1.8: Improper Functions (Exercises)
( \newcommand{\kernel}{\mathrm{null}\,}\)
[exer:1] Suppose g and h are differentiable on [a,b], with
a≤g(y)≤b\quad and\quada≤h(y)≤b,c≤y≤d.
Let f and fy be continuous on [a,b]×[c,d]. Derive Liebniz’s rule:
ddy∫h(y)g(y)f(x,y)dx=f(h(y),y)h′(y)−f(g(y),y)g′(y)+∫h(y)g(y)fy(x,y)dx.
(Hint: Define H(y,u,v)=∫vuf(x,y)dx and use the chain rule.)
[exer:2] Adapt the proof of Theorem [theorem:2] to prove Theorem [theorem:3].
[exer:3] Adapt the proof of Theorem [theorem:4] to prove Theorem [theorem:5].
[exer:4] Show that Definition [definition:3] is independent of c; that is, if ∫caf(x,y)dx and ∫bcf(x,y)dx both converge uniformly on S for some c∈(a,b), then they both converge uniformly on S and every c∈(a,b).
[exer:5]
()0pt0pt14pt 8pt22pt0pt
Show that if f is bounded on [a,b]×[c,d] and ∫baf(x,y)dx exists as a proper integral for each y∈[c,d], then it converges uniformly on [c,d] according to all of Definition [definition:1]–[definition:3].
Give an example to show that the boundedness of f is essential in (a).
[exer:6] Working directly from Definition [definition:1], discuss uniform convergence of the following integrals:
(a) ∫∞0dx1+y2x2dx | (b) ∫∞0e−xyx2dx |
(c) ∫∞0x2ne−yx2dx | (d) ∫∞0sinxy2dx |
(e) ∫∞0(3y2−2xy)e−y2xdx | (f) ∫∞0(2xy−y2x2)e−xydx |
[exer:7] Adapt the proof of Theorem [theorem:6] to prove Theorem [theorem:7].
[exer:8] Use Weierstrass’s test to show that the integral converges uniformly on S:
()0pt0pt14pt 8pt22pt0pt
∫∞0e−xysinxdx, S=[ρ,∞),ρ>0
∫∞0sinxxydx, S=[c,d], 1<c<d<2
∫∞1e−pxsinxyxdx, p>0, S=(−∞,∞)
∫10exy(1−x)ydx, S=(−∞,b),b<1
∫∞−∞cosxy1+x2y2dx, S=(−∞,−ρ]∪[ρ,∞),ρ>0.
∫∞1e−x/ydx, S=[ρ,∞),ρ>0
∫∞−∞exye−x2dx, S=[−ρ,ρ],ρ>0
∫∞0cosxy−cosaxx2dx, S=(−∞,∞)
∫∞0x2ne−yx2dx, S=[ρ,∞),ρ>0, n=0, 1, 2,…
[exer:9]
()0pt0pt14pt 8pt22pt0pt
Show that
Γ(y)=∫∞0xy−1e−xdx
converges if y>0, and uniformly on [c,d] if 0<c<d<∞.
Use integration by parts to show that
Γ(y)=Γ(y+1)y,y≥0,
and then show by induction that
Γ(y)=Γ(y+n)y(y+1)⋯(y+n−1),y>0,n=1,2,3,….
How can this be used to define Γ(y) in a natural way for all y≠0, −1, −2, …? (This function is called the gamma function.)
Show that Γ(n+1)=n! if n is a positive integer.
Show that
∫∞0e−sttαdt=s−α−1Γ(α+1),α>−1,s>0.
[exer:10] Show that Theorem [theorem:8] remains valid with assumption (c) replaced by the assumption that |gx(x,y)| is monotonic with respect to x for all y∈S.
[exer:11] Adapt the proof of Theorem [theorem:8] to prove Theorem [theorem:9].
[exer:12] Use Dirichlet’s test to show that the following integrals converge uniformly on S=[ρ,∞) if ρ>0:
(a) ∫∞1sinxyxydx | (b) ∫∞2sinxylogxdx |
(c) ∫∞0cosxyx+y2dx | (d) ∫∞1sinxy1+xydx |
[exer:13] Suppose g, gx and h are continuous on [a,b)×S, and denote H(x,y)=∫xah(u,y)du, a≤x<b. Suppose also that
limx→b−{supy∈S|g(x,y)H(x,y)|}=0\quad and \quad∫bagx(x,y)H(x,y)dx
converges uniformly on S. Show that ∫bag(x,y)h(x,y)dx converges uniformly on S.
[exer:14] Prove Theorem [theorem:10] for the case where f=f(x,y) is continuous on (a,b]×[c,d].
[exer:15] Prove Theorem [theorem:11] for the case where f=f(x,y) is continuous on (a,b]×[c,d].
[exer:16] Show that
C(y)=∫∞−∞f(x)cosxydx\quad and\quadS(y)=∫∞−∞f(x)sinxydx
are continuous on (−∞,∞) if
∫∞−∞|f(x)|dx<∞.
[exer:17] Suppose f is continuously differentiable on [a,∞), limx→∞f(x)=0, and
∫∞a|f′(x)|dx<∞.
Show that the functions
C(y)=∫∞af(x)cosxydx\quad and\quadS(y)=∫∞af(x)sinxydx
are continuous for all y≠0. Give an example showing that they need not be continuous at y=0.
[exer:18] Evaluate F(y) and use Theorem [theorem:11] to evaluate I:
()0pt0pt14pt 8pt22pt0pt
F(y)=∫∞0dx1+y2x2, y≠0; I=∫∞0tan−1ax−tan−1bxxdx, a, b>0
F(y)=∫∞0xydx, y>−1; I=∫∞0xa−xblogxdx, a, b>−1
F(y)=∫∞0e−xycosxdx, y>0
I=∫∞0e−ax−e−bxxcosxdx, a, b>0
F(y)=∫∞0e−xysinxdx, y>0
I=∫∞0e−ax−e−bxxsinxdx, a, b>0
F(y)=∫∞0e−xsinxydx; I=∫∞0e−x1−cosaxxdx
F(y)=∫∞0e−xcosxydx; I=∫∞0e−xsinaxxdx
[exer:19] Use Theorem [theorem:11] to evaluate:
()0pt0pt14pt 8pt22pt0pt
∫10(logx)nxydx, y>−1,n=0, 1, 2,….
∫∞0dx(x2+y)n+1dx, y>0, n=0, 1, 2, ….
∫∞0x2n+1e−yx2dx, y>0, n=0, 1, 2,….
∫∞0xyxdx, 0<y<1.
[exer:20]
()0pt0pt14pt 8pt22pt0pt
Use Theorem [theorem:11] and integration by parts to show that
F(y)=∫∞0e−x2cos2xydx
satisfies
F′+2yF=0.
Use part (a) to show that
F(y)=√π2e−y2.
[exer:21] Show that
∫∞0e−x2sin2xydx=e−y2∫y0eu2du.
(Hint: See Exercise [exer:20].)
[exer:22] State a condition implying that
C(y)=∫∞af(x)cosxydx\quad and\quadS(y)=∫∞af(x)sinxydx
are n times differentiable on for all y≠0. (Your condition should imply the hypotheses of Exercise [exer:16].)
