
in fact, uniform convergence of (fn0 ) would imply g to be continuous, which is not, in this case.
Here, the hypotheses of Theorem 2.19 are not fulfilled, thus its thesis does not hold.
We end this section with Theorem 2.21, due to Dini1 , and the important Corollary 2.23, a
consequence of the Dini Theorem, very useful in many applications. Theorem and corollary
connect monotonicity and uniform convergence for a sequence of functions; for their proof, we
refer the Reader to [10].
1
Ulisse Dini (1845–1918), Italian mathematician and politician.
2.4. SERIES OF FUNCTIONS 23
Theorem 2.21 (Dini). Let (fn ) be a sequence of continuous functions, converging pointwise to
a continuous function f , defined on the interval [a , b] .
Furthermore, assume that, for any x ∈ [a , b] and for any n ∈ N , it holds fn (x) ≥ fn+1 (x) .
Then fn converges uniformly to f in [a , b] .
Remark 2.22. In Theorem 2.21, hypothesis fn (x) ≥ fn+1 (x) can be replaced with its reverse
monotonicity assumption fn (x) ≤ fn+1 (x) , obtaining the same thesis.
Corollary 2.23 (Dini). Let (fn ) be sequence of nonnegative, continuous and integrable func-
tions, defined on R , and assume that it converges pointwise to f , which is also nonnegative,
continuous and integrable. Suppose further that it is either 0 ≤ fn (x) ≤ fn+1 ≤ f (x) or
0 ≤ f (x) ≤ fn+1 ≤ fn (x) , for any x ∈ R and any n ∈ N . Then:
Z +∞ Z +∞
lim fn (x)dx = f (x)dx .
n→∞ −∞ −∞
Example 2.24. Let us consider an application of Theorem 2.21 and Corollary 2.23. Define
fn (x) = xn sin(πx) , x ∈ [0 , 1] . It is immediate to see that, for any x ∈ [0, 1] :
lim xn sin(πx) = 0 .
n→∞
Moreover, since it is 0 ≤ f (x) ≤ fn+1 ≤ fn (x) for any x ∈ [0 , 1] , the convergence is uniform
and, then: Z 1
lim xn sin(πx)dx = 0 .
n→∞ 0
The following Theorem 2.27, due to Weierstrass, establishes a sufficient condition to ensure the
uniform convergence of a series of functions.
Theorem 2.27 (Weierstrass Theorem on approximation). Let (fn ) be a sequence of functions
defined on [a , b] . Assume that for any n ∈ N , there esists Mn ∈ R such that |fn (x)| ≤ Mn
for any x ∈ [a , b] . Morevover, assume convergence for the numerical series:
∞
X
Mn .
n=1
Proof. For the Cauchy criterion of convergence (Theorem 2.9), the series of functions (2.9)
converges uniformly if and only if, for any ε > 0 , there exist nε ∈ N such that:
m
X
sup fk < ε , for any m > n > nε .
x∈[a ,b]
k=n+1
∞
X
In our case, once ε > 0 is fixed, since the numerical series Mn converges, there exists nε ∈ N
n=1
such that:
m
X
Mk < ε , for any m > n > nε .
k=n+1
Now, we use the triangle inequality:
m
X m
X m
X
sup fk ≤ sup |fk | ≤ Mk < ε .
x∈[a,b]
k=n+1
x∈[a,b]
k=n+1 k=n+1
Theorem 2.15 is useful for swapping between sum and integral of a series, and Theorem 2.19 for
term–by–term differentiability. We now state three further helpful theorems.
Theorem 2.28. If (fn ) is a sequence of continuous functions on [a , b] and if their series (2.9)
converges uniformly on [a , b] , then:
∞ Z b ∞
Z b X !
X
fn (x)dx = fn (x) dx . (2.11)
n=1 a a n=1
Proof. Define:
∞
X n
X
f (x) := fn (x) = lim fm (x) . (2.11a)
n→∞
n=1 m=1
By Theorem 2.14, function f (x) is continuous and:
Z b Xn Z b
f (x)dx = lim fm (x)dx . (2.11b)
a n→∞
m=1 a
We state (without proof) a more general result, that is not based on the uniform convergence,
but only on simple convergence and few other assumptions.
Theorem 2.29. Let (fn ) be a sequence of functions on an interval [a , b] ⊂ R . Assume that
each fn is both piecewise continuous and integrable on I , and that (2.9) converges pointwise,
on I , to a piecewise continuous function f .
Moreover, assume convergence for the numerical (positive terms) series:
X∞ Z b
|fn (x)| dx .
n=1 a
∞
X 1
Our statement follows from the convergence of the infinite series , shown in formula
n2
n=1
(2.71) of § 2.7, later on.
Moreover, if f (x) denotes the sum of the series (2.12), then, due to the uniform convergence:
π ∞
πX ∞ π ∞
1 − cos(n π)
Z Z Z
sin(n x) X 1 X
f (x)dx = dx = sin(n x)dx = .
0 0 n2 n2 0 n3
n=1 n=1 n=1
Theorem 2.31. Assume that (2.9), defined on [a , b] , converges uniformly, and assume that
each fn has continuous derivative fn0 (x) for any x ∈ [a , b] ; assume further that the series of
the derivatives is uniformly convergent. If f (x) denotes the sum of the series (2.9), then f (x) is
differentiable and, for any x ∈ [a , b] :
∞
X
f 0 (x) = fn0 (x) . (2.13)
n=1
The derivatives at the extreme points a and b are obviously understood as right and left deriva-
tives, respectively.
Proof. We present here the proof given in [24]. Let us denote by g(x) , with x ∈ [a , b] , the sum
of the series of the derivatives fn0 (x) :
∞
X
g(x) = fn0 (x) .
n=1
By Theorem 2.14, function g(x) is continuous and, by Theorem 2.15, we can integrate term by
term in [a , x] :
Z x ∞ Z
X x ∞
X ∞
X ∞
X
fn0 (ξ)dξ
g(ξ) dξ = = fn (x) − fn (a) = fn (x) − fn (a) , (2.14)
a n=1 a n=1 n=1 n=1
where linearity of the sum is used in the last step of the chain of equalities. Now, recalling
definition (2.11a) of f (x) , formula (2.14) can be rewitten as:
Z x
g(ξ) dξ = f (x) − f (a) . (2.14a)
a
26 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Differentiating both sides of (2.14a), by the Fundamental Theorem of Calculus2 we obtain g(x) =
f 0 (x) , which means that:
∞
X
0
f (x) = g(x) = fn0 (x) .
n=1
Hence, the proof is completed.
Uniform convergence of series of functions satisfies linearity properties expressed by the following
Theorem 2.32, whose proof is left as an exercise.
Moreover, if h(x) is a continuous function, defined on [a , b] , then the following series is uniformly
convergent:
X∞
h(x)fn (x) .
n=1
Given x0 ∈ R , it is important to find all the values x ∈ R such that the series of functions (2.15)
converges.
Remark 2.34. It is not restrictive, by using a translation, to consider the following simplified–
form power series, obtained from (2.15) with x0 = 0 :
∞
X
an xn . (2.16)
n=0
Obviously, the choice of x in (2.16) determinates the convergence of the series. The following
Lemma 2.35 is of some importance.
Lemma 2.35. If (2.16) converges for x = r0 then, for any 0 ≤ r < |r0 | , it is absolutely and
uniformly convergent in [−r , r] .
2
See, for example, mathworld.wolfram.com/FundamentalTheoremsofCalculus.html
2.5. POWER SERIES: RADIUS OF CONVERGENCE 27
∞
X
Proof. It is assumed the convergence of the numerical series an r0n , that is to say, there
n=0
n
r
exists a positive constant K such that an r0 ≤ K . Since < 1 , then the geometrical
r0
∞
X r n
series converges. Now, for any n ≥ 0 and any x ∈ [−r, r] :
r0
n=0
n
n n
an xn = an r0n x ≤ K x ≤ K r .
r0 r0 r0 (2.17)
By Theorem 2.27, inequality (2.17) implies that (2.16) is uniformly convergent. Due to positivity,
the convergence is also absolute.
From Lemma 2.35 it follows the fundamental Theorem 2.36, due to Cauchy and Hadamard3 ,
which explains the behaviour of a power series:
Theorem 2.36 (Cauchy–Hadamard). Given the power series (2.16), then only one of the fol-
lowing alternatives holds:
As a consequence, by Lemma 2.35, series (2.16) converges for any x ∈] − z, z[ , and, in particular,
it is convergent the series:
X∞
an y n .
n=0
To end the proof, take |y| > r and assume, by contradiction, that it is still convergent the series:
∞
X
an y n .
n=0
If so, using Lemma 2.35, it would follow that series (2.16) converges for any x ∈ ] − |y| , |y| [
and, in particular, it would converge for the number:
|y| + r
> r,
2
which contradicts the assumption r = sup C .
3
Jacques Salomon Hadamard (1865–1963), French mathematician.
28 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Definition 2.37. The interval within which (2.16) converges is called interval of convergence
and r is called radius of convergence.
Theorem 2.38 (Radius of convergence). Consider the power series (2.16) and assume that the
following limit exists:
an+1
` = lim .
n→∞ an
Then:
and apply the ratio test, that is to say, study the limit of the fraction between the (n + 1)–th
term and the n–th term in the series:
|an+1 | |x|n+1
an+1
lim = |x| lim
= |x| ` .
n→∞ |an | |x|n n→∞ an
|an+1 | |x|n+1
lim = 0 < 1.
n→∞ |an | |x|n
If ` > 0 , then:
|an+1 | |x|n+1 1
lim = |x|` < 1 ⇐⇒ |x| < .
n→∞ |an | |x|n `
Eventually, if ` = ∞ , series (2.16) does not converge when x 6= 0 , since it is:
|an+1 | |x|n+1
lim > 1,
n→∞ |an | |x|n
while, for x = 0 , series (2.16) reduces to the zero series, which converges trivially.
Example 2.39. The power series (2.18), known as geometric series, has radius of convergence
r = 1.
X∞
xn . (2.18)
n=0
which means that series (2.18) converges for −1 < x < 1. At the boundary of the interval
of convergence, namely x = 1 and x = −1 , the geometric series (2.18) does not converge. In
conclusion, the interval of convergence of (2.18) is the open interval ] − 1 , 1 [ .
2.5. POWER SERIES: RADIUS OF CONVERGENCE 29
1
Proof. Here, an = , thus:
n
an+1
lim = lim n = 1 .
n→∞ an n→∞ n + 1
Example 2.41. Series (2.20), given below, has infinite radius of convergence:
∞
X xn
. (2.20)
n!
n=0
1
Proof. Since it is an = for any n ∈ N , it follows that:
n!
an+1 1
lim
= lim = 0.
n→∞ an n→∞ n + 1
It is possible to differentiate and integrate power series, as stated in the following Theorem 2.42,
which we include for completeness, as it represents a particular case of Theorems 2.28 and 2.31.
Theorem 2.42. Let f (x) be the sum of the power series (2.16), with radius of convergence r .
The following results hold.
(i) f (x) is differentiable and, for any |x| < r , it is:
∞
X
0
f (x) = n an xn−1 ;
n=1
The radius of convergence of both power series f 0 (x) and F (x) is that of f (x) .
Power series behave nicely with respect to the usual arithmetical operations, as shown in The-
orem 2.43, which states some useful result.
30 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Theorem 2.43. Consider two power series, with radii of convergence r1 and r2 respectively:
∞
X ∞
X
n
f1 (x) = an x , f2 (x) = bn xn . (2.21)
n=0 n=0
We state, without proof, Theorem 2.44, concerning the product of two power series.
Theorem 2.44. Consider the two power series in (2.21), with radii of convergence r1 and r2
respectively. The product of the two power series is defined by the Cauchy formula:
∞
X n
X
n
cn x , where cn = aj bn−j , (2.22)
n=0 j=0
that is:
c0 = a0 b0 ,
c1 = a0 b1 + a1 b0 ,
..
.
cn = a0 bn + a1 bn−1 + · · · + an−1 b1 + an b0 .
Series (2.22) has interval of convergence given by |x| < r = min{r1 , r2 } , and its sum is the
pointwise product f1 (x) f2 (x) .
Since f has derivatives of any order, we may form the limit of (2.23) as n → ∞ ; a condition is
stated in Theorem 2.45 to detect when the passage to the limit is effective.
Theorem 2.45. If f has derivatives of any order in the open interval I , with x0 , x ∈ I , and if:
f (n+1) (ξ)
lim Rn (f (x), x0 ) = lim (x − x0 )n+1 = 0 ,
n→∞ n→∞ (n + 1)!
then:
∞
X f (n) (x0 )
f (x) = y (x − x0 )n . (2.24)
n!
n=0
4
Brook Taylor (1685–1731), English mathematician.
5
Giuseppe Luigi Lagrange (1736–1813), Italian mathematician.
2.6. TAYLOR–MACLAURIN SERIES 31
Remark 2.47. Assuming the existence of the derivatives of any order is not enough to infer
that a function is analytic and, thus, it can be represented with a convergent power series. For
instance, the function:
( 1
e − x2 if x 6= 0
f (x) =
0 if x = 0
has derivatives of any order in x0 = 0 , but such derivates are all zero, therefore the Taylor series
reduces to the zero function. This happens because the Lagrange remainder does not vanish as
n → ∞.
Note that the majority of the functions, that interest us, does not possess the behaviour shown
in Remark 2.47. The series expansion of the most important, commonly used, functions can
be inferred from Equation (2.23), i.e., from the Taylor–Lagrange Theorem. And Theorem 2.45
yields a sufficient condition to ensure that a given function is analytic.
Corollary 2.48. Consider f with derivatives of any order in the interval I = ] a , b [ . Assume
that there exist L , M > 0 such that, for any n ∈ N ∪ {0} and for any x ∈ I :
(n)
f (x) ≤ M Ln . (2.25)
Then, for any x0 ∈ I , function f (x) coincides with its Taylor series in I .
Proof. Assume x > x0 . The Lagrange remainder for f (x) is given by:
f (n+1) (ξ)
Rn (f (x) , x0 ) = (x − x0 )n+1 ,
(n + 1)!
where ξ ∈ (x0 , x) , which can be written as ξ = x0 + α(x − x0 ) , with 0 < α < 1 . Now, using
condition (2.25), it follows:
n+1
L (b − a)
|Rn (f (x) , x0 )| ≤ M.
(n + 1)!
Corollary 2.48, together with Theorem 2.42, allows to find the power series expansion for the
most common elementary functions. Theorem 2.49 concerns a first group of power series that
converges for any x ∈ R .
Theorem 2.49. For any x ∈ R , the exponential power series expansion holds:
∞
X xn
ex = . (2.26)
n!
n=0
∞ ∞
X (−1)n x2n+1 X (−1)n x2n
sin x = , (2.27) cos x = , (2.28)
(2n + 1)! (2n)!
n=0 n=0
∞ ∞
X x2n+1 X x2n
sinh x = , (2.29) cosh x = . (2.30)
(2n + 1)! (2n)!
n=0 n=0
Proof. First, observe that the general term in each of the five series (2.26)–(2.30) comes form
the MacLaurin6 formula.
To show that f (x) = ex is the sum of the series (2.26), let us use the fact that f (n) (x) = ex for
any n ∈ N ; in this way, it is possible to infer that, in any interval [a , b] , the disequality (2.25)
is fulfilled if we take M = 1 and L = max{ex : x ∈ [a, b]} .
To prove (2.27) and (2.28), in which derivatives of the goniometric functions sin x and cos x are
considered, condition (2.25) is immediately verified by taking M = 1 and L = 1 .
Finally, (2.29) and (2.30) are a straightforward consequence of the definition of the hyperbolic
functions in terms of the exponential:
ex + e−x ex − e−x
cosh x = , sinh x = ,
2 2
together with Theorem 2.43.
Theorem 2.50 concerns a second group of power series converging for |x| < 1 .
Theorem 2.50. If |x| < 1 , the following power series expansions hold:
∞ ∞
1 X 1 X
= xn , (2.31) = (n + 1)xn , (2.32)
1−x (1 − x)2
n=0 n=0
∞ ∞
X xn+1 X xn+1
ln(1 − x) = − , (2.33) ln(1 + x) = (−1)n , (2.34)
n+1 n+1
n=0 n=0
∞ ∞
x2n+1 x2n+1
X 1+x X
arctan x = (−1)n , (2.35) ln =2 . (2.36)
2n + 1 1−x 2n + 1
n=0 n=0
1
Proof. To prove (2.31), define f (x) = and build the MacLaurin polynomial of order n ,
1−x
which is Pn ( f (x) , 0 ) = 1 + x + x2 + . . . + xn ; the remainder can be thus estimated directly:
Rn f (x) , n = f (x) − Pn f (x) , 0
1
= − (1 + x + x2 + · · · + xn ) (2.37)
1−x
1 1−x n+1 x n+1
= − = .
1−x 1−x 1−x
Assuming |x| < 1 , we see that the remainder vanishes for n → ∞ , thus (2.31) follows.
Indentity (2.32) can be proven by employing both formula (2.31) and Theorem 2.44, with an =
bn = 1 .
To obtain (2.33), the geometric series in (2.31) can be integrated term by term, using Theorem
2.42; in fact, letting |x| < 1 , we can consider the integral:
Z x
dt
= − ln(1 − x) .
0 1−t
6
Colin Maclaurin (1698–1746), Scottish mathematician.
2.6. TAYLOR–MACLAURIN SERIES 33
Formula (2.33) is then a consequence of formula (2.31). Formula (2.34) can be proven analogously
to (2.33), by considering −x instead of x .
To prove (2.35), we use again formula (2.31) with t = −x , so that we have:
∞
1 X
= (−1)n x2n .
1 + x2
n=0
Integrating, taking |x| < 1 , and using Theorem 2.42, the following result is obtained:
x x ∞
1 1 + x X x2n+1
Z Z
dt 1 1 1
= + dt = ln = .
0 1 − t2 2 0 1+t 1−t 2 1−x 2n + 1
n=0
Observe that the left hand side of (2.40) does not require the numerator to be a natural number.
Therefore, if α ∈ R and if n ∈ N , the generalized binomial coefficient is defined as:
α α · (α − 1) · · · · · (α − n + 1)
= . (2.41)
n n!
From (2.41) an useful property of the generalized binomial coefficient can be inferred, and later
used to expand in power series the function f (x) = (1 + x)α .
Proposition 2.51. For any α ∈ R and any n ∈ N , the following identity holds:
α α α
n + (n + 1) =α . (2.42)
n n+1 n
34 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
By using Proposition 2.51, it is possible to prove the so–called generalised Binomial Theorem
2.52.
Theorem 2.52 (Generalised Binomial). For any α ∈ R and |x| < 1 , the following identity
holds:
∞
α
X α n
(1 + x) = x . (2.43)
n
n=0
Proof. Let us denote f (x) the sum of the generalised binomial series:
∞
X α n
f (x) = x ,
n
n=0
f (x)
g(x) = .