[exer:23] Suppose f is continuously differentiable on [a,∞),
∫∞a|(xkf(x))′|dx<∞,0≤k≤n,
and limx→∞xnf(x)=0. Show that if
C(y)=∫∞af(x)cosxydx\quad and\quadS(y)=∫∞af(x)sinxydx,
then
C(k)(y)=∫∞axkf(x)cosxydx\quad and\quadS(k)(y)=∫∞axkf(x)sinxydx,
0≤k≤n.
[exer:24] Differentiating
F(y)=∫∞1cosyxdx
under the integral sign yields
−∫∞11xsinyxdx,
which converges uniformly on any finite interval. (Why?) Does this imply that F is differentiable for all y?
[exer:25] Show that Theorem [theorem:11] and induction imply Eq. [eq:30].
[exer:26] Prove Theorem [theorem:12].
[exer:27] Show that if F(s)=∫∞0e−sxf(x)dx converges for s=s0, then it converges uniformly on [s0,∞). (What’s the difference between this and Theorem [theorem:13]?)
[exer:28] Prove: If f is continuous on [0,∞) and ∫∞0e−s0xf(x)dx converges, then
lims→s0+∫∞0e−sxf(x)dx=∫∞0e−s0xf(x)dx.
(Hint: See the proof of Theorem 4.5.12, p. 273.)
[exer:29] Under the assumptions of Exercise [exer:28], show that
lims→s0+∫∞re−sxf(x)dx=∫∞re−s0xf(x)dx,r>0.
[exer:30] Suppose f is continuous on [0,∞) and
F(s)=∫∞0e−sxf(x)dx
converges for s=s0. Show that lims→∞F(s)=0. (Hint: Integrate by parts.)
[exer:31]
()0pt0pt14pt 8pt22pt0pt
Starting from the result of Exercise [exer:18](d), let b→∞ and invoke Exercise [exer:30] to evaluate
∫∞0e−axsinxxdx,a>0.
Use (a) and Exercise [exer:28] to show that
∫∞0sinxxdx=π2.
[exer:32]
()0pt0pt14pt 8pt22pt0pt
Suppose f is continuously differentiable on [0,∞) and
|f(x)|≤Mes0x,0≤x≤∞.
Show that
G(s)=∫∞0e−sxf′(x)dx
converges uniformly on [s1,∞) if s1>s0. (Hint: Integrate by parts.)
Show from part (a) that
G(s)=∫∞0e−sxxex2sinex2dx
converges uniformly on [ρ,∞) if ρ>0. (Notice that this does not follow from Theorem [theorem:6] or [theorem:8].)
[exer:33] Suppose f is continuous on [0,∞),
limx→0+f(x)x
exists, and
F(s)=∫∞0e−sxf(x)dx
converges for s=s0. Show that
∫∞s0F(u)du=∫∞0e−s0xf(x)xdx.
Answers to selected exercises
If f(x,y)=1/y for y≠0 and f(x,0)=1, then ∫baf(x,y)dx does not converge uniformly on [0,d] for any d>0.
, (d), and (e) converge uniformly on (−∞,ρ]∪[ρ,∞) if ρ>0; (b), (c), and (f) converge uniformly on [ρ,∞) if ρ>0.
Let C(y)=∫∞1cosxyxdx and S(y)=∫∞1sinxyxdx. Then C(0)=∞ and S(0)=0, while S(y)=π/2 if y≠0.
F(y)=π2|y|;I=π2logab F(y)=1y+1;I=loga+1b+1
F(y)=yy2+1; I=12b2+1a2+1
(d) F(y)=1y2+1;I=tan−1b−tan−1a
(e) F(y)=yy2+1;I=12log(1+a2)
(f) F(y)=1y2+1;I=tan−1a
(−1)nn!(y+1)−n−1 \pi2^{-2n-1}\displaystyle{\binom{2n}{n}}y^{-n-1/2}
(c) \displaystyle{\frac{n!}{2y^{n+1}}} (\log y)^{-2} (d) \displaystyle{\frac{1}{(\log x)^{2}}}
\displaystyle{\int_{-\infty}^{\infty}|x^{n}f(x)|\,dx<\infty}
No; the integral defining F diverges for all y.
\displaystyle{\frac{\pi}{2}}-\tan^{-1}a
Beginning of manual
1. If H(y,u,v)=\displaystyle{\int_{u}^{v}f(x,y)\,dx} then
H_{u}(y,u,v)=-f(u,y), \quad H_{v}(y,u,v)=f(v,y),
and, by Theorem 1, H_{y}(u,v,y) =\displaystyle{\int_{u}^{v}f_{y}(x,y)\,dx}. If
F(y)=H(y,g(y),h(y))=\int_{g(y)}^{h(y)}f(x,y)\,dx,
then
\begin{aligned} F'(y)&=&H_{v}(y, g(y),h(y))h'(y)+H_{u}(y,g(y),h(y))g'(y)+ H_{y}(y,g(y),h(y))\\ &=& f(h(y),y)h'(y)-f(g(y),y)g'(y) +\int_{g(y)}^{h(y)} f_{y}(x,y)\,dx.\end{aligned}
2. Theorem 3 (Cauchy Criterion for Convergence of an Improper Integral II) Suppose g is integrable on every finite closed subinterval of (a,b] and denote
G(r)=\int_{r}^{b}g(x)\,dx,\quad a< r\le b.
Then the improper integral \int_{a}^{b}g(x)\,dx converges if and only if, for each \epsilon >0, there is an r_{0}\in(a,b] such that
\tag{A} |G(r)-G(r_{1})|\le\epsilon,\quad a<r,r_{1}\le r_{0}.
For necessity, suppose \int_{a}^{b}g(x)\,dx=L. By definition, this means that for each \epsilon>0 there is an r_{0}\in (a,b] such that
|G(r)-L|<\frac{\epsilon}{2} \text{\quad and\quad} |G(r_{1})-L|<\frac{\epsilon}{2}, \quad a< r,r_{1}\le r_{0}.
Therefore,
\begin{aligned} |G(r)-G(r_{1})|&=&|(G(r)-L)-(G(r_{1})-L)|\\ &\le& |G(r)-L|+|G(r_{1})-L|\le \epsilon,\quad a< r,r_{1}\le r_{0}.\end{aligned}
For sufficiency, (A) implies that
|G(r)|= |G(r_{1})+(G(r)-G(r_{1}))|\le |G(r_{1})|+|G(r)-G(r_{1})|\le |G(r_{1})|+\epsilon,
a< r_{1}\le r_{0}. Since G is also bounded on the compact set [r_{0},b] (Theorem 5.2.11, p. 313), G is bounded on (a,b]. Therefore the monotonic functions
\overline{G}(r)=\sup\left\{G(r_{1})\, \big|\, a<r_{1}\le r\right\} \text{\quad and\quad} \underline{G}(r)=\inf\left\{G(r_{1})\, \big|\, a<r_{1}\le r\right\}
are well defined on (a,b], and
\lim_{r\to a+}\overline{G}(r)=\overline{L} \text{\quad and\quad} \lim_{r\to a+}\underline{G}(r)=\underline{L}
both exist and are finite (Theorem 2.1.11, p. 47). From (A),
\begin{aligned} |G(r)-G(r_{1})|&=&|(G(r)-G(r_{0}))-(G(r_{1})-G(r_{0}))|\\ &\le &|G(r)-G(r_{0})|+|G(r_{1})-G(r_{0})|\le 2\epsilon,\end{aligned}
so \overline{G}(r)-\underline{G}(r)\le 2\epsilon. Since \epsilon is an arbitrary positive number, this implies that
\lim_{r\to a+}(\overline{G}(r)-\underline{G}(r))=0,
so \overline{L}=\underline{L}. Let L=\overline{L}=\underline{L}. Since
\underline{G}(r)\le G(r)\le \overline{G}(r),
it follows that \lim_{r\to a+} G(r)=L.