(1 + x)α
To prove the thesis, let us show that g(x) = 1 for any |x| < 1 . Differentiating g(x) we obtain:
Thus, g 0 (x) = 0 for any |x| < 1 , which implies that g(x) is a constant function. It follows that
g(x) = g(0) = f (0) = 1 , which proves thesis (2.43).
1
When considering the power series expansion of arcsin, the particular value α = − turns out
2
to be important. Let us, then, study the generalised binomial coefficient (2.41) corresponding to
such an α .
2.6. TAYLOR–MACLAURIN SERIES 35
∞
X (2 n − 1)!!
1
√ = xn , (2.46)
1 − x n=0 (2 n)!!
∞
1 X (2 n − 1)!! n
√ = (−1)n x . (2.47)
1 + x n=0 (2n)!!
Formula (2.46) is implied by Theorem 2.42 and yields the MacLaurin series for arcsin x , while
(2.47) gives the series for arcsinh x , as expressed in the following Theorem 2.55.
Theorem 2.55. Considering |x| < 1 and letting (−1)!! = 1 , then:
∞
X (2 n − 1)!! x2 n+1
arcsin x = , (2.48)
(2 n)!! 2 n + 1
n=0
∞
X (2 n − 1)!! x2 n+1
arcsinh x = (−1)n . (2.49)
(2 n)!! 2 n + 1
n=0
Using the power series (2.46) and the Central Binomial Coefficient formula:
2n (2n − 1)!!
2n
= (2.50)
n n!
provable by induction, a result can be obtained, due to Lehemer7 [26].
1
Theorem 2.56 (Lehmer). If |x| < , then:
4
∞
1 X 2n n
√ = x . (2.51)
1 − 4x n=0 n
Proof. Formula (2.50) yields:
∞ ∞ ∞
2n n X 2n (2n − 1)!! n X 4n (2n − 1)!! n
X
x = x = x .
n n! 2n n!
n=0 n=0 n=0
Notice that from Theorem 2.38 it follows that the radius of convergence of the power series
(2.55) is infinite.
7
Derrick Henry Lehmer (1905–1991), American mathematician.
2.6. TAYLOR–MACLAURIN SERIES 37
Theorem 2.57 (Abel). Denote by f (x) the sum of the power series (2.16), in which we assume
X∞
that the radius of convergenge is r > 0 . Assume further that the numerical series rn an
n=0
converges. Then:
∞
X
lim f (x) = an r n . (2.56)
x→r−
n=0
Proof. The generality of the proof is not affected by the choice r = 1 , as different radii can be
achieved with a straightforward change of variable. Let:
n−1
X
sn = am ;
m=0
then:
∞
X
s = lim sn = an .
n→∞
n=0
Now, observe that a0 = s1 and an = sn+1 − sn for any n ∈ N . If |x| < 1 , then 1 is the
radius of convergence of the power series:
∞
X
sn+1 xn . (2.57)
n=0
∞
X ∞
X ∞
X
n n
(1 − x) sn+1 x = sn+1 x − sn+1 xn+1
n=0 n=0 n=0
X∞ X∞
= sn+1 xn − sn xn (2.58)
n=0 n=1
∞
X ∞
X
n
= s1 + (sn+1 − sn )x = a0 + an xn = f (x) .
n=1 n=1
To obtain thesis (2.56) we have to show that, for any ε > 0 , there exists δε > 0 such that
|f (x) − s| < ε , for any x such that 1 − δε < x < 1 . From (2.58) and using formula (2.31) for the
8
Niels Henrik Abel (1802–1829), Norvegian mathematician.
38 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
ε
Now, fixed ε > 0 , there exists nε ∈ N such that |sn+1 − s| < for any n ∈ N , n > nε ;
2
therefore, using the triangle inequality in (2.59), the following holds for x ∈] − 1, 1[ :
∞
n
X ε X
|f (x) − s| = (1 − x) (sn+1 − s)xn + (sn+1 − s)xn
n=0 n=nε +1
∞
n
X ε X
≤ (1 − x) (sn+1 − s)xn + (1 − x) (sn+1 − s)xn
n=0 n=nε +1
nε ∞
(2.60)
X ε X
≤ (1 − x) |sn+1 − s| |x|n + (1 − x) |x|n
2
n=0 n=nε +1
nε nε
X ε X ε
≤ (1 − x) |sn+1 − s| |x|n + ≤ (1 − x) |sn+1 − s| + .
2 2
n=0 n=0
is continuous and vanishes for x = 1 , it is possible to choose δ ∈]0 , 1[ such that, if 1−δ < x < 1 ,
we have:
nε
X ε
(1 − x) |sn+1 − s| < .
2
n=0
Theorem 2.57 allows to compute, roundly, the sum of many interesting series.
Example 2.58. Recalling the power series expansion (2.33), from Theorem 2.57, with x = 1 ,
it follows:
∞
X (−1)n+1
ln 2 = .
n
n=1
Example 2.59. Recalling the power series expansion (2.35), Theorem 2.57, with x = 1 , allows
finding the sum of the Leibnitz–Gregory9 series:
∞
π X 1
= (−1)n .
4 2n + 1
n=0
9
James Gregory (1638–1675), Scottish mathematician and astronomer.
Gottfried Wilhelm von Leibnitz (1646–1716), German mathematician and philosopher.
2.7. BASEL PROBLEM 39
Example 2.60. Recalling the particular binomial expansion (2.46), Abel Theorem 2.57 implies
that, for x = 1 , the following holds:
∞
1 X (2n − 1)!!
√ = (−1)n .
2 n=0 (2n)!!
π
Using the fact that arccos x = − arcsin x , it is possible to obtain a second series, which gives
2
π.
Example 2.61. Recalling the arcsin expansion (2.48), from Theorem 2.57 it follows:
∞
π X (2n − 1)!! 1
= .
2 (2n)!! 2n + 1
n=0
Example 2.62. We show here two summation formulæ connecting π to the central binomial
coefficients:
2n
∞
X n π
= ; (2.61)
4n (2 n + 1) 2
n=0
2n
∞
X n π
n
= . (2.62)
16 (2 n + 1) 3
n=0
The key to show (2.61) and (2.62) lies in the representation of the central binomial coefficient
(2.50), whose insertion in the left hand side of (2.61) leads to the infinite series:
∞
X (2n + 1)!!
. (2.63)
2n n! (2n + 1)
n=0
We further notice that, from the power expansion of the arcsin function (2.48), it is possible to
infer the following equality:
√ ∞
arcsin x X (2n + 1)!!
√ = xn . (2.64)
x 2n n! (2n + 1)
n=0
The radius of convergence of the power series (2.64) is 1 ; Abel Theorem 2.57 can thus be applied
to arrive to (2.61). It is worth noting that (2.61) can also be obtained using the Lehemer series
(2.51), via the change of variable y = 4 x and integrating term by term.
A similar argument leads to (2.62); here, the starting point is the following power series expan-
sion, which has, again, radius of convergence r = 1 :
2n
∞
arcsin x X n
= n
x2n (2.65)
x 4 (2n + 1)
n=0
1
Equality (2.62) follows by evaluating formula (2.65) at x = .
2
Mengoli10 originally posed this problem in 1644, that takes its name from Basel, birthplace of
π2
Euler11 who first provided the correct solution in [12].
6
There exist several solutions of the Basel problem; here we present the solution of Choe [7],
based on the power series expansion of f (x) = arcsin x , shown in Formula (2.48), as well as on
the Abel Theorem 2.57 and on the following integral Formula (2.67), which can be proved by
induction on m ∈ N :
Z π
2 (2 m)!!
sin2 m+1 t dt = . (2.67)
0 (2 m + 1)!!
The first step towards solving the Basel problem is to observe that, in the sum (2.66), the
attention can be confined to odd indexes only. Namely, if E denotes the sum of the series (2.66),
then E can be computed by considering, separately, the sums on even and odd indexes:
∞ ∞
X 1 X 1
2
+ =E.
(2 n) (2 n + 1)2
n=1 n=0
yielding:
∞
X 1 3
2
= E. (2.68)
(2n + 1) 4
n=0
π2 3 π2
Now, observe that E = ⇐⇒ E= . In other words, the Basel problem is equivalent
6 4 8
to show that:
∞
X 1 π2
= , (2.69)
(2n + 1)2 8
n=0
in which an is the n−th series term, and proving that ρ > 1 . In the case of (2.48), with x = 1 :
This implies, also, that the series (2.48) converges uniformly. The change of variable x = sin t in
π π
both sides of (2.48) yields, when − < t < :
2 2
∞
X (2n − 1)!! sin2n+1 t
t = sin t + . (2.70)
(2n)!! 2n + 1
n=1
10
Pietro Mengoli (1626–1686), Italian mathematician and clergyman from Bologna.
11
Leonhard Euler (1707–1783), Swiss mathematician and physicist.
12
Joseph Ludwig Raabe (1801–1859), Swiss mathematician.
2.8. EXTENSION OF ELEMENTARY FUNCTIONS TO THE COMPLEX FIELD 41
π
Integrating (2.70) term by term, on the interval [0 , ] , and using (2.67), we obtain:
2
∞ Z π
π2 X (2n − 1)!! 2 sin2n+1 t
=1+ dt
8 (2n)!! 0 2n + 1
n=1
∞
X (2n − 1)!! (2n)!! 1
=1+
(2n)!! (2n + 1)!! 2n + 1
n=1
∞ ∞
X 1 X 1
=1+ 2
= .
(2n + 1) (2n + 1)2
n=1 n=0
Definition 2.63.
∞
X zn
ez := . (2.72)
n!
n=0
Equations (2.26) and (2.72) only differ in the fact that, in the latest one, the argument can be
a complex number. Almost all the familiar properties of the exponential still hold, with the one
exception of positivity, which has no sense in the unordered field C . The fundamental property
of the complex exponential is stated in the following Theorem 2.64, due to Euler.
ez = ex+i y
=e ·e
x iy
The last step, above, exploits the real power series expansion for the sine and cosine functions
given in (2.27) and (2.28) respectively.
42 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
The first beautiful consequence of Theorem 2.64 is the famous Euler identity.
Corollary 2.65 (Euler identity).
ei π + 1 = 0 . (2.74)
Proof. First observe that, if x = 0 in (2.73), then it holds, for any y ∈ R :
ei y = cos y + i sin y . (2.75)
Now, with y = π , equation (2.75) follows.
The C–extension of the exponential has an important consequence since the exponential func-
tion, when considered as a function C → C , is no longer one–to–one, but it is a periodic function.
In fact, if z , w ∈ C , then:
ez = ew ⇐⇒ z = w + 2 n π i with n ∈ Z .
Among the infinite logarithms of a complex number, we pin down one, corresponding to the
most convenient argument.
In other words, for a non–zero complex w , the principal determination (or principal value)
Log w is the logarithm whose imaginary part lies in the interval (−π , π] .
Definition 2.72. Given z ∈ C , z 6= 0 , and w ∈ C , the complex power function is defined as:
z w = ew Log z
π
Example 2.73. Compute i i . Applying Definition 2.72: i i = ei Log i . Since arg(i) = and
2
π π π
|i| = 1 , then Log i = i . Finally, i i = ei i 2 = e− 2 .
2
Example 2.74. In C , it is possible solve equations like sin z = 2 , obviously finding complex
solutions. From the sin definition (2.77), in fact, we obtain:
e2 i z − 4 i ei z − 1 = 0 .
Thus: √
ei z = 2 ± 3 i .
2.9 Exercises
2.9.1 Solved exercises
1. Given the following sequence of functions, establish whether it is pointwise and/or uniformly
convergent:
n x + x2
fn (x) = , x ∈ [0, 1] ,
n2
2. Exaluate the pointwise limit of the sequence of functions:
√
fn (x) = n 1 + xn , x ≥ 0.
3. Show that the following sequence of functions converges pointwise, but not uniformly, to
f (x) = 0 :
fn (x) = n x e−n x , x > 0,
5. Show that:
∞
X 1
= ln 2 .
n 2n
n=1
6. Evaluate:
∞
n e−n x
Z
lim dx .
n→∞ 1 1 + nx
1
x(1 − x)
Z
7. Use the definite integral dx to prove that:
0 1+x
∞
X (−1)n 3
= − ln 4 .
(n + 1)(n + 2) 2
n=1
−n
x2
8. Let fn (x) = 1+ , x ≥ 0.
n
x x2 x2
x
lim fn (x) = lim + = lim + lim 2 = 0 + 0 = 0 .
n→∞ n→∞ n n2 n→∞ n n→∞ n
nx + x2
n+1
sup |fn (x) − 0| = sup 2
= .
x∈[0,1] x∈[0,1] n n2
Observe that:
n+1
lim sup |fn (x) − 0| = lim = 0.
n→∞ x∈[0,1] n→∞ n2
lim fn (x) = 1 .
n→∞
13
See, for example,mathworld.wolfram.com/SqueezingTheorem.html
2.9. EXERCISES 45
1 1
Recalling that, here, < 1 , x 6= 0 , we consider a change of variable t = and repeat the
x x √
previous argument (that we followed in the case of a variable t < 1) to obtain lim n tn + 1 =
n→∞
1 , that is: r
n 1
lim + 1 = 1.
n→∞ xn
In other words, for x > 1 , we have shown that:
√
lim fn (x) = lim n 1 + xn = x .
n→∞ n→∞
3. The pointwise limit of sequence fn (x) = n x e−n x is f (x) = 0 , due to the exponential decay
of the factor e−n x . To investigate the possible uniform convergence, we consider:
Differentiating we find:
d
n x e−n x = n e−n x (1 − n x) ,
dx
1
showing that x = n is a local maximizer and the corresponding extremum is:
1 1
fn = .
n e
But this implies that the found convergence cannot be uniform, since:
1
lim sup |fn (x) − f (x)| = 6= 0 .
n→∞ x>0 e
√ √
4. For any x ∈ [−1, 1] and any n ∈ N , it holds that 1 − xn ≤ 2 , thus:
√ √
1 − xn 2
fn (x) = ≤ 2 . (2.79)
n2 n
Now, observe that inequality (2.79) is independent of x ∈ [−1, 1] : this fact, taking the
supremum with respect to x ∈ [−1, 1] , ensures uniform convergence.
n
= xn−1 dx .
n2 0
n=1 n=1
46 CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Since the geometric series, in the right–hand side above, converges uniformly, we can swap
series and integral, obtaining:
∞ Z 1 X ∞
!
X 1 2
= xn−1 dx .
n 2n 0
n=1 n=1
Therefore
∞ Z 1
X 1 2 1
n
= dx .
n2
n=1 0 1−x
n
6. Define the (decreasing) function hn (x) = , with x ∈ [1, +∞) . Then:
1 + nx
n2
h0n (x) = − < 0.
(1 + nx)2
Since:
n
lim =0
x→∞ 1 + n x
and
n n
sup = hn (1) = ,
x∈[1,∞) 1 + n x 1 + n
we can infer that:
n
|hn (x)| ≤ < 1.
1+n
Therefore:
n e−n x
−n x
1 + n x < e .
This shows uniform convergence for fn . We can now invoke Theorem 2.15, to obtain:
Z ∞ Z ∞ Z ∞
n e−n x n e−n x
lim dx = lim dx = 0 dx = 0 .
n→∞ 1 1 + nx 1 n→∞ 1 + n x 1
8. Observe that:
x2
2 2
x x
2 −n
x
−n ln 1 + −n +o
1+ =e n =e n n 2
= e−x (1+o(1))
n
where we have used ln(1 + t) = t + o(t) when t ' 0 . This means that:
−n
x2
2
lim 1+ = e−x .
n→∞ n
We have thus shown point (a). The second statement follows from Corollary 2.23.
Hint. Consider the sequence gn (x) = ln fn (x) and use the power series (2.33) and (2.28).
x + x2 e n x
3. Establish if the sequence of functions (fn )n∈N , defined, for x ∈ R , by fn (x) = ,
1 + enx
converges pointwise and/or uniformly.
∞
X 1 3
4. Show that n
= ln .
n3 2
n=1
Z 1 x+1
5. Show that lim e n dx = 1 .
n→∞ 0
6. Consider the following equality and say if (and why) it is true or false:
Z 1 Z 1
x4 x4
lim dx = lim dx .
n→∞ 0 x2 + n2 2
0 n→∞ x + n
2
n(x3 + x) e−x
7. Let fn (x) = , with x ∈ [0 , 1] .
1 + nx
c. Show that, for any a > 0 , sequence (fn ) converges uniformly to f on [a , 1] , but the
convergence is not uniform on [0 , 1] .
Z 1
d. Evaluate lim fn (x) dx .
n→∞ 0
Hint.
1 1−x 1
2
= 2
+ .
(1 + x)(1 + x ) 2 (1 + x ) 2(1 + x)
∞ ∞
x5
Z X 1
10. Show that x 2 dx = 3
.
0 e −1 n=1
n
√
11. Show that cos z = 2 ⇐⇒ z = 2 n π − i ln 2 ± 3 , n ∈ Z.
3 Multidimensional differential cal-
culus
by
f (x1 , . . . , xj−1 , · , xj+1 , . . . , xn ) .
If g is differentiable at some t0 ∈ (a , b) , then the first–order partial derivative of f at
(xl , . . . , xj−1 , t0 , xj+1 , . . . , xn ) , with respect to xj , is defined by:
∂f
fxj (x1 , . . . , xj−1 , t0 , xj+1 , . . . , xn ) := (x1 , . . . , xj−1 , t0 , xj+1 , . . . , xn )
∂xj
:=g 0 (t0 ) .
Therefore, the partial derivative fxj exists at a point a if and only if the following limit exists:
∂f f (a + h ej ) − f (a)
(a) := lim .
∂xj h→0 h
Higher–order partial derivatives are defined by iteration. For example, when it exists, the second–
order partial derivative of f , with respect to xj and xk , is defined by:
∂2f
∂ ∂f
fxj xk := := .
∂xk ∂xj ∂xk ∂xj
Second–order partial derivatives are called mixed when j 6= k .
Definition 3.1. Let V be a non–empty open subset of Rn , let f : V → R and p ∈ N.
(i) f is said to be C p on V if and only if every k–th order partial derivative of f , with k ≤ p ,
exists and is continuous on V .
49
50 CHAPTER 3. MULTIDIMENSIONAL DIFFERENTIAL CALCULUS
∂f ∂g ∂f
(f g) = f +g .
∂x ∂x ∂x
Example 3.3. By the Mean–Value Theorem2 , if f ( · , y) is continuous on [a, b] and the partial
derivative fx ( · , y) exists on (a , b) , then there exists a point c ∈ (a , b) (which may depend on y
as well as on a and b) such that:
∂f
f (b , y) − f (a , y) = (b − a) (c , y) .
∂x
In most situations, when dealing with higher–order partial derivatives, the order of computa-
tion of the derivatives is, in some sense, arbitrary. This is expressed by the Clairaut3 –Schwarz
Theorem.
∂2 f ∂2 f
(a , b) = (a , b) .
∂y ∂x ∂x ∂y
∂2 f ∂2 f
(a) = (a) .