3. Theorem 5 (Cauchy Criterion for Uniform Convergence II) The improper integral
\int_{a}^{b}f(x,y)\,dx =\lim_{r\to a+}\int_{r}^{b}f(x,y)\,dx
converges uniformly on S if and only if, for each \epsilon>0, there is an r_{0}\in (a,b] such that
\tag{A} \left|\int_{r_{1}}^{r}f(x,y)\,dx\right|< \epsilon, \quad y\in S, \quad a <r,r_{1}\le r_{0}.
Suppose \int_{a}^{b} f(x,y)\,dx converges uniformly on S and \epsilon>0. From Definition 2, there is an r_{0}\in (a,b] such that
\tag{B} \left|\int_{a}^{r}f(x,y)\,dx\right| <\frac{\epsilon}{2} \text{\quad and\quad} \left|\int_{a}^{r_{1}}f(x,y)\,dx\right| <\frac{\epsilon}{2},\, y\in S, \, a< r,r_{1}\le r_{0}.
Since
\int_{r_{1}}^{r}f(x,y)\,dx= \int_{r_{1}}^{b}f(x,y)\,dx- \int_{r}^{b}f(x,y)\,dx
(B) and the triangle inequality imply (A).
For the converse, denote
F(y)=\int_{r}^{b}f(x,y)\,dx.
Since (A) implies that
|F(r,y)-F(r_{1},y)|\le \epsilon, \quad y\in S, \quad a< r, r_{1}\le r_{0},
Theorem 2 with G(r)=F(r,y) (y fixed but arbitrary in S) implies that \int_{a}^{b} f(x,y)\,dx converges pointwise for y\in S. Therefore, if \epsilon>0 then, for each y\in S, there is an r_{0}(y) \in (a,b] such that
\tag{C} \left|\int_{a}^{r}f(x,y)\,dx\right|\le \epsilon, \quad y\in S,\quad a<r\le r_{0}(y).
For each y\in S, choose r_{1}(y)\le \min[{r_{0}(y),r_{0}}]. Then
\int_{a}^{r}f(x,y)\,dx = \int_{a}^{r_{1}(y)}f(x,y)\,dx+ \int_{r_{1}(y)}^{r}f(x,y)\,dx, \quad
so (A), (C), and the triangle inequality imply that
\left|\int_{a}^{r} f(x,y)\,dx\right|\le 2\epsilon,\quad y\in S,\quad a<r\le r_{0}
4. From Definition 3, \int_{a}^{b}f(x,y)\,dx converges uniformly on S if and only if \int_{a}^{c}f(x,y)\,dx and \int_{c}^{b}f(x,y)\,dx both converge uniformly on S, where c\in(a,b). From Theorems 4 and Theorem 5, this is true if and only if, for any \epsilon>0 there are points r_{0} and s_{0} in (a,b) such that
\left|\int_{r}^{r_{1}}f(x,y)\,dx\right|\le \epsilon,\quad y\in S,\quad r_{0}\le r,r_{1}<b
and
\left|\int_{s_{1}}^{s}f(x,y)\,dx\right|\le \epsilon,\quad y\in S,\quad a< s,s_{1}<s_{0}.
These conditions are independent of c.
5. (a) If |f(x,y)|\le M on [a,b]\times [c,d] then
\left|\int_{r_{1}}^{r_{2}}f(x,y)\,dx\right|\le M|r_{2}-r_{1}|
so the Cauchy convergence theorems imply the conclusion.
Define f=f(x,y) on [0,1]\times [0,1] by
f(x,y)= \begin{cases}\displaystyle\frac{1}{y} &\text{if\quad} 0<y\le 1,\\ 1&\text{if\quad} y=0. \end{cases}
Then
\int_{r_{1}}^{r_{2}}f(x,y)\,dx= \begin{cases}\displaystyle\frac{r_{2}-r_{1}}{y} &\text{if\quad} 0<y\le 1,\\ r_{2}-r_{1}&\text{if\quad} y=0. \end{cases}
Therefore f does not satisfy the requirements of Cauchy’s convergence theorems.
6. In all parts I(y) denotes the given integral.
I(0)=\infty. If y\ne0 let u=xy; then I(y)=\displaystyle{\frac{1}{y}\int_{0}^{\infty}\frac{du}{1+u^{2}}}. If \rho>0 and \epsilon >0, choose r so that \displaystyle{\int_{r}^{\infty}\frac{du}{1+u^{2}}}< \rho\epsilon. Then \displaystyle{\frac{1}{|y|}\int_{r}^{\infty}\frac{du}{1+u^{2}}}<\epsilon if |y|\ge \rho, so I(y) converges uniformly on (-\infty, \rho]\bigcup [\rho,\infty) if \rho>0.
I(y)=\infty if y\le0. If y>0 let u=xy; then I(y)=\displaystyle{\frac{1}{y^{3}}\int_{0}^{\infty}e^{-u}u^{2}\,du}. If \rho>0 and \epsilon >0, choose r so that \displaystyle{\int_{r}^{\infty}e^{-u}u^{2}\,du}<\rho^{3}\epsilon. Then \displaystyle{\frac{1}{y^{3}}\int_{r}^{\infty}e^{-u}u^{2}\,du}<\epsilon if y\ge \rho, so I(y) converges uniformly on [\rho,\infty) if \rho>0.
I(y)=\infty if y\le0. If y>0 let u=xy^{1/2}; then I(y)=\displaystyle{y^{-n-1/2}\int_{0}^{\infty}u^{2n}e^{-u}\,du}. If \rho>0 and \epsilon >0, we can choose r so that \displaystyle{\int_{r}^{\infty}u^{2n}e^{-u}\,du}<\epsilon \rho^{n+1/2}. Then y^{-n-1/2}\displaystyle{\int_{r}^{\infty}u^{2n}e^{-u}\,du}<\epsilon if y\ge \rho, so I(y) converges uniformly on S=[\rho,\infty) if \rho>0.
Since I(-y)=-I(y), it suffices to assume that y>0. If u=yx^{2} then I(y)=\displaystyle{\frac{1}{2\sqrt{y}}\int_{0}^{\infty}\frac{\sin u\,du}{\sqrt{u}}}. From Example 3.4.14 (p. 162), this integral converges conditionally. If \rho>0 and \epsilon >0, we can choose r so that \displaystyle{\left|\int_{r}^{\infty}\frac{\sin u\,du}{\sqrt{u}}\right|}<2\epsilon\sqrt{\rho}, so I(y) converges uniformly on (-\infty,-\rho]\bigcup[\rho,\infty) if \rho>0.
If u=y^{2}x then \displaystyle{I(y)=3\int_{0}^{\infty}e^{-u}\,du -\frac{2}{y^{3}}\int_{0}^{\infty} ue^{-u}\,du}. If \rho>0, we can choose r so that \displaystyle{3\int_{r}^{\infty}e^{-u}\,du<\frac{\epsilon}{2}} and \displaystyle{\int_{r}^{\rho}ue^{-u}\,du<\frac{\rho^{3}\epsilon}{2}}. Then
\left|3\int_{r}^{\infty}e^{-u}\,du -\frac{3}{y^{3}}\int_{r}^{\infty} ue^{-u}\,du\right|<\epsilon \text{\quad if\quad} |y|\ge \rho,
so I(y) converges uniformly on (-\infty, \rho]\bigcup [\rho,\infty) if \rho>0.
I(y)=-\infty if y\le0. If y>0, let u=xy; then I(y)=\displaystyle{\frac{1}{y}\int_{0}^{\infty}(2u-u^{2})e^{-u}\,du}. If \rho>0 and \epsilon >0, we can choose r so that \displaystyle{\int_{r}^{\infty}|2u-u^{2}|e^{-u}\,du}<\epsilon \rho, so I(y) converges uniformly on S=[\rho,\infty) if \rho>0.
7. Theorem 7 (Weierstrass’s Test for Absolute Uniform Convergence II) Suppose f=f(x,y) is locally integrable (a,b] and, for some b_{0}\in (a,b],
\tag{A} |f(x,y)| \le M(x), \: y\in S, \: x\in (a,b_{0}],
where
\int_{a}^{b_{0}}M(x)\,dx<\infty.
Then \int_{a}^{b}f(x,y)\,dx converges absolutely uniformly on S.
Denote \int_{a}^{b_{0}}M(x)\,dx=L<\infty. By definition, for each \epsilon>0 there is an r_{0}\in (a,b_{0}] such that
L-\epsilon \le \int_{r}^{b_{0}}M(x)\,dx \le L,\quad a<r\le r_{0}.
Therefore, if a<r_{1}< r\le r_{0}, then
0\le \int_{r_{1}}^{r}M(x)\,dx=\left(\int_{r_{1}}^{b_{0}}M(x)\,dx -L\right)- \left(\int_{r}^{b_{0}}M(x)\,dx -L\right)<\epsilon.
This and (A) imply that
\int_{r_{1}}^{r}|f(x,y)|\,dx\le \int_{r_{1}}^{r} M(x)\,dx <\epsilon, \: y\in S,\: a<r_{1}\le r_{0}\le b.
Now Theorem 5 implies the stated conclusion.
|e^{-xy}\sin x|\le e^{-\rho x} if y\ge\rho and \int_{\rho}^{\infty}e^{-\rho x}\,dx<\infty.
\displaystyle{\int_{0}^{\infty}\frac{\sin x\,dx}{x^{y}}}=I_{1}(y)+I_{2}(y), where
I_{1}(y)=\int_{0}^{1}\frac{\sin x\,dx}{x^{y}} \text{\quad and\quad} I_{2}(y)=\int_{1}^{\infty}\frac{\sin x\,dx}{x^{y}}
are both improper integrals. Since
\sin x= x-\left(\frac{x^{3}}{3!}-\frac{x^{5}}{5!}\right) -\left(\frac{x^{7}}{7!}-\frac{x^{9}}{9!}\right)+ \cdots <x,\quad 0\le x \le 1.
If 0\le 1 and y\le d\le 2, then
\left|\frac{\sin x}{x^{y}}\right|\le x^{1-y}\le x^{1-d} \text{\:so\:} \int_{0}^{1}x^{-1+y}\,dx<\int_{0}^{1}x^{-1+d}\,dx=\frac{1}{2-d}
so I_{1}(y) converges absolutely uniformly on S. Since c>1,
\frac{|\sin x|}{x^{y}}\le x^{-c} \text{\quad and\quad} \int_{1}^{\infty}x^{-c} \,dx=\frac{1}{c-1}\text{\quad if \quad}
I_{2}(y) converges absolutely uniformly on S.
If x\ge 1 then \displaystyle{e^{-px}\left|\frac{\sin xy}{x}\right|\le e^{-px}} for all y and \displaystyle{\int_{1}^{\infty}e^{-px}\,dx<\infty}, since p>0.
\displaystyle{\frac{e^{xy}}{(1-x)^{y}}}\le \displaystyle{\frac{e^{b}}{(1-x)^{b}}}, if 0\le x<1 and y\le b, and \displaystyle{\int_{0}^{1}(1-x)^{-b}\,dx}<\infty if b<1.
If |y|\ge \rho>0 then \displaystyle{\left|\frac{\cos xy}{1+x^{2}y^{2}}\right|\le \frac{1}{1+\rho^{2}x^{2}}} for all x, and \displaystyle{\int_{0}^{\infty}\frac{dx}{1+\rho^{2}x^{2}}}<\infty.
If y\ge \rho>0 then e^{-x/y}\le e^{-x/\rho} and \displaystyle{\int_{0}^{\infty}e^{-x/\rho}\,dx<\infty}.
If |y|\le \rho then e^{xy}e^{-x^{2}}\le e^{x\rho}e^{-x^{2}} and \displaystyle{\int_{-\infty}^{\infty}e^{x\rho}e^{-x^{2}}\,dx}<\infty.
If |x|\ge 1 then |\cos xy-\cos ax|\le 2 and \displaystyle{\int_{1}^{\infty}\frac{2\,dx}{x^{2}}}<\infty.
If y\ge \rho>0 then e^{-yx^{2}}\le e^{-\rho x^{2}} and \displaystyle{\int_{0}^{\infty} x^{2n}e^{-\rho x^{2}}\,dx<\infty}.
9. (a) If 0<x<1 then |x^{y-1}e^{-x}|<x^{c-1}. Therefore, since \displaystyle{\int_{0}^{1}x^{c-1}\,dx}<\infty if c>0, \int_{0}^{1}x^{y-1}e^{-x}\,dx converges uniformly on [c,\infty) if c>0, by Theorem 7. If x>1 then |x^{y-1}e^{-x}|\le x^{d-1}e^{-x} if y\le d. Therefore, since \displaystyle{\int_{1}^{\infty}x^{d-1}e^{-x}}\,dx<\infty, \displaystyle{\int_{1}^{\infty}x^{y-1}e^{-x}}\,dx converges uniformly on (-\infty,d] for every d, by Theorem 6. Hence, \displaystyle{\int_{0}^{\infty}x^{y-1}e^{-x}\,dx} converges uniformly on [c,d] if c>0.
If y>0 then
\Gamma(y)=\int_{0}^{\infty}x^{y-1}e^{-x}\,dx =\frac{x^{y}e^{-x}}{y}\biggr|_{0}^{\infty}+ \frac{1}{y}\int_{0}^{\infty}x^{y}e^{-x}\,dx =\frac{\Gamma(y+1)}{y}. \tag{A}
Therefore
\Gamma(y)=\frac{\Gamma(y+n)}{y(y+1)\cdots (y+n-1)} \tag{B}
is true when n=1. Now suppose it is true for given positive integer n, and replace y by y+n in (A):
\Gamma(y+n)=\frac{\Gamma(y+n+1)}{y+n}.
Substituting this into (B) yields
\Gamma(y)=\frac{\Gamma(y+n+1)}{y(y+1)\cdots (y+n)},
which completes the induction.
If -n <y<-n+1 with n\le 1 then 0<y+n<1 and we can compute \Gamma(y+n) from the definition in part (a):
\Gamma(y+n)=\int_{0}^{\infty}x^{y+n-1}e^{-x}\,dx.
Then we can define \Gamma(y) by (B).