∂xj ∂xk ∂xk ∂xj
Remark 3.6. Existence of partial derivatives does not ensure continuity. As an example, con-
sider: ( xy
if (x , y) 6= (0, 0) ,
f (x , y) = x2 + y 2
0 if (x , y) = (0 , 0) .
This function is not continuous at (0 , 0) , but admits partial derivatives at any (x, y) ∈ R2 ,
since:
f (∆x , 0) − f (0 , 0)
lim = lim 0 = 0 ,
∆x→0 ∆x ∆x→0
and
f (0 , ∆y) − f (0 , 0)
lim = lim 0 = 0 .
∆y→0 ∆y ∆y→0
3.2 Differentiability
In this section, we define what it means for a vector function f to be differentiable at a point
a . Whatever our definition, if f is differentiable at a , then we expect two things:
f (x + h) − f (x)
lim := f 0 (x) .
h→0 h
The definition of differentiability for functions of several variables extends Property (3.1).
f (a + h) − f (a) − d · h
lim =0
h→0 ||h||
(i) f is continuous at a ;
Theorem 3.9. Let V be open in Rn , let a ∈ V and suppose that f : V → R . If all first–order
partial derivatives of f exist in V and are continuous at a , then f is differentiable at a .
The composition of functions follows similar rules, for differentiation, as in the one–dimensional
case. For instance, the Chain Rule 4 holds in the following way. Consider a vector function
g : I → Rn , g = (g1 , . . . , gn ) , defined on an open interval I ⊆ R , and consider f : g(I) ⊆
Rn → R . If each of the components gj of g is differentiable at t0 ∈ I , and if f is differentiable
at a = (g1 (t0 ) , . . . , gn (t0 )) , then the composition ϕ(t) := f (g(t)) is differentiable at t0 , and
we have:
ϕ0 (t0 ) = ∇f (a) · g 0 (t0 ) ,
where · is the dot (inner) product in Rn and:
In order to extend the notion of gradient, we introduce the Jacobian5 matrix associated to a
vector–valued function.
Definition 3.11. Let f : Rn → Rm be a function from the Euclidean n–space to the Euclidean
m–space. f has m real–valued component functions:
f1 (x1 , . . . , xn ) , . . . , fm (x1 , . . . , xn ) .
If the partial derivatives of the component functions exist, they can be organized in an m–by–n
matrix, namely the Jacobian matrix J of f :
x · ∇f (x) = k f (x) .
(i) f (a) is called a local minimum of f if and only if there exists r > 0 such that f (a) ≤ f (x)
for all x ∈ Br (a) , an open ball neighborhood of a ;
(ii) f (a) is called a local maximum of f if and only if there exists r > 0 such that f (a) ≥ f (x)
for all x ∈ Br (a) ;
4
See, for example, mathworld.wolfram.com/ChainRule.html
5
Carl Gustav Jacob Jacobi (1804–1851), German mathematician.
3.4. SUFFICIENT CONDITIONS 53
(iii) f (a) is called a local extremum of f if and only if f (a) is a local maximum or a local
minimum of f .
Remark 3.15. If the first–order partial derivatives of f exist at a, and if f (a) is a local
extremum of f , then ∇f (a) = 0 .
In fact, the one–dimensional function:
As in the one–dimensional case, condition ∇f (a) = 0 is necessary but not sufficient for f (a) to
be a local extremum.
Example 3.16. There exist continuously differentiable functions satisfying ∇f (a) = 0 and
such that f (a) is neither a local maximum nor a local minimum.
Consider, for instance, in the case n = 2 , the following function:
f (x, y) = y 2 − x2 .
It is easy to check that ∇f (0) = 0 , but the origin is a saddle point, as shown in Figure 3.1.
Definition 3.18. Let V ⊆ Rn an open set and let f : V → R be a C 2 function. The Hessian
matrix of f at x ∈ V (or, simply, the Hessian) is the symmetric square matrix formed by the
second–order partial derivatives of f , evaluated at point x :
2
∂ f
H(f )(x) := (x) , for i, j = 1, . . . , n .
∂xi ∂xj
Tests for extrema and saddle points, in the simplest situation of n = 2 , are stated in Theorem
3.19.
Example 3.20. A couple of examples are provided here, and the Reader is invited to verify the
stated results.
Tests for extrema and saddle points, in the general situation of n variables, are stated in Theorem
3.21.
Theorem 3.21. In n variables, a critical point x0 :
Theorem 3.22 (Lagrange multipliers – general case). Let m < n , let V be open in Rn , and
let f , gj : V → R be C 1 on V , for j = 1 , 2 . . . , m . Suppose that:
∂(g1 , . . . , gm )
∂(x1 , . . . , xn )
M = {x ∈ V : gj (x) = 0} .
We will limit the proof of the Lagrange multipliers Theorem 3.22 in a two–dimensional context.
To this aim, it is first necessary to consider some preliminary results; we will resume the proof
in §3.8.
[x , y] := {z ∈ Rn : z = t x + (1 − t) y , 0 ≤ t ≤ 1} .
The one–dimensional Mean–Value theorem (already met in Example 3.3), also called Lagrange
Mean–Value theorem or First Mean–Value theorem, can be extended to the Euclidean space
Rn .
Theorem 3.25 (Implicit Function – case n = 2). Let Ω be an open set in R2 , and let
f : Ω → R be a C 1 function. Suppose there exists (x0 , y0 ) ∈ Ω such that f (x0 , y0 ) = 0 and
fy (x0 , y0 ) 6= 0 .
Then, there exist δ , ε > 0 such that, for any x ∈ (x0 − δ , x0 + δ) there exists a unique
y = ϕ(x) ∈ (y0 − ε , y0 + ε) such that:
f (x, y) = 0 .
Moreover, function y = ϕ(x) is C 1 in (x0 −δ, x0 +δ) and it holds that, for any x ∈ (x0 −δ , x0 +δ) :
fx (x, ϕ(x))
ϕ0 (x) = − .
fy (x, ϕ(x))
Proof. Let us assume that f (x0 , y0 ) > 0 . Since function fy (x , y) is continuos, it is possibile to
find a ball Bδ1 (x0 , y0 ) in which it is verified that (x , y) ∈ Bδ1 (x0 , y0 ) =⇒ fy (x, y) > 0 .
This means that, with an appropriate narrowing of parameters ε and δ , function y 7→ f (x , y)
can be assumed to be an increasing function, for any x ∈ (x0 − δ , x0 + δ) .
In particular, y 7→ f (x0 , y) is increasing and, since f (x0 , y0 ) = 0 by assumption, the following
disequalities are verified, for ε small enough:
Using, again, continuity of f and an appropriate narrowing of δ , we infer that, for any x ∈
(x0 − δ , x0 + δ) :
f (x, y0 + ε) > 0 and f (x, y0 − ε) < 0 .
In conclusion, using continuity of y 7→ f (x , y) and the Bolzano theorem7 on the existence of
zeros, we have shown that, for any x ∈ (x0 −δ, x0 +δ) , there is a unique y = ϕ(x) ∈ (y0 −ε, y0 +ε)
such that:
f (x , y) = f (x , ϕ(x)) = 0 .
To prove the second part of Theorem 3.25, we need to show that ϕ(x) is differentiable. To this
aim, consider h ∈ R such that x + h ∈ (x0 − δ, x0 + δ) . In this way, from the Mean–Value
Theorem 3.24, there exist θ ∈ (0, 1) such that:
0 = f x + h , ϕ(x + h) − f x , ϕ(x)
= fx x + θ h ϕ(x) + θ ϕ(x + h) − ϕ(x) h +
+ fy x + θ h ϕ(x) + θ ϕ(x + h) − ϕ(x) ϕ(x + h) − ϕ(x) ,
thus:
ϕ(x + h) − ϕ(x) fx x + θ h ϕ(x) + θ ϕ(x + h) − ϕ(x)
=− .
h fy x + θ h ϕ(x) + θ ϕ(x + h) − ϕ(x)
The thesis follows by taking, in the equality above, the limit for h → 0 , observing that h →
0 =⇒ θ → 0 , and recalling that f (x , y) is C 1 .
We are now ready to state the Implicit function Theorem 3.26 in the general n–dimensional
case; here, Ω is an open set in Rn × R , thus (x, y) ∈ Ω means that x ∈ Rn and y ∈ R .
Theorem 3.26 (Implicit Function – general case). Let Ω ⊆ Rn × R be open, and let f ∈
C 1 (Ω , R) . Assume that there exists (x0 , y0 ) ∈ Ω such that f (x0 , y0 ) = 0 and fy (x0 , y0 ) 6= 0 .
Then, there exist an open ball Bδ (x0 ) , an open interval (y0 − ε , y0 + ε) and a function ϕ :
(y0 − ε , y0 + ε) → R , such that:
7
Bernard Placidus Johann Nepomuk Bolzano (1781–1848), Czech mathematician, theologian and philosopher.
For the theorem of Bolzano see, for example, mathworld.wolfram.com/BolzanoTheorem.html
3.8. PROOF OF THEOREM 3.22 57
f (x, y) = 0 ⇐⇒ y = ϕ(x) ;
Theorem 3.27 (Lagrange multipliers – case n = 2). Let A ⊂ R2 be open, and let f , g : A → R
be C 1 functions. Consider the subset of A :
M = {(x , y) ∈ A : g(x , y) = 0} .
Assume that ∇g(x , y) 6= 0 for any (x , y) ∈ M . Assume further that (x0 , y0 ) ∈ M is a maximum
or a minimum of f (x , y) for any (x , y) ∈ M .
Then, there exists λ ∈ R such that:
∇f (x0 , y0 ) = λ∇g(x0 , y0 ) .
Proof. Since ∇g(x0 , y0 ) 6= 0 , we can assume that gy (x0 , y0 ) 6= 0, . Thus, from the Implicit
function Theorem 3.25, there exist ε , δ > 0 such that, for x ∈ (x0 −δ , x0 +δ) , y ∈ (y0 −ε , y0 +ε) ,
it holds:
g(x , y) = g(x , ϕ(x)) = 0 .
Consider the function x 7→ f x , ϕ(x) := h(x) , for x ∈ (x0 − δ , x0 + δ) . By assumption, h(x)
admits an extremum in x = x0 , therefore its derivative in x0 vanishes. Using the Chain Rule,
it follows:
0 = h0 (x0 ) = fx x0 , ϕ(x0 ) + fy x0 , ϕ(x0 ) ϕ0 (x0 ) .
(3.2)
Again, use the Implicit function Theorem 3.25, which gives:
0 gx x0 , ϕ(x0 )
ϕ (x0 ) = − .
gy x0 , ϕ(x0 )
we evaluate:
00 00
Lxx Lxy gx
00 00
Λ = det Lxy Lyy gy .
gx gy 0
Then:
max f (x , y) = xa y 1−a
(3.3)
subject to p x + q y − c = 0
Eliminating m from the first two equations, by subtraction, we obtain the two–by-two linear
system in the variables x , y :
(
(1 − a) p x − a q y = 0 ,
px + qy − c = 0 .
Solving the 2 × 2 system and recovering m , from m = (a xa−1 y 1−a ) p−1 , we find the critical
point:
ac
x=
p
c (1 − a)
y=
q
m = (1 − a)1−a aa p−a q a−1
3.9. SUFFICIENT CONDITIONS 59
c − a c −a
a−2
c − a c 1−a
a−1
ac ac
(a − 1) a (1 − a) a p
p q p q
a c a c − a c −a
a q
a−1 −a
c − ac
ac p q
(1 − a) a − q
p q c
p q 0
Eliminating m from the first two equations, by substitution, we obtain the two–by-two linear
system in the variables x , y : (y
=6,
x
x1/4 y 3/4 = 1 .
Solving the 2 × 2 system and recovering m from m = (4 y 1/4 )/(3 x1/4 ) , the critical point is
found:
4 × 21/4
x = 6−3/4 , y = 61/4 , m= .
33/4
The Bordered Hessian is:
3 m y 3/4 y 3/4
3m
−
16 x7/4 16 x3/4 y 1/4 4 x3/4
3m 3 m x1/4 3 x1/4
Λ= −
.
16 x3/4 y 1/4 16 y 5/4 4 y 1/4
y 3/4 3 x1/4
0
4 x3/4 4 y 1/4
Evaluating Λ at the critical point and computing its determinant:
3 (3)1/4
det Λ = − ,
23/4
we see that we found a minimum.
Example 3.30. The problem presented here is typical in the determination of an optimal
investment portfolio in Corporate Finance.
We seek to minimize f (x, y) = x2 + 2 y 2 + 3 z 2 + 2 xz + 2 y z with the constraints:
x+y+z =1 2x + y + 3z = 7.
60 CHAPTER 3. MULTIDIMENSIONAL DIFFERENTIAL CALCULUS
L(x , y , z ; m , n) = x2 + 2 y 2 + 3 z 2 + 2 x z + 2 y z − m (x + y + z − 1) − n (2 x + y + 3 z − 7) ,
x = 0, y = −2 , z = 3, m = −10 , n = 8.
The convexity of the objective function ensures that the found solution is the absolute minimum.
Though this statement should be proved rigorously, we do not treat it here.
4 First order equations: general the-
ory
Our goal, in introducing ordinary differential equations, is to provide a brief account on methods
of explicit integration, for the most common types of ordinary differential equations. However,
it is not taken for granted the main theoretical problem, concerning existence and uniqueness
of the solution of the Initial Value Problem, modelled by (4.3). Indeed, the proof of the Picard–
Lindelöhf Theorem 4.17 is presented in detail: to do this, we will use some notions from the
theory of uniform convergence of sequences of functions, already discussed in Theorem 2.15. An
abstract approach followed, for instance, in Chapter 2 of [39], is avoided here.
In the following Chapter 5, we present some classes of ordinary differential equations for which,
using suitable techniques, the solution can be described in terms of known functions: in this
case, we say that we are able to find an exact solution of the given ordinary differential equation.
61
62 CHAPTER 4. FIRST ORDER EQUATIONS: GENERAL THEORY
Definition 4.1. Given f : Ω ⊂ R2 → R , being Ω an open set, the initial value problem (also
called Cauchy problem) takes the form:
(
y 0 = f (x , y) , x∈I ,
(4.3)
y(x0 ) = y0 , x 0 ∈ I , y0 ∈ J ,
where I × J ⊆ Ω are intervals, and where we have simply denoted y in place of y(x) .
Remark 4.2. We say that differential equations are studied by quantitative or exact methods
when they can be solved completely, that is to say, all their solutions are known and could be
written in closed form, in terms of elementary functions or, at times, in terms of special functions
(or in terms of inverses of elementary and special functions).
We can also follow a reverse approach, in the sense that, as illustrated in Example 4.4, given a
geometrical locus, we obtain its ordinary differential equation.
y = α x2 . (4.5)
Any parabola in the family has the y-axes as common axis, with vertex in the origin. Differen-
tiating, we get:
y0 = 2 α x . (4.6)
Eliminating α from (4.5) and (4.6), we obtain the differential equation:
2y
y0 = . (4.7)
x
This means that any parabola in the family is solution to the differential equation (4.7).
4.1. PRELIMINARY NOTIONS 63
Remark 4.7. In the Lotka–Volterra case (4.8), it is n = 2 , thus y = (y1 , y2 ) , the open set
is Ω = R × (0 , +∞) × (0 , +∞) and the continuos function is f (x , y) = f (x , y1 , y2 ) =
y1 (a − b y2 ) , y2 (c y1 − d) .
The rigorous definition of initial value problem for a differential equation of order n is provided
below.
Definition 4.8. Consider an open set Ω ⊆ R × Rn , where n ≥ 1 is integer. Let F : Ω → R
be a scalar continuous function of (n + 1)–variables. Let further (x0 , b) ∈ Ω and I be an open
interval such that x0 ∈ I . Finally, denote b = (b1 , . . . , bn ) .
Then, a real function s : I → R is a solution of the initial value problem:
y (n) = F (x , y , y 0 , y 00 , · · · , y (n−1) )
y(x0 ) = b1
y 0 (x0 ) = b2 (4.10)
...
(n−1)
y (x0 ) = bn
64 CHAPTER 4. FIRST ORDER EQUATIONS: GENERAL THEORY
if:
(i) s ∈ C n (I) ;
(2) all solutions to (4.11) can be expressed as functions of the family itself, i.e., they take the
form y(x ; c1 , . . . , cn ) .
Remark 4.10. Systems of first–order differential equations like (4.9) and equations of order n
like (4.10) are intimately related. Given the n-th order equation (4.10), in fact, an equivalent
system can be build, that has form (4.9), by introducing a new vector variable z = (z1 , . . . , zn )
and considering the system of differential equations:
z10 = z2
z20 = z3
... (4.12)
0
zn−1 = zn
0
zn = F (x , z1 , z2 , . . . , zn )
System (4.12) can be represented in the vectorial form (4.9), simply by setting z 0 = (z10 , . . . , zn0 ) ,
b = (b1 , . . . , bn ) and:
z2
z3
f (x , z) =
... .
zn
F (x , z1 , z2 , . . . , zn )
Form Remark 4.10, the following Theorem 4.11 can be inferred, whose straightforward proof is
omitted.
Theorem 4.11. Function s is solution of the n–th order initial value problem (4.10) if and
only if the vector function z solves system (4.12), with the initial conditions (4.13).
Remark 4.12. It is also possible to go in the reverse way, that is to say, any system of n
differential equations, of first order, can be transformed into a scalar differential equation of order
4.2. EXISTENCE OF SOLUTIONS: PEANO THEOREM 65
n . We illustrate this procedure with the Lotka-Volterra system (4.8). The first step consists in
computing the second derivative, with respect to x , of the first equation in (4.8):
Then, the values of y10 and y20 from (4.8) are inserted in (4.8a), yielding:
y100 = y1 (a − b y2 )2 + b y2 (d − c y1 ) .
(4.8b)
Thirdly, using again the first equation in (4.8), y2 is expressed in terms of y1 and y10 , namely:
a y1 − y10
y10 = y1 (a − b y2 ) =⇒ y2 = . (4.8c)
b y1
Finally, (4.8c) is inserted into (4.8b), which provides the second–order differential equation for
y1 :
y 02
y100 = a y1 − y10 (d − c y1 ) + 1 .
(4.8d)
y1
Theorem 4.13. Consider the rectangle R = [x0 −a , x0 +a]×[y0 −b , y0 +b] , and let f : R → R
be continuos. Then, the initial value problem (4.3) admits at least a solution in a neighborhood
of x0 .
To extend Theorem 4.13 to systems of ordinary differential equations, the rectangle R is replaced
by a parallelepiped, obtained as the Cartesian product of a real interval with an n–dimensional
closed ball.
Remark 4.15. Under the sole continuity assumption, a solution needs not to be unique. Con-
sider, for example, the initial value problem:
(
y 0 (x) = 2 |y(x)| ,
p
(4.14)
y(0) = 0 .