The assertion is true if n=1, since
\Gamma(2)=\int_{0}^{\infty}xe^{-x}\,dx = -xe^{-x}\biggr|_{0}^{\infty}+\int_{0}^{\infty} e^{-x}\,dx = 1.
If \Gamma(n+1)=n! for some n\ge 1, then
\begin{aligned} \Gamma(n+2)&=&\int_{0}^{\infty}x^{n+1}e^{-x}\,dx= -e^{-x} x^{n+1}\biggr|_{0}^{\infty}+(n+1) \int_{0}^{\infty}x^{n+1}e^{-x}\,dx\\ &=&(n+1)\Gamma(n+1)=(n+1)n!=(n+1)!,\end{aligned}
which completes the induction proof.
The change of variable x=st yields
\int_{0}^{\infty}e^{-st}t^{\alpha}\,dt=\frac{1}{s^{\alpha+1}} \int_{0}^{\infty}x^{\alpha}e^{-x}\,dx=\frac{1}{s^{\alpha+1}} \Gamma(\alpha+1),
from the definition of the Gamma function.
Since |g_{x}(x,y)| is monotonic with respect to x,
\tag{A} \int_{r}^{r_{1}}|g_{x}(x,y)|\,dx=|g(r_{1},y)-g(r,y)|,\quad a\le r<r_{1}<b.
From Assumption (a) of Theorem 8, if \epsilon>0 there is an r_{0}\in [a,b) such that
|g(s,y)|\le \epsilon,\quad y\in S,\quad r_{0}\le s<b.
Therefore, (A) implies that
\int_{r}^{r_{1}}|g_{x}(x,y)\,dx\le 2\epsilon,\quad y\in S,\quad r_{0}\le r\le r_{1}<b.
Now Theorem 4 implies that \int_{a}^{b}|g(x,y)|\,dx converges uniformly on S, which is assumption (c) of Theorem 8.
) If g, g_{x}, and h are continuous on (a,b]\times S then
\int_{a}^{b}g(x,y)h(x,y)\,dx
converges uniformly on S if the following conditions are satisfied:
()0pt0pt14pt 8pt22pt0pt
\displaystyle{\lim_{x\to a+}\left\{\sup_{y\in S}|g(x,y)|\right\}=0};
There is a constant M such that
\sup_{y\in S}\left|\int_{x}^{b}h(u,y)\,du\right| \le M, \quad a< x\le b;
\int_{a}^{b}|g_{x}(x,y)|\,dx converges uniformly on S.
If
\tag{A} H(x,y)=\int_{x}^{b}h(u,y)\,du
then integration by parts yields
\begin{aligned} \int_{r_{1}}^{r}g(x,y)h(x,y)\,dx&=&-\int_{r_{1}}^{r}g(x,y)H_{x}(x,y)\,dx \\ &=&-g(r,y)H(r,y)+g(r_{1},y)H(r_{1},y)\\ &&+\int_{r_{1}}^{r}g_{x}(x,y)H(x,y)\,dx.\end{aligned}
Therefore, since assumption (b) and (A) imply that |H(x,y)|\le M, (x,y)\in (a,b]\times S,
\tag{B} \left|\int_{r_{1}}^{r}g(x,y)h(x,y)\,dx\right|\le M\left(2\sup_{a<x\le r} |g(x,y)|+\int_{r_{1}}^{r}|g_{x}(x,y)|\,dx\right)
on [r_{1},r]\times S. Now suppose \epsilon>0. From assumption (a), there is an r_{0} \in [a,b) such that |g(x,y)|<\epsilon on S if a< x \le r_{0} \le b. From assumption (c) and Theorem 5, there is an s_{0}\in (a,b] such that
\int_{r_{1}}^{r}|g_{x}(x,y)|\,dx<\epsilon,\quad y\in S,\quad a<r_{1}<r\le s_{0}.
Therefore (B) implies that
\left|\int_{r_{1}}^{r}g(x,y)h(x,y)\right| < 3M\epsilon,\quad y\in S,\quad a<r_{1}<r\min(r_{0},s_{0})
Now Theorem 5 implies the stated conclusion.
Denote F(y)=\displaystyle{\int_{1}^{\infty}\frac{\sin xy}{x^{y}}} and, with 1\le r< r_{1},
\begin{aligned} F(r,r_{1},y)=\int_{r}^{r_{1}}\frac{\sin xy}{x^{y}}\,dx &=& -\frac{\cos xy}{yx^{y}}\biggr|_{r}^{r_{1}}- \int_{r}^{r_{1}}\frac{\cos xy}{x^{y+1}}\,dx\\ &=& \frac{\cos r y}{yr^{y}}-\frac{\cos r_{1}y}{yr_{1}^{y}}- \int_{r}^{r_{1}}\frac{\cos xy}{x^{y+1}}\,dx.\end{aligned}
Therefore
|F(r,r_{1},y)|\le \frac{2}{yr^{y}}+\int_{r}^{r_{1}}x^{-y-1}\,dx<\frac{3}{yr^{y}}, \quad r,y>0.
Now Theorem 4 implies that F(y) converges uniformly on [\rho,\infty) if \rho>0.
(b) Denote F(y)=\displaystyle{\int_{2}^{\infty}\frac{\sin xy}{\log x}\,dx} and, with 2\le r< r_{1},
F(r,r_{1},y)=\int_{r}^{r_{1}}\frac{\sin xy}{\log x}\,dx =-\frac{\cos xy}{y\log x}\biggr|_{r}^{r_{1}}- \frac{1}{y}\int_{r}^{r_{1}}\frac{\cos xy}{x(\log x)^{2}}\,dx.
Therefore
|F(r,r_{1},y)|\le \frac{1}{y}\left|\frac{2}{\log r}+\int_{r}^{r_{1}} \frac{dx}{x(\log x)^{2}}\right|\le \frac{3}{y\log r}.
Now Theorem 4 implies that F(y) converges uniformly on [\rho,\infty) if \rho>0.
(c) Denote F(y)=\displaystyle{\int_{0}^{\infty}\frac{\cos xy}{x+y^{2}}}, and, with 0<r<r_{1},
F(r,r_{1},y)=\int_{r}^{r_{1}}\frac{\cos xy}{x+y^{2}}\,dx= \frac{1}{y}\left(\frac{\sin xy}{x+y^{2}}\biggr|_{r}^{r_{1}} +\int_{r}^{r_{1}}\frac{\sin xy}{(x+y^{2})^{2}}\,dx\right),
so
|F(r,r_{1},y)|\le \frac{3}{y(r+y^{2})}.
Now Theorem 4 implies that F(y) converges uniformly on [\rho,\infty) if \rho>0.
(d) Denote F(y)=\displaystyle{\int_{0}^{\infty}\frac{\sin xy}{1+xy}}, and, with 0<r<r_{1},
F(r,r_{1},y)=\int_{r}^{r_{1}}\frac{\sin xy}{1+xy}\,dx= -\frac{\cos xy}{y(1+xy)}\biggr|_{r}^{r_{1}}-\int_{r}^{r_{1}} \frac{\cos xy}{y^{2}(1+xy)^{2}}\,dx,
so
|F(r,r_{1},y)|\le \frac{3}{y(1+ry)}.
Now Theorem 4 implies that F(y) converges uniformly on [\rho,\infty) if \rho>0.
13. Integration by parts yields
\begin{aligned} \int_{r}^{r_{1}}g(x,y)h(x,y)\,dx&=&\int_{r}^{r_{1}}g(x,y)H_{x}(x,y)\,dx\\ &=&g(r_{1},y)H(r_{1},y)-g(r,y)H(r,y)\\ &&-\int_{r}^{r_{1}}g_{x}(x,y)H(x,y)\,dx,\end{aligned}
so
\left|\int_{r}^{r_{1}}g(x,y)h(x,y)\,dx\right|\le 2\sup_{x\ge r}\left\{\left\{\sup_{y\in S}|g(x,y)H(x,y)|\right\}\right\}+ \left|\int_{r}^{r_{1}}g_{x}(x,y)H(x,y)\,dx\right|.
Now suppose \epsilon\ge 0. From our first assmption, there is an s_{0}\in [a,b) such that
\sup_{x\ge r}\left\{\left\{\sup_{y\in S}|g(x,y)H(x,y)|\right\}\right\}<\epsilon, \quad s_{0}\le r<b.
Since \int_{a}^{b}g_{x}(x,y)H(x,y)\,dx converges uniformly on S, Theorem 4 implies that there is an r_{0}\in [a,b) such that
\left|\int_{r}^{r_{1}}g_{x}(x,y)H(x,y)\,dx\right|\le \epsilon, \quad y\in S, \quad r_{0}\le r<r_{1}<b.
Therefore,
\left|\int_{r}^{r_{1}}g(x,y)h(x,y)\,dx\right|\le 2\epsilon, \quad y\in S, \quad \max(r_{0},s_{0})\le r<r_{1}<b.
Now Theorem 4 implies that \int_{a}^{b}g_{x}(x,y)h(x,y)\,dx converges uniformly on S.
14. Theorem 10 If f=f(x,y) is continuous on (a,b]\times [c,d] and
\tag{A} F(y)=\int_{a}^{b}f(x,y)\,dx
converges uniformly on [c,d], then F is continuous on [c,d]. Moreover,
\tag{B} \int_{c}^{d}\left(\int_{a}^{b}f(x,y)\,dx\right)\,dy =\int_{a}^{b}\left(\int_{c}^{d}f(x,y)\,dy\right)\,dx.
We will first show that F in (A) is continuous on [c,d]. Since F converges uniformly on [c,d], Definition 1 implies that if \epsilon>0, there is an r \in (a,b] such that
\left|\int_{a}^{r}f(x,y)\,dx\right|\le \epsilon, \quad c \le y \le d.
Therefore, if y and y_{0} are in [c,d], then
\begin{aligned} |F(y)-F(y_{0})|&=& \left|\int_{a}^{b}f(x,y)\,dx-\int_{a}^{b}f(x,y_{0})\,dx\right|\\ &\le&\left|\int_{r}^{b}[f(x,y)-f(x,y_{0})]\,dx\right|+ \left|\int_{a}^{r}f(x,y)\,dx\right|\\ &&+\left|\int_{a}^{r}f(x,y_{0})\,dx\right|\\\end{aligned}
so
\tag{C} |F(y)-F(y_{0})| \le \int_{r}^{b}|f(x,y)-f(x,y_{0})|\,dx +2\epsilon.
Since f is uniformly continuous on the compact set [r,b]\times [c,d] (Corollary 5.2.14, p. 314), there is a \delta>0 such that
|f(x,y)-f(x,y_{0})|<\epsilon
if (x,y) and (x,y_{0}) are in [r,b]\times [c,d] and |y-y_{0}|<\delta. This and (C) imply that
|F(y)-F(y_{0})|<(r-a)\epsilon +2\epsilon<(b-a+2)\epsilon
if y and y_{0} are in [c,d] and |y-y_{0}|<\delta. Therefore F is continuous on [c,d], so the integral on left side of (B) exists. Denote
\tag{D} I= \int_{c}^{d}\left(\int_{a}^{b}f(x,y)\,dx\right)\,dy.
We will show that the improper integral on the right side of (B) converges to I. To this end, denote
I(r)= \int_{r}^{b}\left(\int_{c}^{d}f(x,y)\,dy\right)\,dx.
Since we can reverse the order of integration of the continuous function f over the rectangle [r,b]\times [c,d] (Corollary 7.2.2, p. 466),
I(r)=\int_{c}^{d}\left(\int_{r}^{b}f(x,y)\,dx\right)\,dy.
From this and (D),
I-I(r)=\int_{c}^{d}\left(\int_{a}^{r}f(x,y)\,dx\right)\,dy.
Now suppose \epsilon>0. Since \int_{a}^{b}f(x,y)\,dx converges uniformly on [c,d], there is an r_{0}\in (a,b] such that
\left|\int_{a}^{r}f(x,y)\,dx\right|<\epsilon, \quad a<r<r_{0}, %###
so |I-I(r)|<(d-c)\epsilon, a<r<r_{0}. Hence,
\lim_{r\to a+}\int_{r}^{b}\left(\int_{c}^{d}f(x,y)\,dy\right)\,dx= \int_{c}^{d}\left(\int_{a}^{b}f(x,y)\,dx\right)\,dy,
which completes the proof of (B).
15. Theorem 11 Let f and f_{y} be continuous on (a,b]\times [c,d], and suppose that
F(y)=\int_{a}^{b}f(x,y)\,dx
converges for some y_{0} \in [c,d] and
G(y)=\int_{a}^{b}f_{y}(x,y)\,dx
converges uniformly on [c,d]. Then F converges uniformly on [c,d] and is given explicitly by
F(y)=F(y_{0})+\int_{y_{0}}^{y} G(t)\,dt,\quad c\le y\le d.
Moreover, F is continuously differentiable on [c,d] and
\tag{A} F'(y)=G(y), \quad c \le y \le d,
where F'(c) and f_{y}(x,c) are derivatives from the right, and F'(d) and f_{y}(x,d) are derivatives from the left.
Let
F_{r}(y)=\int_{r}^{b}f(x,y)\,dx, \quad a\le r<b,\quad c \le y \le d.
Since f and f_{y} are continuous on [r,b]\times [c,d], Theorem 1 implies that
F_{r}'(y)=\int_{r}^{b}f_{y}(x,y)\,dx, \quad c \le y \le d.
Therefore
\begin{aligned} F_{r}(y)&=&F_{r}(y_{0})+\int_{y_{0}}^{y}\left( \int_{r}^{b}f_{y}(x,t)\,dx\right)\,dt\\ &=&F(y_{0})+\int_{y_{0}}^{y}G(t)\,dt \\&&+(F_{r}(y_{0})-F(y_{0})) -\int_{y_{0}}^{y}\left(\int_{a}^{r}f_{y}(x,t)\,dx\right)\,dt, \quad c \le y \le d.\end{aligned}
Therefore,
\tag{B} \left|F_{r}(y)-F(y_{0})-\int_{y_{0}}^{y}G(t)\,dt\right| \le |F_{r}(y_{0})-F(y_{0})| +\left|\int_{y_{0}}^{y} \int_{a}^{r}f_{y}(x,t)\,dx\right|\,dt.
Now suppose \epsilon>0. Since we have assumed that \lim_{r\to a+}F_{r}(y_{0})=F(y_{0}) exists, there is an r_{0} in (a,b) such that
|F_{r}(y_{0})-F(y_{0})|<\epsilon,\quad a<r< r_{0}.