The zero function y(x) = 0 is a solution of (4.14), which is solved, though, by function y(x) =
x |x| as well. Moreover, for each pair of real numbers α < 0 < β , the following ϕα ,β (x) function
solves (4.14) too:
2
−(x − α)
if x<α,
ϕα ,β (x) = 0 if α ≤ x ≤ β ,
(x − β) 2 if x>β .
In other words, the considered initial value problem admits infinite solutions. This phenomenon
is known as Peano funnel.
1
Giuseppe Peano (1858–1932), Italian mathematician and glottologist.
66 CHAPTER 4. FIRST ORDER EQUATIONS: GENERAL THEORY
R = [x0 , x0 + a] × [y0 − b , y0 + b] .
Proof. The proof is somewhat long, so we present it splitted into four steps.
First step. Let n ∈ N . Define the sequence of functions (un ) by recurrence:
u0 (x) = y0 ,
Z x
un+1 (x) = y0 +
f X , un (X) dX .
x0
We want to show that x , un (x) ∈ R for any x ∈ [x0 , x0 + α] . To this aim, it is enough to
prove that, for n ≥ 0 , the following inequality is verified:
It is precisely here that we can understand the reason of the peculiar definition (4.16) of the
number α , as such a choice turns out appropriate in correctly defining each (and any) term in
the sequence un . It also highlights the local nature of the solution of the initial value problem
(4.3).
Second step. We now show that (un ) converges uniformly on [x0 , x0 + α] . The identity:
n
X
un = u0 + (u1 − u0 ) + · · · + (un − un−1 ) = u0 + (uk − uk−1 )
k=1
suggests that any sequence (un ) can be thought of as an infinite series: its uniform convergence,
thus, can be proved by showing that the following series (4.18) converges totally on [x0 , x0 + α] :
∞
X
(uk − uk−1 ) . (4.18)
k=1
Ln−1 |x − x0 |n
|un (x) − un−1 (x)| ≤ M , for any x ∈ [x0 , x0 + α] . (4.19)
n!
We proceed by induction. For n = 1 , the bound is verified, since:
We now prove (4.19) for n + 1 , assuming that it holds true for n . Indeed:
Z x
|un+1 (x) − un (x)| =
f X , un (X) − f X , un−1 (X) dX
Z xx0
≤ f X , un (X) − f X , un−1 (X) dX
x0
Z x
≤L |un (X) − un−1 (X)| dX
x0
Ln−1 x
Ln |x − x0 |n+1
Z
≤M |X − x0 |n dX = M .
n! x0 (n + 1)!
Therefore (4.19) is proved and implies that series (4.18) is totally convergent for [x0 , x0 + α] ;
in fact:
∞ ∞ ∞
X X X Ln−1 αn
(un − un−1 ) ≤ sup |un − un−1 | ≤ M
n!
n=1 n=1 [x0 ,x0 +α] n=1
∞
M X (L α)n M
eα L − 1 < +∞ .
= =
L n! L
n=1
Third step. We show that the limit of the sequence of functions (un ) solves the initial value
problem (4.3). From the equality:
Z x
lim un+1 (t) = lim y0 + f X , un (X) dX ,
n→∞ n→∞ x0
since f X , un (X) ≤ M ensures uniform convergence for f X , un (X) .
n∈N
Now, differentiating both sides of (4.20), we see that u(x) is solution of the initial value problem
(4.3).
Fourth step. We have to prove uniqueness of the solution of (4.3). By contradiction, assume
that v ∈ C 1 ([x0 , x0 + α] , R) solves (4.3) too. Thus:
Z x
v(x) = y0 + f X , v(X) dX .
x0
As before, it is possible to show that, for any n ∈ N and any x ∈ [x0 , x0 + α] , the following
inequality holds true:
Ln |x − x0 |n
|u(x) − v(x)| ≤ K , (4.21)
n!
where K is given by:
K= max |u(x) − v(x)| .
x∈[x0 ,x0 +α]
Indeed: Z x
|u(x) − v(x)| ≤ f X , u(X) − f x , v(X) dX ≤ K L |x − x0 | ,
x0
which proves (4.21) for n = 1 . Using induction, if we assume that (4.21) is satisfied for some
n ∈ N , then:
Z x
|u(x) − v(x)| ≤ f X , u(X) − f X , v(X) dX
x0
Z x Z x
Ln (X − x0 )n
≤L |u(X) − v(X)| dX ≤ L K dX .
x0 x0 n!
After calculating the last integral in the above inequality chain, we arrive at:
Ln+1 (x − x0 )n+1
|u(x) − v(x)| ≤ K ,
(n + 1)!
which proves (4.21) for index n + 1 . By induction, (4.21) holds true for any n ∈ N .
We can finally end our demonstration of Theorem 4.17. In fact, by taking the limit n → ∞ in
(4.21), we obtain that, for any x ∈ [x0 , x0 + α] , the following inequality is verified:
|u(x) − v(x)| ≤ 0 ,
Remark 4.18. Let us go back to Remark 4.15. In such ap situation, where the initial value
problem (4.14) has multiple solutions, function f (x , y) = 2 |y| does not fulfill the Lipschitz
continuity property. In fact, taking, for istance, y1 , y2 > 0 yields:
|f (y1 ) − f (y2 )| 2
= √ √ ,
|y1 − y2 | y1 − y2
which is unbounded.
The proof of Theorem 4.17, based on successive Picard iterates, is also useful in some simple
situations, where it allows to compute an approximate solution of the initial value problem (4.3).
This is illustrated by Example 4.19.
4.3. EXISTENCE AND UNIQUENESS: PICARD–LINDELÖHF THEOREM 69
x2 x4 x6 x8 (−1)n x2n
yn (x) = 1 − + − + + ··· + .
1! 2! 3! 4! n!
The sequence of Picard–Lindelhöf iterates converges only if it also converges the series:
m
X (−1)n x2n
y(x) := lim .
m→∞ n!
n=0
it follows:
∞
X (−x2 )n 2
y(x) = = e−x .
n!
n=0
We will show later, in Example 5.4, that this function is indeed the solution to (4.22).
is solved by the function y(x) = tanh x , so that its interval of existence is R . It is worth noting
that the similar initial value problem:
(
y0 = 1 + y2
y(0) = 0
behaves differently,
since it is solved by y(x) = tan x and has, therefore, interval of existence
given by − π2 , π2 . Moreover, in this latter case, taking the limit at the boundary of the interval
of existence yields:
limπ y(x) = limπ tan x = ±∞.
x→± 2 x→± 2
This is not a special situation. Even when the interval of existence is bounded, for some theoret-
ical reason that we present later, in detail, the solution can be unbounded; this case is referred
to as a blow-up phenomenon.
R = [x0 , x0 + a] × B b (y 0 ) .
The vector–valued version of the Picard–Lindelöhf Theorem 4.17 is represented by the following
Theorem 4.21, whose proof is omitted, as it is very similar to that of Theorem 4.17.
b
M = max ||f || , α = min {a , }. (4.23)
R M
is also a solution of y 0 = f (x , y) .
4.3. EXISTENCE AND UNIQUENESS: PICARD–LINDELÖHF THEOREM 71
Rn
Ω
y1
y0
x0 x1 R
Definition 4.23 (Maximal domain solution). If u ∈ C 1 (I , Rn ) solves the initial value problem
(4.9), we say that u has maximal domain (or that u does not admit a continuation) if there
exists no function v ∈ C 1 (J , Rn ) which also solves (4.9) and such that I ⊂ J .
The existence of the maximal domain solution to IVP (4.9) can be understood euristically, as it
comes from indefinitely repeating the continuation procedure. Establishing it with mathematical
rigor is beyond the aim of these lecture notes, since it would require notions from advanced
theoretical Set theory, such as Zorn’s Lemma4 .
We end this section stating, in Theorem 4.24, a result on the asymptotic behaviour of a solution
with maximal domain, in the particular case where Ω = I × Rn , being I an open interval. Such
a result explains what observed in Example 4.20, though we do not provide a proof of Theorem
4.24.
Theorem 4.24. Let f be defined on an open set I × Rn ⊂ R × Rn . Given (x0 , y 0 ) ∈ I × Rn ,
assume that function y is a maximal domain solution of the initial value problem:
(
y 0 = f (x , y) ,
y(x0 ) = y 0 .
Denote (α , ω) the maximal domain of y . Then, one of two possibility holds respectively for α
and for ω :
4
See, for example, mathworld.wolfram.com/ZornsLemma.html
72 CHAPTER 4. FIRST ORDER EQUATIONS: GENERAL THEORY
In the previous Chapter 4 we exposed the general theory, concerning conditions for existence and
uniqueness of an initial value problem. Here, we consider some important particular situations,
in which, due to the structure of certain kind of scalar ordinary differential equations, it is
possible to establish methods to determine their explicit solution
where a(x) and b(y) are continuous functions, respectively defined on intervals Ia and Ib ,
such that x0 ∈ Ia and y0 ∈ Ib .
To obtain existence and uniqueness in the solution of (5.1), we have to assume that b(y0 ) 6= 0 .
b(y) 6= 0 , (5.2)
then the unique solution to (5.1) is function y(x) , defined implicitly by:
Z y Z x
dz
= a(s) ds . (5.3)
y0 b(z) x0
∂F (x, y) 1
= ,
∂y b(y)
it follows:
∂F (x0 , y0 ) 1
= 6= 0 .
∂y b(y0 )
73
74 CHAPTER 5. FIRST ORDER EQUATIONS: EXPLICIT SOLUTIONS
We can thus invoke the Implicit function theorem 3.25 and infer the existence of δ , ε > 0 for
which, given any x ∈ (x0 − δ , x0 + δ) , there is a unique C 1 function y = y(x) ∈ (y0 − ε , y0 + ε)
such that:
F (x, y) = 0
and such that, for any x ∈ (x0 − δ, x0 + δ) :
0 Fx x , y(x) a(x)
y (x) = − = = a(x) b(y) .
Fy x , y(x) 1
b(y)
Function y , implicitly defined by (5.3), is thus a solution of (5.1); to complete the proof, we
still have to show its uniqueness. Assume that y1 (x) and y2 (x) are both solutions of (5.1), and
define: Z y
dz
B(y) := ;
y0 b(z)
then:
d y10 (x) y20 (x)
B y1 (x) − B y2 (x) = −
dx b y1 (x) b y2 (x)
a(x) b y1 (x) a(x) b y2 (x)
= − = 0.
b y1 (x) b y2 (x)
Notice that
we used the fact that both y1 (x) and y2 (x) are assumed to solve (5.1). Thus
B y1 (x) − B y2 (x) is a constant function, and its constant value is zero, since y1 (x0 ) =
y2 (x0 ) = y0 . In other words, we have shown that, for any x ∈ Ia :
B y1 (x) − B y2 (x) = 0 ,
At this point, using the Mean–Value Theorem 3.24, we infer the existence of a number X(x)
between the integration limits y2 (x) and y1 (x) , such that:
1
y1 (x) − y2 (x) = 0 .
b X(x)
Example 5.4. Consider once more the IVP studied, using successive approximations, in Ex-
ample 4.19: (
y 0 (x) = −2 x y(x) ,
y(0) = 1 .
Setting a(x) = −2 x , b(y) = y , x0 = 0 , y0 = 1 in (5.3) leads to:
Z y Z x
1 2
dz = (−2 z) dz ⇐⇒ ln y = −x2 ⇐⇒ y(x) = e−x .
1 z 0
5.1. SEPARABLE EQUATIONS 75
In the next couple of examples, some interesting, particular cases of separable equations are
considered.
Example 5.5. The choice b(y) = y in (5.3) yields the particular separable equation:
(
y 0 (x) = a(x) y(x) ,
(5.4)
y(x0 ) = y0 ,
thus: Z x
a(s) ds
x0
y = y0 e .
x
For instance, if a(x) = − , the initial value problem:
2
( x
y 0 (x) = − y(x)
2
y(0) = 1
has solution:
x2
y(x) = e− 4 .
Example 5.6. In (5.3), let b(y) = y 2 , which leads to the separable equation:
(
y 0 (x) = a(x) y 2 (x) ,
(5.5)
y(x0 ) = y0 ,
has solution
1
y= .
1 + x2
We now provide some practical examples, recalling that a complete treatment needs, both,
finding the analytical expression of the solution and determining the maximal solution domain.
y(−1) = 0 .
76 CHAPTER 5. FIRST ORDER EQUATIONS: EXPLICIT SOLUTIONS
From (5.3):
Z y(x) Z x
(z + 1)dz = (s − 1)ds ,
0 0
performing the relevant computations, we get:
1 2 1
y (x) + y(x) = x2 − x ,
2 2
so that: (
p x − 2,
y(x) = −1 ± (x − 1)2 =
−x .
Now, recall that x lies in a neighborhood of zero and that the initial condition requires y(0) = 0 ;
it can be inferred, therefore, that y(x) = −x must be chosen. To establish the maximal domain,
observe that x − 1 vanishes for x = 1 ; thus, we infer that x < 1 .
Example 5.9. As a varies in R , investigate the maximal domain of the solutions to the initial
value problem: (
u0 (x) = a 1 + u2 (x) cos x ,
u(0) = 0 .
Form (5.3), to obtain:
Z u(x) Z x
dz
= a cos s ds.
0 1 + z2 0
After performing the relevant computations, we get:
arctan u(x) = a sin x . (5.6)
It is clear that the Range of the right hand–side of (5.6) is [−a , a] . To obtain a solution defined
π
on R , we have to impose that a < . In such a case, solving with respect to u yields:
2
u(x) = tan (a sin x) .
π π
Viceversa, when a ≥ , since there exists x ∈ R+ for which a sin x = , then, the obtained
2 2
solution is defined in (−x , x) , and x is the minimum positive number verifying the equality
π
a sin x = .
2
5.1. SEPARABLE EQUATIONS 77
5.1.1 Exercises
1. Solve the following separable equations:
x
ex
y 0 (x) = sin2 y(x) , y 0 (x) = ,
y(x) (d) (1 + ex ) cosh y(x)
(a) r
π y(0) = 0 ,
y(0) =
,
2
ex
y 0 (x) = 2 x ,
y 0 (x) = ,
(b) cos y(x) (e) (1 + ex ) cos y(x)
y(0) = 0 ,
y(0) = 5π ,
2x
y 0 (x) =
,
x2 2
cos y(x) 0
y (x) = y (x) ,
(c) (f) 1+x
y(0) = π ,
y(0) = 1 .
4
Solutions:
p 1
(1 + e2 x ) ,
(a) ya (x) = arccos(−x2 ) , (d) yd (x) = arcsinh ln 2
(e) ye (x) = arcsin ln 21 (1 + e2 x ) ,
(b) yf (x) = arcsin x2 + 5 π ,
2
(c) yc (x) = arcsin √1 + x2 ,
(f) yb (x) = .
2 2(1 + x − ln(1 + x)) − x2
e2x
0
y (x) = y(x)
√ 4 + e2x
y(0) = 5
√
is y(x) = 4 + e2 x . What is the maximal domain of such a solution?
y 0 (x) = x sin x
1 + y(x)
y(0) = 0
√
is y(x) = 2 sin x − 2 x cos x + 1 − 1 . Find the maximal domain of the solution.
y 3 (x)
0
y (x) =
1 + x2
y(0) = 2
2
is y(x) = √ . Find the maximal domain of the solution.
1 − 8 arctan x
5. Show that the solution to the initial value problem:
(
y 0 (x) = (sin x + cos x) e−y(x)
y(0) = 1
is y(x) = ln(1 + e + sin x − cos x) . Find the maximal domain of the solution.
78 CHAPTER 5. FIRST ORDER EQUATIONS: EXPLICIT SOLUTIONS
is y(x) = tan x ln(x2 + 1) − 2 x + 2 arctan x + π4 . Find the maximal domain of the solution.
1
is y(x) = − √
3
. Find the maximal domain of the solution.
1 − 2x3
8. Solve the initial value problem:
00 0
y (x) = (sin x) y (x)
y(0) = 1
0
y (1) = 0
Hint. Set z(x) = y 0 (x) and solve the equation z 0 (x) = (sin x) z(x) .
y 0 = (y − 1) (y − 2) . (5.7)
Equation (5.7) is separable, so we can easily adapt formula (5.3), using indefinite integrals and
adding a constant of integration:
y−2
ln = x + c1 .
y−1
Solving for y , the general solution to (5.7) is obtained:
1
y(x) = 1 + (5.8)
1 − cex
where we set c = ec1 .
Observe that the two constant functions y = 1 and y = 2 are solutions of equation (5.7).
Observe further that y = 2 is obtained from (5.8) taking c = 0 , thus such a solution is a
particular solution to (5.7). Viceversa, solution y = 1 cannot be obtained using the general
solution (5.8); for this reason this solution is called singular.
Singular solutions of a differential equation can be found with a computational procedure, illus-
trated in Remark 5.10.
Remark 5.10. Given the differential equation (4.2), suppose that its general solution is given
by Φ(x , y , c) = 0 . When there exists a singular integral of (4.2), it can be detected eliminating
c from the system:
Φ(x , y , c) = 0 ,
∂Φ (5.9)
(x , y , c) = 0 .
∂c
5.3. HOMOGENEOUS EQUATIONS 79
leads to x u0 (x) + u(x) = f x , x u(x) = f 1 , u(x) , where we used the fact that f (α x , α y) =
x y(x)
y(2) = 2 .
x2 + y 2
f (x, y) = .
xy
u0 (x) = 1 ,
x u(x)
u(2) = 1 ,
yielding: p
2 ln |x| + 1 − ln 4 .
u(x) =
√ 2
Observe that the solution is defined on the interval x > eln 4−1 = √ .
e
At this point, going back to our original initial value problem, we arrive at the solution of the
homogeneous problem:
p 2
y(x) = x 2 ln |x| + 1 − ln 4 , x> √ .
e
5.3.1 Exercises
Solve the following initial values problems for homogeneous equations:
y 2 (x) y(x)
1
y 0 (x) = , y 0 (x) =
y(x) + x e x ,
2
x − x y(x) x
1. 3.
y(1) = 1 ,
y(1) = −1 ,
3
1 15 x + 11 y(x)
0 2 y 0 (x) = −
y (x) = − 2 x y(x) + y (x) ,
,
2. x 4. 9 x + 5 y(x)
y(1) = 1 , y(1) = 1 ,
Solutions:
√
1 + 3x2 − 1 3. y(x) = −x ln (e − ln x) ,
1. y(x) = ,
x x 1
2x 4. = ,
2 23/5 y 2/5 y 3/5
2. y(x) = 2 , 1+ 3+
3x − 1 x x
5.4. QUASI HOMOGENEOUS EQUATIONS 81
where
a b
det 6 0.
= (5.14)
α β
In this situation, the linear system:
(
ax + by + c = 0
(5.15)
αx + βy + γ = 0
y 0 = 3 x + 4 ,
y−1
y(0) = 2 .
4
whose solution is x1 = − , y1 = 1 . In the second step, the change variable is performed:
3
(
X = x + 43 ,
Y =y−1 ,
and simplifying:
8
3 x + x = y 2 − 2y
2
3
yields:
p
y =1± 3 x2 + 8 x + 1 .