Since we have assumed that G(y) converges for y\in[c,d], there is an r_{1} \in (a,b] such that
\left|\int_{a}^{b}f_{y}(x,t)\,dx\right|<\epsilon, \quad t\in[c,d], \quad a<r\le r_{1}.
Therefore, (B) yields
\left|F_{r}(y)-F(y_{0})-\int_{y_{0}}^{y}G(t)\,dt\right|< \epsilon(1+|y-y_{0}|) \le \epsilon(1+d-c)
if a<r<\min(r_{0},r_{1}) and t\in[c,d]. Therefore F(y) converges uniformly on [c,d] and
F(y)=F(y_{0})+\int_{y_{0}}^{y}G(t)\,dt, \quad c \le y \le d.
Since G is continuous on [c,d] by Theorem 10, (A) follows from differentiating this (Theorem 3.3.11, p. 141).
16. Since
|f(x)\cos xy|\le |f(x)|,\quad |f(x)\sin xy|\le |f(x)|,\text{\quad and\quad} \int_{-\infty}^{\infty} |f(x)|\,dx<\infty,
Theorems 6 and 7 imply that \int_{-\infty}^{\infty}f(x)\cos xy\,dx and \int_{-\infty}^{\infty}f(x) \sin xy\,dx converge uniformly on (-\infty,\infty), so Theorem 10 implies that C(y) and S(y) are continuous on (-\infty,\infty).
If y\ne0, integrating by parts yields
\begin{aligned} C(y)&=&f(x)\frac{\sin xy}{y}\biggr|_{a}^{\infty}-\frac{1}{y} \int_{a}^{\infty}f'(x)\sin xy \,dx\\ &=&-f(a)\frac{\sin ay}{y} -\frac{1}{y}\int_{a}^{\infty}f'(x)\sin xy \,dx\end{aligned}
and
\begin{aligned} S(y)&=&-f(x)\frac{\cos xy}{y}\biggr|_{a}^{\infty}+\frac{1}{y} \int_{a}^{\infty}f'(x)\cos xy \,dx \\ &=&f(a)\frac{\cos ay}{y}+ \frac{1}{y}\int_{a}^{\infty}f'(x)\cos xy \,dx,\end{aligned}
since \lim_{x\to\infty} f(x)=0. From Exercise [exer:17] with f replaced by f', \int_{1}^{\infty}f'(x)\cos xy \,dx and \int_{1}^{\infty}f'(x)\cos xy\,dx are continuous on (-\infty,\infty). Therefore C(y) and S(y) are continuous on (-\infty,0)\cup(0,\infty).
To see that C and S are not necessarily continuous at y=0, let a=1 and f(x)=1/x, so
\lim_{x\to\infty}f(x)=0\text{\quad and\quad} \int_{1}^{\infty}|f'(x)|=\int_{1}^{\infty}\frac{dx}{x^{2}}=1.
Then
C(y)=\lim_{r\to\infty}\int_{1}^{r}\frac{\cos xy}{x}\,dx \text{\quad and\quad} S(y)=\lim_{r\to\infty}\int_{1}^{r}\frac{\sin xy}{x}\,dx,\quad y\ne0.
If y>0 make the change of variable u=xy to see that
C(y)=\lim_{r\to\infty}\int_{y}^{ry}\frac{\cos u}{u}\,du= \int_{y}^{\infty}\frac{\cos u}{u}\,du
and
S(y)=\lim_{r\to\infty}\int_{y}^{ry}\frac{\sin u}{u}\,du. S(y)=\int_{y}^{\infty}\frac{\sin u}{u}\,du.
Therefore \lim_{y\to 0+}C(y)=\infty, so C is not continuous at y=0. Since S(0)=0 and \lim_{y\to 0+}S(y)= \displaystyle{\int_{0}^{\infty}\frac{\sin u}{u}\,du}\ne 0, S is not continuous at y=0.
The integral diverges if y=0. If y\ne0 substitute u=|y|x to obtain
F(y)=\int_{0}^{\infty}\frac{dx}{1+x^{2}y^{2}}= \frac{1}{|y|}\int_{0}^{\infty}\frac{du}{1+u^{2}} =\frac{1}{|y|}\tan^{-1}u\biggr|_{0}^{\infty}=\frac{\pi}{2|y|}, \tag{A}
so F(y) converges for all y\ne0. To test for uniform convergence, suppose |y|>0 and 0<r<r_{1}. Then
\int_{r}^{r_{1}}\frac{dx}{1+x^{2}y^{2}} =\frac{1}{|y|}\int_{r|y|}^{r_{1}|y|} \frac{du}{1+u^{2}} <\frac{1}{\rho}\int_{r\rho}^{\infty}\frac{du}{1+u^{2}}
if |y|\ge \rho. If \epsilon>0 there is an \alpha>0 such that \displaystyle{\frac{1}{\rho}\int_{\alpha}^{\infty}\frac{du}{1+u^{2}}}<\epsilon. Therefore \displaystyle{\int_{r}^{r_{1}}\frac{dx}{1+x^{2}y^{2}}}<\epsilon if \alpha/\rho<r<r_{1}. Now Theorem 4 implies that F(y) converges uniformly on (-\infty,-\rho]\cup[\rho,\infty) if \rho>0.
To evaluate
I=\displaystyle{\int_{0}^{\infty}\frac{\tan^{-1}ax-\tan^{-1}bx}{x}\,dx},
we note that
\frac{\tan^{-1}ax-\tan^{-1}bx}{x}=\int_{b}^{a}\frac{dy}{1+x^{2}y^{2}}.
Therefore
I=\int_{0}^{\infty}\,dx \int_{b}^{a}\frac{dy}{1+x^{2}y^{2}} =\int_{b}^{a}\,dy\int_{0}^{\infty}\frac{dx}{1+x^{2}y^{2}} =\frac{\pi}{2}\int_{b}^{a}\frac{dy}{y}=\frac{\pi}{2}\log\frac{a}{b},
where the second equality is valid because of the uniform convergence of F(y) on the closed interval with endpoints a and b, and the third equality follows from (A).
F(y) is a proper integral if y\ge 0 and it diverges if y\le -1. If -1<y<0, then
F(y)=\int_{0}^{1}x^{y}\,dx=\frac{x^{y+1}}{y+1}\biggr|_{0}^{1}=\frac{1}{y+1} \tag{A}
is convergent. Since
\int_{0}^{r}x^{y}\,dx=\frac{x^{y+1}}{y+1}\biggr|_{0}^{r}= \frac{r^{y+1}}{y+1}.
and
\frac{\partial}{\partial y}\left (\frac{r^{y+1}}{y+1}\right) =\frac{r^{y+1}}{y+1}\left(\log r-\frac{1}{y+1}\right)<0 \text{\quad if \quad} 0<r\le 1 \text{\quad and\quad} y>-1,
it follows that
\left|\int_{0}^{r}x^{y}\,dx\right|\le \frac{r^{\rho+1}}{\rho +1} \text{\quad if\quad} 0<r\le 1\text{\quad and\quad} -1<\rho\le y.
Therefore, Theorem 5 implies that F(y) converges uniformly on [\rho,\infty) if \rho>-1.
Now Theorem 11 implies that
I=\displaystyle{\int_{0}^{1}\frac{x^{a}-x^{b}}{\log x}\,dx} = \int_{0}^{1}\,dx \int_{b}^{a}x^{y}\,dy =\int_{b}^{a}\,dy\int_{0}^{1}x^{y}\,dx =\int_{b}^{a}\frac{dy}{y+1}=\log\frac{a+1}{b+1}.