The worked–out Example 5.15 illustrate the procedure to be followed if, when considering equa-
tion (5.13), condition (5.14) is not fulfilled.
y 0 = − x + y + 1 ,
2x + 2y + 1
y(1) = 2 .
In this situation, since the two equations in the system are proportional, the change of variable
to be employed is: (
t = x,
z = x+y.
The given differential equation is, hence, transformed into the separable one:
z
z 0 = ,
2z + 1
z(1) = 3 .
Thus:
2 z − 6 + ln z − ln 3 = t − 1 =⇒ x + 2 y + ln(x + y) = 5 + ln 3 .
Observe that, in this example, it is not possible to express the dependent variable y in an
elementary way, i.e., in terms of elementary functions.
Φ x , y(x) = y 2 + (x2 − 1) y − x3 = c .
(5.17)
Thus (5.17) is an implicit solution for the differential equation (5.16); if an initial condition is
assigned, we can determine c .
It is not always possible to determine an explicit solution, expressed in terms of y . In the
particular situation of Example 5.16, though, this is feasible and we are able to find an explicit
solution. For instance, setting y(0) = 1 , we get c = 0 and:
1 p
y(x) = 1 − x2 + x4 + 4 x3 − 2 x2 + 1 .
2
Let us, now, leave the particular case of Example 5.16, and return to the general situation, i.e.,
consider ordinary differential equation of the form:
M (x , y) + N (x , y) y 0 = 0 . (5.18)
We call exact the differential equation (5.18), if there exists a function Φ(x , y) such that:
∂Φ ∂Φ
= M (x , y) , = N (x , y) , (5.19)
∂x ∂y
and the (implicit) solution to an exact differential equation is constant:
Φ(x , y) = c .
In other words, finding Φ(x , y) constitues the central task in determining whether a differential
equation is exact and in computing its solution.
Establishing a necessary condition for (5.18) to be exact is easy. In fact, if we assume that (5.18)
is exact and that Φ(x , y) satisfies the hypotheses of Theorem 3.4, then the equality holds:
∂ ∂Φ ∂ ∂Φ
= .
∂x ∂y ∂y ∂x
1
See, for example, mathworld.wolfram.com/ChainRule.html
84 CHAPTER 5. FIRST ORDER EQUATIONS: EXPLICIT SOLUTIONS
The result in Theorem 5.17 speeds up the search of solution for exact equations.
y 0 = − M (x , y) ,
N (x , y) (5.21)
y(x0 ) = y0 , (x0 , y0 ) ∈ Q .
∂M (x , y)
M (x , y) = 6 x + y 2 =⇒ = 2y,
∂y
∂N (x , y)
N (x , y) = 2 x y + 1 =⇒ = 2y.
∂x
Formula (5.22) then yields:
Z x Z x
M (t , 1) dt = (6 t + 1) dt = −4 + x + 3 x2 ,
Z 1y Z1 y
N (x , s) ds = (2 x s + 1) ds = −1 − x + y + x y 2 .
1 1
Hence, the solution to the given initial value problem is implicitly defined by:
x y 2 + y + 3 x2 − 5 = 0
3 y e3 x − 2 x
0
y =− ,
e3 x
y(1) = 1 .
5.6. INTEGRATING FACTOR FOR NON EXACT EQUATIONS 85
Here, it holds:
∂M (x , y)
M (x , y) = 3 y e3 x − 2 x =⇒ = 3 e3 x ,
∂y
∂N (x , y)
N (x , y) = e3 x =⇒ = 3 e3 x .
∂x
Using formula (5.22):
Z x Z x
M (t , 1) dt = (3 e3 t − 2 t) dt = −x2 + e3 x − e3 + 1 ,
Z 1y Z1 y
N (x , s) ds = (e3 x ) ds = (y − 1) e3 x .
1 1
−x2 + e3 x − e3 + 1 + (y − 1) e3 x = 0 y = e−3 x x2 + e3 − 1 .
=⇒
5.5.1 Exercises
1. Solve the following initial value problems, for exact equations:
3 2
y 0 = 9 x − 2 x y ,
y 0 = − 2 x + 3 y ,
(a) 3x + y − 1 (d) x2 + 2 y + 1
y(1) = 2 , y(0) = −3 ,
2
2
y 0 = 9 x − 2 x y ,
y 0 = 2 x y + 4 ,
(e) 2 (3 − x2 y)
(b) x2 + 2 y + 1
y(−1) = 8 ,
y(0) = −3 ,
2xy
2
− 2x
y 0 = 2 x y + 4 , 2
0 = 1+x
(f) y ,
(c) 2 (3 − x2 y)
2 − ln (1 + x2 )
y(−1) = −8 , y(0) = 1 .
2. Using the method for exact equation, described in this § 5.5, prove that the solution of the
initial value problem:
1 − 3 y 3 e3 x y
y 0 =
3 x y 2 e3 x y + 2 y e3 x y
y(0) = 1
is implicitly defined by y 2 e3 x y −x = 1 , and verify this result using the Dini Implicit function
Theorem 3.25.
∂ ∂
M (x, y) + N (x, y) y 0 = 0 , N (x, y) = M (x , y) .
∂x ∂y
The case is more frequent, though, in which condition (5.19) is not satisfied, so that we are
unable to express the solution of the given differential equation in terms of the known functions.
There is, however, a general method of solution which, at times, allows the solution of the
general differential equation to be formulated using the known functions. In formula (5.18a)
86 CHAPTER 5. FIRST ORDER EQUATIONS: EXPLICIT SOLUTIONS
below, although it can hardly be considered as an orthodox procedure, we split the derivative
y 0 and, then, rewrite (5.18) in the so–called Pfaffian2 form:
M (x , y) dx + N (x , y) dy = 0 . (5.18a)
We do not assume condition (5.20). In this situation, there exists a function µ(x , y) such that,
multiplying both sides of (5.18a) by µ , an equivalent equation is obtained which is exact,
namely:
µ(x , y) M (x, y) dx + µ(x , y) N (x, y) dy = 0 . (5.18b)
This represents a theoretical statement, in the sense that it is easy to formulate conditions that
need to be satisfied by the integrating factor µ , namely:
∂ ∂
µ(x , y) N (x, y) = µ(x , y) M (x, y) . (5.23)
∂x ∂y
Evaluating the partial derivatives (and employing a simplified subscript notation for partial
derivatives), the partial differential equation for µ is obtained:
M (x , y) µy − N (x , y) µx = Nx (x , y) − My (x , y) µ . (5.23a)
Notice that solving (5.23a) may turn out to be harder than solving the original differential
equation (5.18a). However, depending on the particular structure of the functions M (x , y) and
N (x , y) , there exist favorable situations in which it is possibile to detect the integrating factor
µ(x , y) , provided that some restrictions are imposed on µ itself. In the following Theorems 5.20
and 5.23, we describe what happens when µ depends on one variable only.
Theorem 5.20. Equation (5.18a) admits an integrating factor µ depending on x only, if the
quantity:
My (x , y) − Nx (x , y)
ρ(x) = (5.24)
N (x , y)
also depends on x only. In this case, it is:
Z
ρ(x) dx
µ(x) = e , (5.25)
Proof. Assume that µ(x, y) is a function of one variable only, say, it is a function of x only,
thus:
dµ
µ(x , y) = µ(x) , µx = = µ0x , µy = 0 .
dx
In this situation, equation (5.23a) reduces to:
N (x , y) µ0x = My (x , y) − Nx (x , y) µ ,
(5.23b)
that is:
µ0x My (x , y) − Nx (x , y)
= . (5.23c)
µ N (x , y)
Now, if the left hand–side of (5.23c) depends on x only, then (5.23c) is separable: solving it
leads to the integrating factor represented in thesis (5.25).
2
Johann Friedrich Pfaff (1765–1825), German mathematician.
5.6. INTEGRATING FACTOR FOR NON EXACT EQUATIONS 87
x (x − y) (5.26)
y(1) = 3 .
Equation (5.26) is not exact, nor separable. Let us rewrite it in Pfaffian form, temporarily
ignoring the initial condition:
(3 x y − y 2 ) dx + x (x − y) dy = 0 . (5.26a)
My (x , y) − Nx (x , y) 3 x − 2 y − (2 x − y) 1
= = ,
N (x , y) x (x − y) x
which is a function of x only. The hypotheses of Theorem 5.20 are fulfilled, and the integrating
factor comes from (5.25): Z
1
dx
x
µ(x) = e = x.
Multiplying equation (5.26a) by the integrating factor x , we form an exact equation, namely:
(3 x2 y − x y 2 ) dx + x2 (x − y) dy = 0 . (5.26b)
M1 (x , y) = x M (x , y) = 3 x2 y − x y 2 ,
N1 (x, y) = x N (x , y) = x2 (x − y) ,
and employ them in equation (5.22), which also incorporates the initial condition:
Z x Z y
M1 (t , 3) dt + N1 (x , s) ds = 0 ,
1 3
that is: Z x Z y
2
(−9 t + 9 t ) dt + (x2 − x s) ds = 0 .
1 3
Evaluating the integrals:
x2 y 2 3
x3 y − + = 0.
2 2
Solving for y , and recalling the initial condition, leads to the solution of the initial value problem
(5.26): √
3 + x4
y =x+ .
x
Example 5.22. Consider the following initial value problem, in which the differential equation
is not exact nor separable:
2
y 0 = − 4 x y + 3 y − x ,
x (x + 2y) (5.27)
y(1) = 1 .
Since the hypotheses of Theorem 5.20 are fulfilled, the integrating factor is given by (5.25):
Z
2
dx
x
µ(x) = e = x2 .
M1 (x , y) = x M (x , y) = 4 x3 y + 3 x2 y 2 − x3 ,
N1 (x , y) = x N (x , y) = x3 (x + 2 y) ,
we can use them into equation (5.22), which also incorporates the initial condition, obtaining:
Z x Z y
M1 (t , 1) dt + N1 (x , s) ds = 0 ,
1 1
that is: Z x Z y
2 3
(3 t + 3 t ) dt + x2 (2 s + x) ds = 0 .
1 1
Evaluating the integrals yields:
x4 7
x4 y − + x3 y 2 − = 0 .
4 4
Solving for y and recalling the initial condition, we get the solution of the initial value problem
(5.27): √
x5 + x4 + 7 x
y= − .
2x3/2 2
We examine, now, the case in which the integrating factor µ is a function of y only. Given the
analogy with Theorem 5.20, the proof of Theorem 5.23 is not provided here.
Theorem 5.23. Equation (5.18a) admits an integrating factor µ depending on y only, if the
quantity:
Nx (x , y) − My (x , y)
ρ(y) = (5.28)
M (x , y)
also depends on y only. In this case, the integrating factor is:
Z
ρ(y) dy
µ(y) = e , (5.29)
y 2 (x + y + 1) dx + x y (x + 3 y + 2) dy = 0 .
5.6. INTEGRATING FACTOR FOR NON EXACT EQUATIONS 89
M1 (x , y) = x M (x , y) = y 2 (x + y + 1) ,
N1 (x , y) = x N (x , y) = x y (x + 3 y + 2) ,
and employ them into equation (5.22), which also incorporates the initial condition, obtaining:
Z x Z y
M1 (t , 1) dt + N1 (x , s) ds ,
1 1
that is: Z x Z y
(2 + t) dt + s x (2 + 3 s + x) ds = 0 .
1 1
x2 y 2 5
+ x y3 + x y2 − = 0 .
2 2
To end this § 5.6, let us consider the situation of a family of differential equations for which an
integrating factor µ is available.
Theorem 5.25. Let Q = { (x , y) ∈ R2 | 0 < a < x < b , 0 < c < x < d } , and let f1 and
f2 be C 1 functions on Q , such that f1 (x y) − f2 (x y) 6= 0 . Define the functions M (x , y) and
N (x , y) as:
M (x , y) = y f1 (x y) , N (x , y) = x f2 (x y) .
Then:
1
µ(x , y) =
x y f1 (x y) − f2 (x y)
is an integrating factor for:
y f1 (x y)
y0 = − .
x f2 (x y)
Proof. It suffices to insert the above expressions of µ , M and N into condition (5.23) and
verify that it gets satisfied.
5.6.1 Exercises
1. Solve the following initial value problems, using a suitable integrating factor.
2
y 0 = − y (x + y) ,
y 0 = 3 x + 2 y ,
(a) x + 2y − 1 (e) 2xy
y(1) = 1 , y(1) = 1 ,
0 y − 2 x3
y 0 = − y
2
, y = ,
(b) x (y − ln x) (f) x
y(1) = 1 , y(1) = 1 ,
y 0 = y
y 0 = y ,
, (g) 3
y − 3x
(c) y − 3x − 3 y(0) = 1 ,
y(0) = 0 ,
3 x
( y 0 = − y + 2 y e ,
y0 = x − y , (h) ex + 3 y 2
(d)
y(0) = 0 , y(0) = 0 .
90 CHAPTER 5. FIRST ORDER EQUATIONS: EXPLICIT SOLUTIONS
Let functions a(x) and b(x) be continuous on the interval I ⊂ R . The first–order differential
equation (5.31) is called linear, since y is represented by a polynomial of degree 1 . We can
establish a formula for its integration, following a procedure that is similar to what we did for
separable equations.
Theorem 5.26. The unique solution to (5.31) is:
Z x Z t
a(t) dt − a(s) ds
!
Z x
y(x) = e x0 y0 + b(t) e x0 dt
x0
where: Z x
A(x) = a(s) ds . (5.33)
x0
Proof. To arrive at formula (5.32), we first examine the case b(x) = 0 , for which (5.31) reduces
to the separable (and linear) equation:
(
y 0 (x) = a(x) y(x) ,
(5.34)
y(x0 ) = c ,
To find the solution to the more general differential equation (5.31), we use the method of
Variation of Parameters 3 , due to Lagrange: we assume that c is a function of x and search
for c (x) such that the function:
y(x) = c(x) eA(x) (5.36)
becomes, indeed, a solution of (5.31). To this aim, differentiate (5.36):
from which:
c0 (x) = b(x) e−A(x) . (5.37)
Integrating (5.37) between x0 and x , we obtain:
Z x
c(x) = b(t) e−A(t) dt + K ,
x0
3
See, for example, mathworld.wolfram.com/VariationofParameters.html
5.7. LINEAR EQUATIONS OF FIRST ORDER 91
Evaluating y(x0 ) and recalling the initial condition in (5.31), we see that y0 = K . Thesis (5.32)
thus follows.
Remark 5.27. An alternative proof to Theorem 5.26 can be provided, using Theorem 5.20 and
the integrating factor procedure. In fact, if we assume:
then:
My − Ny
= −a(x) ,
N
which yields the integrating factor µ(x) = e−A(x) , with A(x) defined as in (5.33). Considering
the following exact equation, equivalent to (5.31):
Remark 5.28. The general solution of the linear differential equation (5.31) can be described
when a particular solution of it is known, together with the general solution of the linear and
separable equation (5.34).
If y1 and y2 are both solutions of (5.31), in fact, there exist v1 , v2 ∈ R such that:
Z x
y1 (x) = eA(x) v1 + b(t) e−A(t) dt ,
x0
Z x
y2 (x) = eA(x) v2 + b(t) e−A(t) dt .
x0
Subtracting, we obtain:
y1 (x) − y2 (x) = v1 − v2 eA(x) ,
which means that y1 − y2 has the form (5.35) and, therefore, solves (5.34). Now, using the fact
that y1 is a solution of (5.31), the general solution to (5.31) can be written as:
and all this is equivalent to saying that the general solution y = y(x) of (5.31) can be written
in the form:
y − y1
= c, c ∈ R.
y2 − y1
Example 5.29. Consider the equation:
3
y 0 (x) = 3 x2 y(x) + x ex ,
y(0) = 1 .
92 CHAPTER 5. FIRST ORDER EQUATIONS: EXPLICIT SOLUTIONS
3
Here, a(x) = 3 x2 and b(x) = x ex . Using (5.32)–(5.33), we get:
Z x Z x
A(x) = a(s) ds = 3 s2 ds = x3
x0 0
and
x x
x2
Z Z
3 3
b(t) e−A(t) dt = t et e−t dt = ,
x0 0 2
so that:
x2
x3
y(x) = e 1+ .
2
Example 5.30. Consider the equation:
(
y 0 (x) = 2 x y(x) + x ,
y(0) = 2 .
and
" 2
#x 2
x x
e−t 1 e−x
Z Z
−A(t) −t2
b(t) e = te dt = − = − .
0 0 2 2 2
0
5.7.1 Exercises
Solve the following initial value problems for linear equations.
1 1
x
0
y (x) = − y(x) + , y 0 (x) =
2
y(x) + 1 ,
1 + x 2 1 + x 2
3. 1 + x
1.
y(0) = 0 ,
y(0) = 0 ,
1
y 0 (x) = − sin x y(x) + sin x ,
y 0 (x) = y(x) + x2 ,
2. 4. x
y(0) = 0 ,
y(1) = 0 ,
5.8. BERNOULLI EQUATION 93
1
3
y 0 (x) = 3 x2 y(x) + x ex ,
y 0 (x) = x −
y(x) ,
5. 6. 3x
y(0) = 1 ,
y(1) = 1 .
is known as Bernoulli equation for the fact that Jacob Bernoulli (1654–1705) first solved it. In
the article [32] the historical process which led to solve the famous Bernoulli differential equation
is exposed in detail.
The change of dependent variable v(x) = y 1−α (x) transforms (5.38) into a linear equation:
The next three examples illustrate how the change of variable works.
Here α = 4 , and the change of variable is v(x) = y −3 (x) , i.e., y = v −1/3 , leading to:
1 −4/3 0 1 1
− v v = − v −1/3 − v −4/3 .
3 x x
1 0 1 1
− v =− v− .
3 x x
Simplifying and recalling the initial condition, we obtain a linear initial value problem:
v 0 = 3 v + 3 ,
x x
v(2) = 1 ,
1 3
solved by v(x) = x − 1 . Hence, the solution to the original problem is:
4
r
4
y(x) = 3 3 .
x −4
We have to solve a Bernoulli equation, with exponent α = 5 . The change of variable is:
1 − 1
y(x) = v(x) 1−5 = v(x) 4 ,
solved by:
x + x3
v(x) = .
2
Recovering y(x) , we find: r
4 2
y(x) =
x + x3
that is defined for x > 0 .
Remark 5.35. Bernoulli equation (5.38) can also be solved using the same approach used for
linear equations, that is, imposing a solution of the form:
so that:
1
1−α
1−α
c(x) = (1 − α) F (x) + c0 ,
where: Z x
F (x) = b(z) e(α−1) A(z) dz .
x0
5.8. BERNOULLI EQUATION 95
Remark 5.36. There is one more way to solve equation (5.38) transforming it into a separable
equation changing the dependent variable as follows
where, as usual Z
A(x) = a(x) dx
Now expressing y 0 in (5.40b ) using the fact that y solves (5.38) we arrive at
Therefore solution y of (5.38) is obtained inserting u from (5.40d ) in (5.40) and solving for y:
Z 1
1−α
A(x) (α−1)A(x)
y=e (1 − α) b(x)e dx + c (5.40e )
5.8.1 Exercises
Solve the following initial value problems for Bernoulli equations.
y 0 = 2 x y + 2√y ,
y 0 = − 1 − 1 y 2 ,
1. x ln x 3. 1+x
y(2) = 1 , y(0) = 1 ,
( 2 (
y 0 = x2 y + x5 + x2 y 3 , y0 = y + x y2 ,
2. 4.
y(0) = 1 , y(0) = 1 .
(
y 0 (x) − y(x) + (cos x) y(x)2 = 0 ,
5.
y(0) = 1 .
96 CHAPTER 5. FIRST ORDER EQUATIONS: EXPLICIT SOLUTIONS
6 First order equations: advanced
topics
The solving strategy is based on knowing one particular solution y1 (x) of (6.1). Then, it is
assumed that the other solutions of (6.1) have the form y(x) = y1 (x) + u(x) , where u(x) is an
unknown function, to be found, and that solves the associated Bernoulli equation:
Notice that, in this latter way, we combine together two substitutions: the first one maps the
Riccati1 equation into a Bernoulli equation; the second one linearizes the Bernoulli equation.
Example 6.1. Knowing that y1 (x) = 1 solves the Riccati equation:
1+x 1
y0 = − + y + y2 , (6.3)
x x
we want to show that the general solution to equation (6.3) is:
x2 e x
y(x) = 1 + .
c + (1 − x)ex
Let us use the change of variable:
1
y(x) = 1 + ,
u(x)
to obtain the linear equation:
1 2
u0 (x) = − − 1+ u(x) . (6.3a)
x x
To solve it, we proceed as we learned. First, compute A(x) :
e−x
Z
2
A(x) = − 1+ dx = −x − 2 ln x =⇒ eA(x) = .
x x2
1
Jacopo Francesco Riccati (1676–1754), Italian mathematician and jurist.
97
98 CHAPTER 6. FIRST ORDER EQUATIONS: ADVANCED TOPICS
Then, form:
Z Z Z
−A(x) 1 2 x
b(x) e dx = − x e dx = − x ex dx = (1 − x) ex .
x
e−x x
c e−x + 1 − x c + (1 − x) ex
u(x) = c + (1 − x) e = = .
x2 x2 x2 e x
1 x2 ex
y(x) = 1 + =1+ .
u(x) c + (1 − x) ex
y(x)
y 0 (x) = −x5 + + x3 y 2 (x) , (6.4)
x
2x
y= + x.
2 x5
2 c e− 5 −1
Remark 6.3. In applications, it may be useful to state conditions on the coefficient functions
a(x) , b(x) and c(x) , to the aim that the relevant Riccati equation (6.1) is solved by some
particular function having simple form. The following list summarizes such conditions and, for
each one, the correspondent simple–form solution y1 .
1. Monomial solution:
if a(x) + xn−1 x b(x) + c(x) xn+1 − n = 0 , then y1 (x) = xn .
2. Exponential solution:
if a(x) + en x b(x) + c(x) en x − n = 0 , then y1 (x) = en x .
4. Sine solution: if a(x) + b(x) sin(n x) + c(x) sin2 (n x) − n cos(n x) = 0 , then y1 (x) =
sin(n x) .
5. Cosine solution: if a(x) + b(x) cos(n x) + c(x) cos2 (n x) + n sin(n x) = 0 , then y1 (x) =
cos(n x) .
6.1. RICCATI EQUATION 99
A consequence of Theorem 6.4 is the so–called Cross–Ratio property of the Riccati equation,
illustrated in the following Corollary 6.5.
Corollary 6.5. Given any four solutions y1 , . . . , y4 of the Riccati equation (6.1), their Cross–
Ratio is constant and is given by the quantity:
y4 − y2 y3 − y1
. (6.6)
y4 − y1 y3 − y2
Proof. Relation (6.5) implies that, if y4 is a fourth solution of (6.1), then:
y4 − y2 y3 − y2
=c ,
y4 − y1 y3 − y1
which, since c is constant, demonstrates thesis (6.6).
u0 = A0 (x) + A1 (x) u2
is known as reduced form of the Riccati equation (6.1). Functions A0 (x) and A1 (x) are related
to functions a(x) , b(x) and c(x) appearing in (6.1). In fact, if B(x) is a primitive of b(x) , i.e.,
B 0 (x) = b(x) , the change of variable:
This can be seen by computing u0 = e−B(x) y 0 − y B 0 (x) from (6.7) and then substituting, in
the factor y 0 − y B 0 (x) , the equalities B 0 (x) = b(x) and y 0 = a(x) + b(x) y + c(x) y 2 , and finally
y = eB(x) u .
Sometimes, given a Riccati equation, its solution can be obtained by simply transforming it to
its reduced form. Example 6.6. illustrates this fact.
Example 6.6. Consider the initial value problem for the Riccati equation:
y 0 = 1 − 1 y + x y 2 ,
2x
y(1) = 0 .
u(1) = 0 ,
c(x) 2 B(x)
e
a(x)
has to be constant and equal to a certain real number λ . The topic of separability of the Riccati
equation is presented in [1, 35, 36, 38, 40].
where P (x) and Q(x) are given continuos functions, defined on an interval I ⊂ R . The term
linear indicates the fact that the unknown function y = y(x) and its derivatives appear in
polynomial form of degree one.
6.1. RICCATI EQUATION 101
The second–order linear differential equation (6.8) is equivalent to a particular Riccati equation.
We follow the fair exposition given in Chapter 15 of [33]. Let us introduce a new variable
u = u(x) , setting: Z
y = e−U (x) , with U (x) = − u(x) dx . (6.9)
simplifying which leads to the non–linear Riccati differential equation of first order:
Viceversa, to find a linear differential equation, of second order, that is equivalent to the first–
order Riccati equation (6.1), let us proceed as follows. Consider the transformation:
w0
y=− , (6.10)
c(x) w
b(x) w0 w2
a(x) − + .
c(x) w c(x) (w0 )2
x 1
y 00 − 2
y0 + y = 0, with − 1 < x < 1. (6.11)
1−x 1 − x2
x 1
P (x) = − , Q(x) = ,
1 − x2 1 − x2
1 x
u0 = 2
+ u + u2 . (6.11a)
1−x 1 − x2
102 CHAPTER 6. FIRST ORDER EQUATIONS: ADVANCED TOPICS
To obtain the reduced form of (6.11a), we employ the transformation (6.7), observing that, here,
such a transformation works in the following way:
Z
x 1
b(x) = =⇒ B(x) = b(x) dx = − ln(1 + x2 )
1 − x2 2
p
=⇒ v = e−B(x) u = 1 − x2 u .
tan (arcsin x + c)
u(x) = √ .
1 − x2
To get the solution of the linear equation (6.11), we reuse relation (6.9).
First, a primitive U (x) of u(x) must be found:
tan arcsin x + c
Z Z
U (x) = u(x) dx = √ dx = − ln cos(arcsin x + c) − K .
1 − x2
where ±K is a constant whose sign can be made positive. Then, using (6.9), we can conclude
that the general solution to (6.11) is:
ln cos(arcsin x+c) +K
y(x) = e = eK cos(arcsin x + c) .
6.1.4 Exercises
1. Knowing that y1 (x) = 1 is a solution of:
x2 y 00 + 3 x y 0 + y = 0 .
Hint: Transform the equation into a Riccati equation, in reduced form, and then use the fact
that v1 (x) = x2 is a particular solution.
(1 + x2 ) y 00 − 2 x y 0 + 2 y = 0 .
Hint: Transform the equation into a Riccati equation, then solve it, using the fact that u1 =
1
− is a particular solution.
x
This last condition ensures uniqueness for the (x, y)–solution of the system:
(
X = X(x, y) ,
Y = Y (x, y) .
Dx Y (x, y) Yx + Yy yx0
=
Dx X(x, y) Xx + Xy yx0
i.e.,
Yx + f (x, y) Yy
YX0 = . (6.14a)
Xx + f (x, y) Xy
The right hand–side of (6.14a) contains (x, y) . To complete the coordinate change, we have to
revert the mapping, solving the system:
( (
X = X(x, y) , x = x̂(X, Y ) ,
=⇒
Y = Y (x, y) , y = ŷ(X, Y ) ,
Later, Joseph Liouville (1809-1882) proved in [27] that, beyond Bernoulli’s (6.17), equation
(6.16) cannot be solved in terms of elementary functions. In the general situation, solution of
(6.16) are expressed through Bessel functions, that we will study in section 8.4.
Before illustrate the Bernoulli argument we provide the integration of (6.16) in the “limit
case”of (6.17) i.e. taking the limit as n → ∞ which means α = −2. We therefore deal with
a
y0 = + by 2 (6.18)
x2
Change the dependent variable by:
1 z0
y= =⇒ y 0 = − 2
z z
then inserting in (6.18) we arrive at
z 2
z 0 = −a −b (6.18’)
x
Use a second change of variable:
z
u=
x
being u the new dependent varaible, obtaing the separable equation
−b − u − au2
u0 = (6.18”)
x
The integration of (6.18”) is elementary and depends on the sign of ∆ = 1 − 4ab. As an example
if a = 1 and b = −6 we get: Z Z
du 1
2
= +c
6−u−u x
thus using partial fraction:
u+3
= (cx)5
u−2
To end the process we go back to the original variable using
z 1 1
u= ,z= =⇒ u =
x y xy
b
y 0 = ay 2 + , a, b ∈ R (6.19)
x4
We use the change of variable suggested in [23]:
1 1
x =
X =
x X
a =⇒ X 1
(6.20)
Y =
,
y = X − .
x (1 + axy)
Y a
106 CHAPTER 6. FIRST ORDER EQUATIONS: ADVANCED TOPICS
a2 b
YX0 = + a = a + bY 2 (6.19a)
x2 (axy + 1)2
Solution of (6.19a) is therefore obtained with the following integration which depends on the
sign of ab as follows
1 pa
Z
dY √
arctan b Y +c if ab > 0
+ c = ab (6.19b)
a + bY 2 1 p a
√
arctanh −b Y + c if ab < 0
−a b
Using (6.19b) we have the solution of the transformed differential equation (6.19a):
r
a √
tan a b X + c if ab > 0
b
Y = r
√ (6.19c)
a
− − tanh −a b X + c if ab < 0
b
How and why Daniel’s change of variables separates the variables in Riccati’s equation (6.19)?
The answer is found, on one hand by returning to the equation (6.16) with arbitrary exponent
α and on the other by parameterizing the change of variables.
Changing variable in (6.16) as
α+3
1
X = x x = 3+α
X
a =⇒ 2 1 (6.21)
Y =
aX − α+3 − Y X − α+3
x(a + xy) y=
aY
We arrive at the transformed equation
α+4
aX − α+3 + bY 2
YX0 =− (6.22)
α+3
Not having specified the parameters allows us to understand α = −4 is “special case”. Moreover
we see that change of variables (6.21) transforms equation (6.16) into an equation of the same
kind, but with a different exponent of x.
Consequently, iterating the process we identify the family of exponents that, after a finite
number of steps, produces a separable equation.
6.2. CHANGE OF VARIABLE 107
It is easily seen that by iterating the transformation of variables the sequence of the exponents
of the independent variable is
α+4 3α + 8 5α + 12 7α + 16 α(2n − 1) + 4n
α, − ,− ,− ,− ..., − , ...
α+3 2α + 5 3α + 7 4α + 9 αn + 2n + 1
Henceforth the iteration will produce a separable is α is such that there is a positive integer n
such that
4n
α(2n − 1) + 4n = 0 =⇒ αn =
1 − 2n
For instance for n = 2 that is α = −8/3 solution of (6.16) is
√ √ √
3√ab
3 3
x ab cot 3x + k + 3ab − x2/3
if ab > 0
√ √
x5/3 − 3x4/3 ab cot 3√ ab
3x + k
y= √ √ √
3 √−ab
3 3
x −ab coth 3x + k + 3ab − x2/3
if ab < 0
√ √
−ab
x5/3 − 3x4/3 −ab coth 3 √
3x + k
6.2.2 Exercises
1. Prove that the differential equation:
y − 4 x y 2 − 16 x3
yx0 =
y 3 + 4 x2 y + x
YX0 = −2 X .
p y
where X(x , y) = 4 x2 + y 2 and Y (x , y) = arctan .
2x
Then, use this fact to integrate the original differential equation.
y 3 + x2 y − y − x
yx0 =
x y 2 + x3 + y − x
YX0 = Y (1 − Y 2 ) ,
y p
where X(x , y) = arctan and Y (x , y) = x2 + y 2 .
x
Then, use this fact to integrate the original differential equation.
y 3 ey
y0 = , (6.23)
y 3 ey − ex
108 CHAPTER 6. FIRST ORDER EQUATIONS: ADVANCED TOPICS
X(x , y) = − 1 ,
y
Y (x , y) = ex−y .
y2
0 y+1
y = 3+ ,
x x (6.24)
y(1) = 0 ,
The general form of a differential equation of order n ∈ N was briefly introduced in equation
(4.10) of Chapter 4. The current Chapter is devoted to the particular situation of linear equations
of second order:
a(x) y 00 + b(x) y 0 + c(x) y = d(x) , (7.1)
where a , b , c and d are continuous real functions of the real variable x ∈ I , being I an interval
in R and a(x) 6= 0 . Equation (7.1) may be represented, at times, in operational notation:
M y = d(x) ,
where M : C 2 (I) → C(I) is a differential operator that acts on the function y ∈ C 2 (I) :
In this situation, existence and uniqueness of solutions are verified, for any initial value problem
associated to (7.1).
Before dealing with the simplest case, in which the coefficient functions a(x) , b(x) , c(x) are
constant, we examine general properties, that hold in any situation. We will study some variable
coefficient equations, that are meaningful in applications. Our treatment can be easily extended
to equation of any order; for details, refer to Chapter 5 of [34] or Chapter 6 of [2].
Ly = 0. (7.4a)
where u(x) is the new dependent variable, while f (x) is a function to be specified, in order to
simplify computations. We find:
L(f u) = f u00 + 2 f 0 + p f u0 + f 00 + p f 0 + q f u = r .
(7.6)
109
110 CHAPTER 7. LINEAR EQUATIONS OF SECOND ORDER
f 00 + p f 0 + q f = 0 . (7.7)
In this way, equation (7.6) becomes easily solvable, since it reduces to a first–order linear equation
in the unknown v = u0 :
f v0 + 2 f 0 + p f v = r .
(7.6a)
At this point, if any particular solution to the homogeneous equation (7.4a) is available, the
solution of the non–homogeneous equation (7.4) can be obtained.
The set of solutions to a homogeneous equation forms a two–dimensional vector space, as illus-
trated in the following Theorem 7.1. The first, and easy, step is to recognize, that given two
solutions y1 and y2 of (7.4a), their linear combination:
y = α 1 y1 + α 2 y2 , α1 , α2 ∈ R ,
and
α1 y10 + α2 y20 = (α1 y1 + α2 y2 )0 .
This shows that α1 y1 + α2 y2 is indeed a solution to (7.4a). This demonstrates also that, when
α2 = 0 , any multiple of one solution of (7.4a) solves (7.4a) too. By iteration, it holds that any
linear combination of solutions of (7.4a) solves (7.4a) too.
Theorem 7.1. Consider the linear differential operator L : C 2 (I) → C(I) , defined by (7.5).
Then, the kernel of L has dimension 2 .
Proof. Fix x0 ∈ I and define the linear operator T : ker(L) → R2 , which maps each function
y ∈ ker(L) onto its initial value, evaluated at x0 , i.e.:
T y = y(x0 ) , y 0 (x0 ) .
The existence and uniqueness Theorem 4.17 means that T y = (0 , 0) implies y = 0 . Hence, by
the theory of linear operators, this mean that T is one–one operator and it holds:
If condition (7.12) holds only when all the αk are zero (i.e., αk = 0 for all k = 1 , . . . , n), then
functions f1 , . . . , fn are linearly independent.
We now provide a sufficient condition for linear independence of a set of functions. Let us as-
sume that f1 , . . . , fn are n–times differentiable. Then, from equation (7.12), applying successive
differentiations, we can form a system of n linear equations in the variables α1 , . . . , αn :
α1 f1 + α2 f2 + . . . + αn fn = 0,
α1 f10 + α2 f20 + ... + αn fn0 = 0,
α1 f100 + α2 f200 + ... + αn fn00 = 0, (7.13)
.. ..
. .
(n−1) (n−1) (n−1)
α1 f1 +α2 f2 + · · · + αn fn = 0.
For example, functions f1 (x) = sin2 x , f2 (x) = cos2 x , f3 (x) = sin(2 x) , are linearly indepen-
dent on I = R , since their Wronskian is non–zero:
f1 (x) f2 (x) f3 (x)
W (f1 , f2 , f3 )(x) = det f10 (x) f20 (x) f30 (x)
f100 (x) f200 (x) f300 (x)
sin2 x cos2 x
sin(2 x)
= det 2 cos x sin x −2 cos x sin x 2 cos(2 x) = 4 .
2 2 2 2
2 cos x − sin x) 2 sin x − cos x) −4 sin(2 x)
Theorem 7.2. Let functions y1 (x) and y2 (x) , defined on the interval I , be solutions to the
linear differential equation (7.4a). Then, a necessary and sufficient condition, for y1 and y2 to
be linearly independent, is provided by their Wronskian being non–zero on I .
Differentiating, we obtain:
d
W 0 (x) = W (y1 , y2 )(x) = y1 (x) y200 (x) − y2 (x) y100 (x) .
dx
Since y1 and y2 are solution to (7.4a), recalling that we assume a(x) 6= 0 in (7.1), it holds:
W 0 = −p(x) W . (7.15)
W (x0 ) = 0 ∀ x0 ∈ I .
By assumption, the determinant of this homogeneous system is zero, hence the system admits a
T
non trivial solution α1 , α2 , with α1 , α2 not simultaneously null. Now, define the function:
Since y(x) is a linear combination of solutions y1 (x) and y2 (x) of (7.4a), then y(x) is also a
solution to (7.4a). And since, by construction, (α1 , α2 ) solves (7.17), then it also holds that:
At this point, from the existence and uniqueness of the solutions of the initial value problem:
(
Ly = 0 ,
y(x0 ) = y 0 (x0 ) = 0 ,
Putting together Theorems 7.1 and 7.2, it is possible to establish if a pair of solutions to (7.4a)
is a basis for the set of solutions to the equation (7.4a), as illustrated in Theorem 7.3.
Theorem 7.3. Consider the linear differential operator L : C 2 (I) → C(I) defined in (7.4a). If
y1 and y2 are two independent elements of ker(L) , then any other element of ker(L) can be
expressed as a linear combination of y1 and y2 :
x2 (1 + x) y 00 − x (2 + 4 x + x2 ) y 0 + (2 + 4 x + x2 ) y = 0 , (7.18)
a second independent solution to (7.18) can be found, by seeking a solution of the form:
Now, introducing v(x) = u0 (x) , we see that v has to satisfy the first–order linear separable
differential equation:
x+2
v0 = v,
x+1
which is solved by:
v(x) = c (1 + x) ex .
We can assume c = 1 , since we are only interested in finding one particular solution of (7.18).
Function u(x) is then found by integration:
Z
u(x) = (1 + x) ex dx = x ex .
Therefore, a second solution to (7.18) is y2 (x) = x2 ex , where, again, we do not worry about
the integration constant. Functions y1 , y2 form an independent set of solutions to (7.18) if their
Wronskian:
x 2 ex
y1 (x) y2 (x) x
det 0 = det = x2 (x + 1) ex .
y1 (x) y20 (x) 1 x (x + 2) ex
is different from zero. Now, observe that the differential equation (7.18) has to be considered
in one of the intervals (−∞ , −1) , (−1 , 0) , (0, +∞) , where the leading coefficient x2 (1 + x) of
(7.18) does not vanish. On such intervals, the Wronskian does not vanish as well, thus f1 , f2
are linearly independent. In conclusion, the general solution to (7.18) is:
y(x) = c1 x + c2 x2 ex , c1 , c2 ∈ R . (7.19)
The procedure illustrated in Example 7.4 can be repeated in the general case. For simplicity, we
recall (7.4a) written in the explicit form:
If a solution y1 (x) of (7.20) is known, we look for a second solution of the form:
Equation (7.21) is a first–order separable equation in the unknown u0 , exactly in the same way
as in the Example 7.4, and it can be integrated to obtain the second solution to (7.20).
The search for a second independent solution to (7.20) can also be pursued using the Wronskian
equation (7.16), without explicitly computing two solutions of (7.20). This is stated in the
following Theorem 7.5.
7.1. HOMOGENEOUS EQUATIONS 115
Theorem 7.5. If y1 (x) is a non–vanishing solution of the second–order equation (7.20), then
a second independent solution is given by:
Z t
−p(s) ds
Z x x0
e
y2 (x) = y1 (x) dt , (7.22)
x0 y12 (t)
Proof. Given the assumption that y1 is a non–vanishing function, rewrite the Wronskian as:
y1 y20 − y2 y10
y y2
W (y1 , y2 ) = det 10 0 = y1 y20 − y2 y10 = y12 ,
y1 y2 y12
setting to zero the constant of integration W (x0 ) . At this point, thesis (7.22) follows from
inserting equation (7.16) into (7.24).
Example 7.6. Consider again Example 7.4. The solution y1 (x) = x of (7.18) can be used in
formula (7.22), to detect a second solution to such equation. Observe that, in this case:
−x (2 + 4 x + x2 ) 2 + 4 x + x2
p(x) = = − ,
x2 (1 + x) x (1 + x)
so that: Z
−p(x) dx = x + 2 ln x + ln(1 + x) ,
and: Z
−p(s) ds
e = x2 (1 + x) ex .
The second solution is, hence:
Z
y2 (x) = x (1 + x) ex dx = x2 ex ,
in accordance with the solution y2 found using the method order reduction in Example 7.4.
Example 7.7. Find the general solution of the homogeneous linear differential equation of
second order:
1 1
y 00 + y 0 + 1 − y = 0.
x 4 x2
We seek a solution of the form y1 = xm sin x . For such a solution, the first and second derivatives
are: (
y10 = xm−1 x cos x + m sin x ,
y(x) = eλ x ,
and imposing that y(x) solves (7.25), leads to the algebraic equation, called characteristic equa-
tion of (7.25):
a λ2 + b λ + c = 0 . (7.26)
The roots of (7.26) determine solutions of (7.25). Namely, if the discriminant ∆ = b2 − 4 a c
is positive, so that equation (7.26) admits two distinct real roots λ1 and λ2 , then the general
solution to (7.25) is:
y = c1 eλ1 x + c2 eλ2 x . (7.27)
The independence of solutions y1 = c1 eλ1 x and y2 = c2 eλ2 x follows from the analysis of their
Wronskian, which is non–vanishing for any x ∈ R :
√
y1 y2 (λ1 +λ2 ) x ∆ −b x
W (y1 , y2 )(x) = det 0 0 = (λ2 − λ1 ) e = e a 6= 0 .
y1 y2 a
7.1. HOMOGENEOUS EQUATIONS 117
When ∆ < 0 , equation (7.26) admits two distinct complex conjugate roots λ1 = α + i β and
λ2 = α − i β , and the general solution to (7.25) is:
y = eα x c1 cos(β x) + c2 sin(β x) .
(7.28)
Forming the complex exponential of λ1 and that of λ2 , two complex valued functions z1 , z2
are obtained:
Set, for example, y1 = eα x cos(β x) and y2 = eα x sin(β x) . Then, the real solution presented
in (7.28) is a linear combination of the real functions y1 and y2 , which are are independent,
since their Wronskian is non–vanishing:
y1 y2
W (y1 , y2 )(x) = det 0 = β e2 α x 6= 0 .
y1 y20
When ∆ = 0 , equation (7.26) has one real root with multiplicity 2 , and the correspondent
solution to (7.25) is:
b
y1 = e − 2 a x
.
In this situation, we need a second independent solution, that is obtained from formula (7.22)
of Theorem 7.5, using the just found y1 and with p(s) built for equation (7.26), thus:
Z
b
− dx
Z a
b e b
y2 = e− 2 a x b
dx = x e− 2 a x .
e− a x
In other words, when ∆ = 0 the general solution of (7.25) is:
b
y = e− 2 a x
c1 + c2 x . (7.29)
Note that the knowledge of the Wronskian expression is useful in the study of non–homogeneous
differential equations too, as it will be shown in § 7.2.
and the general solution can be expressed as y(x) = c1 y1 (x) + c2 y2 (x) . Now, forming the
initial conditions: (
y(0) = c1 y1 (0) + c2 y2 (0) = c1 ,
√
y 0 (0) = c1 y10 (0) + c2 y20 (0) = c1 + c2 5 ,
we see that constants c1 and c2 must verify:
c1 = 0 ,
(
c1 = 0 ,
√ =⇒ 1
c1 + c2 5 = 1 , c2 = √ .
5
In conclusion, the considered initial value problem is solved by:
ex √
y(x) = √ sin 5x .
5
When ∆ < 0 , there are two complex conjugate roots m1,2 = α ± i β of (7.31); then, setting for
example y1 = xα cos(β ln x) and y2 = xα sin(β ln x) , the Wronskian does not vanish:
y1 y2
W (y1 , y2 )(x) = det 0 = β x2α−1 6= 0 , for x > 0 .
y1 y20
When ∆ = 0 , equation (7.31) has one real root m of multiplicity 2 ; in this case y1 = xm and
y2 = y1 ln x ; again, the Wronskian does not vanish:
y1 y2
W (y1 , y2 )(x) = det 0 = x2 m−1 6= 0 , for x > 0 .
y1 y20
Notice, again, that knowing the Wronskian turns out useful also when studying the non–
homogeneous differential equations case (refer to § 7.2).
7.1. HOMOGENEOUS EQUATIONS 119
m (m − 1) − 2 m + 2 = 0 ,
m (m − 1) − m + 5 = 0 ,
with:
1 0 1
I(x) = q(x) − p (x) − p2 (x) . (7.33)
2 4
In fact, assuming the knowledge of a solution to (7.4a) of the form y = f u leads to, in the
same way followed to obtain equation (7.6):
Function f (x) does not vanish, and has first and second derivatives given by:
fp f (p2 − 2 p0 )
f0 = − , f 00 = .
2 4
Hence, f can be simplified out in (7.34), yielding the reduced form:
00 1 0 1 2
u + q− p − p u = 0, (7.35)
2 4
which is stated in (7.32)–(7.33). Equation (7.32) is called the normal form of equation (7.4a).
Function I(x) , introduced in (7.33), is called invariant of the homogeneous differential equation
(7.4a) and represents a mathematical invariant, in the sense expressed by the following Theorem
7.13.
L1 y = y 00 + p1 y 0 + q1 y = 0 (7.36)
L2 y = y 00 + p2 y 0 + q2 y = 0 , (7.37)
by the the change of dependent variable y = f u , then the invariants of (7.36)–(7.37) coincide:
1 0 1 1 1
I1 = q1 − p1 − p21 = q2 − p02 − p22 = I2 .
2 4 2 4
Viceversa, when equations (7.36) and (7.37) admit the same invariant, either equation can be
transformed into the other one, by:
Z
1
−
2 p1 (x) − p2 (x) dx
y(x) = u(x) e .
Remark 7.14. We can transform any second–order linear differential equation into its normal
form. Moreover, if we are able to solve the equation in normal form, then we can obtain, easily,
the general solution to the original equation. The next Example 7.15 clarifies this idea.
Example 7.15. Consider the homogeneous differential equation, depending on the real positive
parameter a :
00 2 0 2 2
y − y + a + 2 y = 0. (7.38)
x a
Hence:
2 2 1 0 1
p(x) = − , q(x) = a2 + =⇒ I(x) = q(x) − p (x) − p2 (x) = a2 .
x a2 2 4
In this example, the invariant is not dependent of x , and the normal form is:
u00 + a2 u = 0 . (7.39)
The general solution to (7.39) is u(x) = c1 cos(a x)+c2 sin(a x) , where c1 , c2 are real parameter,
and the solution to the original equation (7.38) is:
Z
1
−
2 p(x) dx
y(x) = u(x) e = c1 x cos(a x) + c2 x sin(a x) .
7.1. HOMOGENEOUS EQUATIONS 121
Example 7.16. Find the general solution of the homogeneous linear differential equation of
second order:
y 00 − 2 tan(x) y 0 + y = 0 .
The first step consists in transforming the given equation into normal form, with the change of
variable: Z Z
1
−
2 p(x) dx tan x dx
u
y=u e =u e = .
cos x
The normal form is:
u00 + 2 u = 0 ,
The normal form clarifies the structure of the solutions to a constant–coefficient, linear equation
of second–order.
Remark 7.17. Consider the constant–coefficient equation (7.25). The change of variable:
b
y = u e− 2 a x
b2 − 4 a c
u00 − u = 0, (7.40)
4 a2
since, in this case:
b c 1 0 1 2 −b2 + 4 a c
p=− , q=− =⇒ I =q− p − p = .
a a 2 4 4 a2
The normal form (7.40) explains the nature of the following formulæ (7.41), (7.42) and (7.43),
namely describing the structure of the solution to a constant–coefficient homogeneous linear
differential equation of second order. In the following, the discriminant is ∆ = b2 − 4 a c and
c1 , c2 are constant:
(i) if ∆ > 0 ,
√ √
b
−2a x ∆ ∆
y(x) = e c1 cosh x + c2 sinh x ; (7.41)
2a 2a
(ii) if ∆ < 0 ,
√ √
b
−2a x −∆ −∆
y(x) = e c1 cos x + c2 sin x ; (7.42)
2a 2a
(iii) if ∆ = 0 ,
b
y(x) = e− 2 a x (c1 x + c2 ) . (7.43)
122 CHAPTER 7. LINEAR EQUATIONS OF SECOND ORDER
L (y − yp ) = L y − L yp = r − r = 0 .
This means that y − yp can be express by a linear combination of y1 and y2 . Hence, thesis
(7.45) is proved.
Theorem 7.19. Let y1 and y2 be two independent solutions of the homogeneous equation
associated to (7.44). Then, a particular solution to (7.44) has the form:
where: Z Z
y2 (x) r(x) y1 (x) r(x)
k1 (x) = − dx , k2 (x) = dx . (7.47)
W (y1 , y2 )(x) W (y1 , y2 )(x)
Proof. Assume that y1 and y2 are independent solutions of the homogeneous equation associ-
ated to (7.44), and look for a particular solution of (7.44) in the desired form:
yp = k1 y1 + k2 y2 ,
where k1 , k2 are two C 1 functions to be determined. Computing the first derivative of yp yields:
Let us impose a first condition on y1 and y2 , i.e., impose that they verify:
At this point, imposing that yp solves equation (7.44) leads to forming the following expression
(in which variable x is discarded, to ease the notation):
yp00 + p yp0 + q yp = k1 y100 + k2 y200 + k10 y10 + k20 y20 + p k1 y10 + k2 y20 + q k1 y10 + k2 y20
Equations (7.49) and (7.50) form a 2 × 2 linear system, in the variables k10 , k20 , that admits a
unique solution:
y2 r y1 r
k10 = − , k20 = , (7.51)
W (y1 , y2 ) W (y1 , y2 )
since its coefficient matrix is the Wronskian W (y1 , y2 ) , which does not vanish, given the as-
sumption that y1 and y2 are independent. Thesis (7.47) follows by integration of k10 , k20 in
(7.51).
Example 7.20. In Example 7.4, we showed that the general solution of the homogeneous
equation (7.18) has the form (7.19), i.e., c1 x + c2 x2 ex , c1 , c2 ∈ R . Here, we use Theorem 7.19
to find the general solution of the non–homogeneous equation:
2
x2 (1 + x) y 00 − x (2 + 4 x + x2 ) y 0 + (2 + 4 x + x2 ) y = x2 (1 + x) . (7.52)
x (2 + 4 x + x2 ) 0 (2 + 4 x + x2 )
y 00 − y + y = x2 (1 + x) .
x2 (1 + x) x2 (1 + x)
x3
Z Z
k1 (x) = − x2 dx = − , k2 (x) = x e−x dx = −(1 + x) e−x .
3
Hence, the general solution of (7.52) is:
x4
y(x) = c1 x + c2 x2 ex − − x2 (1 + x) .
3
Example 7.21. Consider the initial value problem:
(
x2 y 00 − x y 0 + 5 y = x2 ,
y(1) = y 0 (1) = 0 ,
In Example 7.12, the associated homogeneous equation was considered, of which two independent
solutions were found, namely y1 = x cos(2 ln x) and y2 = x sin(2 ln x) , whose Wronskian is 2 x .
Now, writing the given non–homogenous differential equation in explicit form:
1 0 5
y 00 − y + 2 y = 1,
x x
124 CHAPTER 7. LINEAR EQUATIONS OF SECOND ORDER
(a) If K(α) 6= 0 , it means that α is not a root of the characteristic equation. Then, a
particular solution of (7.53) has the form:
yp = eα x Qn (x) , (7.54)
yp = xs eα x Rn (x) , (7.55)
K(λ) = λ2 + 2 λ + 1 = 0 ,
thus, we are in situation (1a) and we look for a solution of the form (7.54), that is
yp (x) = s0 x2 + s1 x + s2 .
Differentiating: (
yp0 (x) = 2 s0 x + s1 ,
yp00 (x) = 2 s0 ,
and imposing that yp solves the differential equation, we obtain:
s0 x2 + (4 s0 + s1 ) x + 2 s0 + 2 s1 + s2 = x2 + x .
yp = x 2 − 3 x + 4 .
Finally, solving the associated homogeneous equation, we obtain the required general solution:
Observe, first, that the general solution of the associated homogeneous equation is:
Observe, further, that the characteristic equation has roots ±i , Hence, we are in situation (2b)
and we look for a solution of the form (7.57), that is:
a x2 y 00 + b x y 0 + c y = 0 (7.30)
To solve it, in a different way, we use a change of independent variable from x to ξ. This means
that we take a C 2 bijective function φ : R → R+ which realizes the variable transformation:
x = eξ ⇐⇒ ξ = ln x (7.59)
or in other terms
z(ln x) = y(x) (7.60)
We now impose that function y(x) introduced in (7.60) is indeed a solution to (7.30); with this
aim in mind we compute the first and second derivative in (7.60) , using (of course) the chain
rule:
1 0
y 0 (x) = z (ln x) (7.61)
x
1 1
y 00 (x) = − 2 z 0 (ln x) + 2 z 00 (ln x) (7.62)
x x
7.3. CHANGE OF INDEPENDENT VARIABLE 127
7.26. Keeping in mind how we proceed in examples 7.24 and 7.25 we face the general situation.
So consider once more equation
with the general change of independent variable that now we represent with
Inserting (7.71) we obtain the transformation of (7.4) (which we put in explicit form)
Notice that coefficients of the new linear differential operator in (7.72) are expressed as functions
of the (old) variable x. To complete the variabile transformation they have to be converted to
the new ξ variable using x = ϕ−1 (ξ).
In both examples 7.24 and 7.25 the transformed equation (7.72) turned out to be a constant
coefficient equation. Of course this nice situation does not come out of the blu, but is the
consequence of a very specific condition that depends on the coefficients p(x) and q(x) of the
original equation. In general, it is not possibile to transform a variable coefficient equation into
a constant coefficient one using only a variable tranformation involving solely the independent
variable.
For brevity we denote the coefficients of the transformed differential operator (7.72) as
To establish conditions under which (7.72) becomes an equation with constant coefficients, it is
necessary at least that qb to be a constant function, hence we impose the following condition
Moreover form (7.73) we have to assume q(x) ≥ 0 also. Thus, after this choice we see that (7.72)
can become a constant coefficient equation if and only if pb also is a constant function. Therefore,
inserting (7.73) in the expression defining pb we see that the expression
1 2p(x)q(x) + q 0 (x)
2c [q(x)]3/2
must also individuate a constant function. Summing up we can state the following result.
2p(x)q(x) + q 0 (x)
(7.74)
[q(x)]3/2
7.4. HARLEY DIFFERENTIAL RESOLVENT 129
transforms equation (7.4) to an equation with constant coefficients. If the expression (7.74) is
not constant, no change of idependent variable will reduce (7.4) to an equation with constant
coefficients.
we have
1 q 0 (x) + 2p(x)q(x) 8
p(x) = 8x − , q(x) = 20x2 =⇒ 3/2
=√
x [q(x)] 5
Hence form theorem 7.27 we know that the change of independent variable
Z x√
ϕ(x) = c 20s2 ds
0
transforms the given equation into an equation with constant coeffcients. Of course we can
choose√c ∈ R in order to keep the transformation as simple as possible. In this case we take
c = 2/ 20 so that we can use ϕ(x) = x2 , x ≥ 0 as a change of independent variable, which gives
the constant coeffcient equation
z 00 + 4z 0 + 5z = 0.
Solution of this equation is
z(ξ) = e−2ξ (c1 cos ξ + c2 sin ξ)
and since ξ = ϕ(x) = x2 the general solution of the given equation is
2
y(x) = e−2x c1 cos x2 + c2 sin x2 .
y 3 − 3y + 2x = 0 (7.76)
in the unknown y being x a parameter. Equation (7.76) individuates an implicit function y = y(x)
which will be determinated solving a differential equation. Using implicit differentiation we see
that this function satisfies
2 1
y0 = − (7.77)
3 y2 − 1
Now we “solve” for y 2 in (7.76) obtaining
2x 2(y − x)
y2 = 3 − =⇒ y 2 − 1 =
y y
130 CHAPTER 7. LINEAR EQUATIONS OF SECOND ORDER
1 y y 2 − x2 − 2(y − x)
y 1 (1 − x2 )y y 2 + xy − 2
= 2
= 2
= (7.81)
y−x 1−x y−x 1−x y−x 1 − x2
Inserting (7.81) in the differential equation (7.78) we obtain
y 2 + xy − 2
y0 = − (7.82)
3(1 − x2 )
Now we try to use equality (7.81): first we multiply both sides of (7.84) for 3(1 − x2 ) so that:
9 1 − x2 y 00 − 9xy 0 + y = 0
(7.86)
7.4. HARLEY DIFFERENTIAL RESOLVENT 131
x 1 q 0 (x) + 2p(x)q(x)
p(x) = − , q(x) = =⇒ =0
1 − x2 9(1 − x2 ) [q(x)]3/2
Thus we can change the indpendent variable obtaining a constant coefficient equation using
Z x
c 1 c
ξ = ϕ(x) = √ ds = arcsin x
3 0 1−s 2 3
1
z 00 + z = 0 =⇒ z = c1 cos(ξ/3) + c2 sin(ξ/3)
9
Going back to the orginal variable we have the general solution of (7.86)
1 1
y(x) = c1 cos arcsin x + c2 sin arcsin x (7.87)
3 3
√
1 1 1
y0 (x) = 2 sin arcsin x , y1,2 (x) = ± 3 cos arcsin x − sin arcsin x
3 3 3
-1
-2
-2 -1 0 1 2
7.5 Exercises
1. Solve the following second order variable coefficient linear differential equations:
00 2 2
(a) y − 1+ y0 + y = 0 ;
x x
2x 2
(b) y 00 − y0 + y = 0;
1 + x2 1 + x2
1
(c) y 00 − y 0 − 4 x2 y = 0 .
x
m
Hint: y = ex .
2. Solve the following initial value problems, using the transformation of the given differential
equation in normal form:
00
0 2
y + 2 sin x y + (sin x) + cos x − 6
y = 0,
x2
(a)
y(1) = 0 ,
y 0 (1) = 1 ,
2 0
00
y + x y + y = 0 ,
(b) y(1) = 1 ,
0
y (1) = 0 .
5. Finding the appropriate change of variable following theorem 7.27 find the general solution
of the following differential equations
3 0 1
(a) y 00 + y + 6y=0
x x
(b) xy 00 + (4x2 − 1)y 0 + 4x3 y = 2x3
8 Series solutions
Up to this point we illustrate various techniques for solving second-order linear equations. How-
ever, several equations are not tractable using the methods we have presented: for instance
simple equation like y 00 = xy defies all the methods of solution. However for a large class of
equations it is still possible to express solution in terms of powers series. At this purpose we
refer to the concept exposed in section 2.5. Thus in what follows we will work with analytic
functions that is f : I → R, f ∈ C ∞ being I a real open interval such that for x0 ∈ I the Taylor
series
∞
X (x − x0 )n
f (n) (x0 )
n!
n=0
converges in some neighborhood of x0 . Most elementary functions like sin x, cos x and ex are an-
alytic ewerywhere, moreover sums, differences and products of these functions are too. Quotient
of analytic functions are analytic at all points where the denominator does not vanish.
where a0 and a1 are arbitrary constants and y1 and y2 are independent functions that are analytic
at x = 0.
To evaluate the coefficients an as indacated in theorem 8.2 we indicate a five steps procedure,
known as power series method.
Step 1. Evaluate the derivatives of the power series
∞
X
y= an xn (8.1)
n=0
X∞
y0 = nan xn−1 (8.2)
n=0
X∞
y 00 = n(n − 1)an xn−2 (8.3)
n=0
133
134 CHAPTER 8. SERIES SOLUTIONS
Step 2. Collect powers of x and set the coefficient of each power equal to zero.
Step 3. The equation obtained by setting the coefficients of xn to zero in Step 2. will contain
aj terms for a finite number of j’s. Solve the equation for the aj term having the largest index.
The resulting equation is known as the recurrence formula for the given differential equation.
Step 4. Use the recurrence to determine aj , j = 2, 3, . . . in terms od a0 and a1 .
Step 5. Insert the coefficients determinated in Step 4. into (8.1) to express the solution of the
differential equation Ly = 0.
We illustrate these five steps with some working examples. The fist one is very simple in
terms of computation, but it completely represents the procedure just described
Example 8.3. Consider, the example is inspired by [34] section 107, the (constant coefficient)
differential equation
y 00 + 4y = 0
Step 1.
∞
X ∞
X
y 00 + 4y = n(n − 1)an xn−2 + 4 an xn = 0 (S1)
n=0 n=0
Step 2. Change the indexes of the terms in the second series at right hand side of (S1) so that
the series will involve the power xn−2 in the general term, thus (S1) becomes
∞
X ∞
X
y 00 + 4y = n(n − 1)an xn−2 + 4 an−2 xn−2 = 0 (S2)
n=0 n=2
summation starts from n = 2 because the first two terms in the first series in (S2) are both zero.
Step 4. Since condition
n(n − 1)an + 4an−2
holds true, solving for an we obtain, for n ≥ 2
4
an = − an−2
n(n − 1)
thus a0 and a1 are free and arbitrary. So to start the recursion fix a0 , a1 ∈ R so that
4 4
a2 = − a0 a3 = − a1
2·1 3·2
4 4
a4 = − a2 a5 = − a3
4·3 5·4
4 4
a6 = − a4 a7 = − a5
6·5 7·6
in general
4 4
a2k = − a2k−2 a2k+1 = − a2k−1
2k · (2k − 1) (2k + 1) · 2k
Multiplying side by side the relations of even index, we obtain
4k
a2 · a4 · · · · · a2k = (−1)k a0 · a2 · · · · · a2k−2
(2k)!
which simplifies to
4k
a2k = (−1)k a0 , k ≥ 1
(2k)!
8.1. SOLUTION AT ORDINARY POINTS 135
4k
a2k+1 = (−1)k a1 , k ≥ 1
(2k + 1)!
Step 5. Now recalling the assumed series (8.1) for y we rewrite it in the form
∞
X ∞
X
y =a0 + a2k x2k + a1 x + a2k+1 x2k+1
k=1 k=1
∞ ∞
! !
X (−1)k 4k X (−1)k 4k 2k+1
=a0 1+ x2k + a1 x+ x
(2k)! (2k + 1)!
k=1 k=1
∞ ∞
X (−1)k 2k a1 X (−1)k
=a0 (2x) + (2x)2k+1
(2k)! 2 (2k + 1)!
k=0 k=0
=a0 cos(2x) + a1 sin(2x)
Example 8.3 is interesting in order to understand how the method of searching for solutions
works, the fact that, for the particular equation considered, the solution can be easily obtained
with other methods, it allows to check the validity of this new approach. Now let’s see how to
use the power series method to solve an equation with variable coefficients,
(1 − x2 )y 00 − 6xy 0 − 4y = 0
The first difference with the previous example lies in the fact the coefficients are analytical if
|x| < 1, so here power series solution (8.1) is considered for |x| < 1. Inserting (8.1), (8.2) and
(8.3) into the equation we obtain
∞
X ∞
X ∞
X ∞
X
n(n − 1)an xn−2 − n(n − 1)an xn − 6nan xn − 4an xn = 0
n=0 n=0 n=0 n=0
or
∞
X ∞
X
n(n − 1)an xn−2 − (n + 1)(n + 4)an xn = 0
n=0 n=0
n = 0 =⇒ 0 · a0 = 0, n = 1 =⇒ 0 · a1 = 0.
136 CHAPTER 8. SERIES SOLUTIONS
follows that
∞ ∞
! !
X
2k
X 2k + 3 2k+1
y = a0 1+ (k + 1)x + a1 x+ x
3
n=1 n=1
It is possible to verify with the usual convergence tests that both series converge for |x| < 1.
Observe that in this particular example both power series are summed in terms of elementary
functions, in fact
∞ ∞
X
2k
X 2k + 3 a0 a1 3x − x3
y = a0 (k + 1)x + a1 x2k+1 = +
3 (1 − x2 )2 3 (1 − x2 )2
n=0 n=0
y 00 − xy = 0 (8.4)
However, the process for the solution in power series results more complicated with respect to
the previous examples 8.3 and 8.4. Let’s start the usual power series procedure inserting (8.1)
and (8.3) into the Airy equation (8.4)
∞
X ∞
X
n−2
n(n − 1)an x − an xn+1 = 0 (8.4a)
n=0 n=0
8.2. AIRY DIFFERENTIAL EQUATION 137
From (8.4a) see immediately infer that we must have a2 = 0 and then we can rewrite (8.4a) as
∞
X ∞
X
n(n − 1)an xn−2 − an xn+1 = 0 (8.4b)
n=3 n=0
Thus the sequence (an ) of the coefficients of the power series solution of equation (8.4) verifies
condition
1
an+3 = an
(n + 3)(n + 2) (8.5)
a0 , a1 , a2 ∈ R, with a2 = 0.
Equation (8.5) identifies a recurrence of the third order, which means that three indexed families
a3n , a3n+1 , a3n+2 are generated by (8.5) starting from a0 , a1 , a2 . Since we saw that a2 = 0 we
see that a3n+2 = 0 for any n ∈ N, while a3n and a3n+1 are determinated as follows:
1 1
a3 = a0 a4 = a1
6·5 7·6
1 1
a6 = a3 a7 = a4
9·8 10 · 9
1 1
a9 = a6 a10 = a7
12 · 11 13 · 12
...
In general
1 1
a3n = a3n−3 , a3n+1 = a3n−3
3n(3n − 1) (3n + 1)3n
Thus, multiplying, as we did in examples 8.3 and 8.4, by columns, we obtain, for n ≥ 1:
a0 a1
a3n = n , a3n+1 = n (8.6)
Y Y
3k(3k − 1) (3k + 1)3k
k=1 k=1
Both products in (8.6) are computed in closed form using Euler Gamma function, in fact, first
write product which define a3n as:
n n
Y
2n
Y 1
3k(3k − 1) = 3 n! k−
3
k=1 k=1
so that
n 2
32n n! Γ n +
Y
3
3k(3k − 1) = 2
(8.7)
k=1
Γ 3
138 CHAPTER 8. SERIES SOLUTIONS
Definition 8.5. The singular point x0 is a regular singular point of equation Ly = 0 if both
functions (x − x0 )p(x) and (x − x0 )2 q(x) are analytic at x0 . Singular points that are not regular
are called irregular.
Assuming that x0 = 0 is a regular singular point of Ly = 0. The power series method has to
be modified following the so called Frobenius method. Here we limit to a brief mention, see for
instance [8] or [34] for an extensive treatment. In presence of regular singular points we search
power series solutions of the form
∞
X
y = xα an xn (8.10)
n=0
being α a real constant, which will be determinated by a quadratic equation, called indicial
equation.
We illustrate the procedure with a quite popular example, since is exposed also in [34] and
in [11].
x0 = 0 is a regular singular point for (8.11). We search for a solution of the form
∞
X
y= an xn+α (8.12)
n=0
∞
X
00
y = (n + α)(n + α − 1)an xn+α−2 (8.14)
n=0
To individuate α we equate the coeffcient of the lowest power of x in (8.15) to zero, here xα−1 ,
obtaining in this way the indicial equation
2α(α − 1) a0 + α a0 = 0 (8.16)
Solving (8.16) we obtain α = 0 and α = 21 ; notice, in view to the analysis of the general case,
that the difference beteween the two roots of the indicial equation is not an integer. To each
values of α corresponds a solution. Equating to zero the powers of xn+α in (8.15) we obtain the
recurrence rule for the coefficient of the power series solution to (8.11):
3−n
an = an−1
n(2n − 1)
which gives
n=1 a1 = 2a0
1 1
n=2 a2 = a1 = a0
6 3
n≥2 an = 0
If α = 21 (8.17) gives
(2n − 5)
an = − an−1
2n(2n + 1)
which, for the first values of n gives
3
n=1→ a1 = a0
2·1·3
1
n=2→ a2 = a1
2·2·5
−1
n=3→ a3 = a2
2·3·7
−3
n=4→ a4 = a3
2·4·9
−5
n=5→ a5 = a4
2 · 5 · 11
−7
n=6→ a6 = a5
2 · 6 · 13
140 CHAPTER 8. SERIES SOLUTIONS
(2n − 5)
an = − an−1
2n(2n + 1)
(−1)n 3a0
an =
2n n!(2n − 3)(2n − 1)(2n + 1)
x2 y 00 + xy 0 + (x2 − ν 2 )y = 0 (8.18)
Solutions can be found with the Frobenius method, which consists in finding functions of the
form
∞
X
y= an xn+α , a0 6= 0
n=0
∞
X
y0 = (n + α)an xn+α−1
n=0
X∞
y 00 = (n + α)(n + α − 1)an xn+α−2
n=0
going on
∞
X ∞
X
2 n+α
(n + α)(n + α − 1) + (α + n) − ν an x + an xn+α+2 = 0
n=0 n=0
8.4. BESSEL EQUATION 141
and finally
∞
X ∞
X
(α + n)2 − ν 2 an xn+α + an−2 xn+α = 0
n=0 n=2
Every coefficient of the powers of x has to be zero. Thus, we have the following equations:
α 2 − ν2 a = 0
0
(α + 1)2 − ν 2 a1 = 0
(α + n)2 − ν 2 an + an−2 = 0
a0
a2 = −
22 · 1 · (1 + ν)
a2
a4 = − 2
2 · 2 · (2 + ν)
a4
a6 = − 2
2 · 3 · (3 + ν)
a6
a8 = − 2
2 · 4 · (4 + ν)
......
(−1)n a0 (−1)n a0
a2n = n = a2n =
22n Y 22n n! (1 + ν)n
n! (k + ν)
k=1
Now since the solution of BE has, for the previsous argument the structure
∞ ∞
X X (−1)n Γ(1 + ν)
y= a2n x2n+ν = a0 x2n+ν
22n n! Γ(1 + ν + n)
n=0 n=0
the Bessel function, Jν (x) of the first kind of order ν and argument x is obtained choosing
1
a0 =
2ν Γ(1 + ν)
This time, we need to impose that 2ν is not a positive integer. Using the ratio test it is possible
to show that the series defining Bessel functions of the first kind of order ν and −ν are absolutely
convergent for all x ∈ R (as a matter of fact also in C).
Jν and J−ν are linearly independent: they form a basis for the vector space of the solutions
of (8.18)
Bibliography
[1] J.L. Allen and F.M. Stein. On solution of certain Riccati differential equations. Amer.
Math. Monthly, 71:1113–1115, 1964.
[2] Tom Mike Apostol. Calculus: Multi Variable Calculus and Linear Algebra, with Applications
to Differential Equations and Probability. John Wiley & Sons, New York, 1969.
[3] Daniel J. Arrigo. Symmetry analysis of differential equations: an introduction. John Wiley
& Sons, New York, 2015.
[4] E. Artin and M. Butler. The gamma function. Holt, Rinehart and Winston, New York,
1964.
[5] Daniel Bernoulli. Exercitationes quaedam mathematicæ. Apud Dominicum Lovisam, Venice,
1724.
[6] D. Brannan. A First Course in Mathematical Analysis. Cambridge University Press, Cam-
bridgde, 2006.
∞
X 1 π2
[7] B.R. Choe. An elementary proof of = . American Mathematical Monthly,
n2 6
n=1
94(7):662–663, 1987.
[8] E.A. Coddington and N. Levinson. Theory of ordinary differential equations. Tata McGraw-
Hill Education, 1955.
[9] Lothar Collatz. Differential Equations: An Introduction With Applications. John Wiley
and Sons, 1986.
[10] P. Duren. Invitation to Classical Analysis, volume 17. American Mathematical Society,
Providence, 2012.
[11] Refaat El Attar. Ordinary Differential Equations. Lulu Press Incorporated, 2006.
[13] O.J. Farrell and B. Ross. Solved problems: gamma and beta functions, Legendre polynomials,
Bessel functions. Macmillan, New York, 1963.
[14] Angelo Favini, Ermanno Lanconelli, Enrico Obrecht, and Cesare Parenti. Esercizi di analisi
matematica: Equazioni Differenziali, volume 2. Clueb, 1978.
[16] A. Ghizzetti, A. Ossicini, and L. Marchetti. Lezioni di complementi di matematica. 2nd ed.
Libreria eredi Virgilo Veschi, Roma, 1972.
143
144 BIBLIOGRAPHY
[17] Robert Harley. On the theory of the transcendental solution of algebraic equations. Quar-
terly Journal of Pure and Applied Mathematics, 5:337–361, 1862.
[18] Robert Harley. On differential resolvents. Proceedings of the London Mathematical Society,
1(1):35–42, 1865.
[19] Phillip Hartman. Ordinary Differential Equations 2nd Edition. Siam, 2002.
[20] O. Hijab. Introduction to calculus and classical analysis. Springer, New York, 2011.
[21] P. Hydon. Symmetry methods for differential equations: a beginner’s guide. Cambridge
University Press, Cambridge, 2000.
[25] A.M. Legendre. Traité des fonctions elliptiques et des intégrales Euleriennes, volume 2.
Huzard-Courcier, Paris, 1826.
[26] D.H. Lehmer. Interesting series involving the central binomial coefficient. The American
Mathematical Monthly, 92(7):449–457, 1985.
[27] Joseph Liouville. Remarques nouvelles sur l’equation de Riccati. Journal de mathématiques
pures et appliquées, 6:1–13, 1841.
[28] A.R. Magid. Lectures on differential galois theory. Notices of the American Mathematical
Society, 7:1041–1049, 1994.
[29] C.C. Maican. Integral Evaluations Using the Gamma and Beta Functions and Elliptic
Integrals in Engineering: A Self-study Approach. International Press of Boston Inc., Boston,
2005.
[30] A.M. Mathai and H.J. Haubold. Special Functions for Applied Scientists. Springer, New
York, 2008.
[32] Adam Parker. Who solved the bernoulli equation and how did they do it? College Mathe-
matical Journal, 44:89–97, 2013.
[33] Earl David Rainville. Intermediate differential equations. Macmillan, New York, 1964.
[34] Earl David Rainville and Phillip E. Bedient. Elementary differential equations. 6th ed.
Macmillan, New York, 1981.
[35] P.R.P. Rao. The Riccati differential equation. Amer. Math. Monthly, 69:995–996, 1962.
[36] P.R.P. Rao and V.H. Hukidave. Some separable forms of the Riccati equation. Amer. Math.
Monthly, 75:38–39, 1968.
[37] L. Schwartz. Mathematics for the physical sciences. Addison-Wesley, New York, 1966.
[38] H. Siller. On the separability of the Riccati differential equation. Math. Mag., 43:197–202,
1970.
[40] J.S.W. Wong. On solution of certain Riccati differential equations. Math. Mag., 39:141–143,
1966.
[41] R.C. Wrede and M.R. Spiegel. Schaum’s Outline of Advanced Calculus. McGraw-Hill, New
York, 2010.