\displaystyle{F(y)=\int_{0}^{\infty} e^{-yx}\cos x \,dx=\frac{y}{y^{2}+1}}. Since
\left|\int_{r}^{\infty}e^{-yx} \cos x\,dx\right|\le \int_{r}^{\infty}e^{-xy}\,dx=\frac{e^{-yr}}{y},
Theorem 4 (or Theorem 6) implies that F(y) converges uniformly on [\rho,\infty) if \rho>0. Therefore, Theorem implies that if a, b>0 then
\begin{aligned} I&=&\int_{0}^{\infty}\frac{e^{-ax}-e^{-bx}}{x}\cos x\,dx =\int_{0}^{\infty}\cos x\,dx\int_{a}^{b}e^{-yx}\,dy \\ &=&\int_{a}^{b} \,dy\int_{0}^{\infty}e^{-yx}\cos x\,dx =\int_{a}^{b}\frac{y}{y^{2}+1}\,dy=\frac{1}{2}\log\frac{b^{2}+1}{a^{2}+1}.\end{aligned}
\displaystyle{F(y)=\int_{0}^{\infty} e^{-yx}\sin x \,dx=\frac{1}{y^{2}+1}}. Since
\left|\int_{r}^{\infty}e^{-yx} \sin x\,dx\right|\le \int_{r}^{\infty}e^{-yx}\,dx=\frac{e^{-yr}}{y},
Theorem 4 (or Theorem 6) implies that F(y) converges uniformly on every [\rho,\infty) if \rho>0. Therefore, if a, b>0 then Theorem 11 implies that
\begin{aligned} I&=&\int_{0}^{\infty}\frac{e^{-ax}-e^{-bx}}{x}\sin x\,dx =\int_{0}^{\infty}\sin x\,dx\int_{a}^{b}e^{-yx}\,dy \\ &=&\int_{a}^{b} \,dy\int_{0}^{\infty}e^{-yx}\sin x\,dx =\int_{a}^{b}\frac{1}{y^{2}+1}\,dy=\tan^{-1}b-\tan^{-1}a.\end{aligned}
(e) \displaystyle{F(y)=\int_{0}^{\infty} e^{-x}\sin xy \,dx=\frac{y}{y^{2}+1}}. Since
\left|\int_{r}^{\infty}e^{-x} \sin xy\,dx\right|\le \int_{r}^{\infty}e^{-x}\,dx=e^{-r},
Theorem 4 (or Theorem 6) implies that F(y) converges uniformly on [\rho,\infty) if \rho>0. Therefore Theorem 11 implies that
\begin{aligned} I&=&\int_{0}^{\infty}e^{-x}\frac{1-\cos ax}{x}\,dx =\int_{0}^{\infty}e^{-x}\,dx\int_{0}^{a}\sin xy\,dy\\ &=&\int_{0}^{a}\,dy\int_{0}^{\infty}e^{-x}\sin xy\,dx =\int_{0}^{a} \frac{y}{y^{2}+1}\,dy=\frac{1}{2}\log(1+a^{2}).\end{aligned}
\displaystyle{F(y)=\int_{0}^{\infty} e^{-x}\cos xy \,dx=\frac{1}{y^{2}+1}}. Since
\left|\int_{r}^{\infty}e^{-x} \cos xy\,dx\right|\le \int_{r}^{\infty}e^{-x}\,dx=e^{-r},
Theorem 4 (or Theorem 6) implies that F(y) converges uniformly on [\rho,\infty) if \rho>0. Therefore Theorem 11 implies that
\begin{aligned} I&=&\int_{0}^{\infty}e^{-x}\sin ax\,dx =\int_{0}^{\infty}e^{-x}\,dx\int_{0}^{a}\cos xy\,dy\\ &=&\int_{0}^{a}\,dy\int_{0}^{\infty}e^{-x}\cos xy\,dx =\int_{0}^{a} \frac{1}{y^{2}+1}\,dy=\tan^{-1}a.\end{aligned}
19. (a) We start with
F(y)=\int_{0}^{1} x^{y}\,dx =\frac{1}{y+1}\quad y>-1. \tag{A}
Formally differentiating this yields
F^{(n)}(y)=\int_{0}^{1}(\log x)^{n} x^{y}\,dx =\frac{(-1)^{n}n!}{(y+1)^{n+1}},\quad y>-1. \tag{B}
To justify this we will show by induction that the improper integrals
I_{n}(y)=\int_{0}^{1}(\log x)^{n} x^{y}\,dx,\quad n=0,1,2, \dots
converge uniformly on [\rho,\infty) if \rho>-1. We begin with n=0:.
\int_{0}^{r}x^{y}\,dx = \frac{x^{y+1}}{y+1}\biggr|_{r_{1}}^{r}=\frac{r^{y+1}}{y+1}\le \frac{r^{y+1}}{\rho+1},\quad -1<\rho\le y.
so I_{0}(y)=F(y) converges uniformly on [\rho,\infty) if \rho>-1. Now suppose that I_{n}(y) converges uniformly on [\rho,\infty). Integrating by parts yields
\begin{aligned} \int_{r_{1}}^{r}(\log x)^{n+1}x^{y}\,dx&=& \frac{r^{y+1}(\log r)^{n+1}-r_{1}^{y+1}(\log r_{1})^{n+1}} {y+1}\\ &&-\frac{n+1}{y+1} \int_{r_{1}}^{r}(\log x)^{n}x^{y}\,dx, \quad -1<y<\infty.\end{aligned}
Letting r_{1}\to 0 yields
\tag{C} \int_{0}^{r}(\log x)^{n+1} x^{y}\,dx =\frac{r^{y+1}(\log r)^{n+1}}{y+1} -\frac{n+1}{y+1}\int_{0}^{r}(\log x)^{n}x^{y}\,dx.
Since the integral on the right converges, it follows that the integral on the left converges; in fact
\int_{0}^{1}(\log x)^{n+1}x^{y}\,dx= -\frac{n+1}{y+1} \int_{0}^{1}(\log x)^{n}x^{y}\,dx.
We must still show that the integral on the left converges uniformly on [\rho,\infty) if
\rho>-1. To this end, note from (C) that
\tag{D} \left|\int_{0}^{r}(\log x)^{n+1} x^{y}\,dx\right| \le \left|\frac{r^{\rho+1}(\log r)^{n+1}}{\rho+1}\right| +\frac{n+1}{\rho+1}\left|\int_{0}^{r}(\log x)^{n}x^{y}\,dx\right|
if y\ge \rho, Now suppose \epsilon>0. Since \displaystyle{\lim_{r\to0+}r^{\rho+1}(\log r)^{n+1}=0}, there is an r_{1}\in (0,1) such that
\left|\frac{r^{\rho+1}(\log r)^{n+1}}{\rho+1}\right| \le \frac{\epsilon}{2} \text{\quad if \quad} 0<r<r_{1}.
Since \int_{0}^{r}(\log x)^{n}x^{y}\,dx is uniformly convergent (by our induction assumption), there is r_{2}\in (0,1) such that
\frac{n+1}{\rho+1}\left|\int_{0}^{r}(\log x)^{n}x^{y}\,dx\right|\le \frac{\epsilon}{2},\quad y\ge \rho,
Now (D) implies that
\left|\int_{0}^{r}(\log x)^{n+1} x^{y}\,dx\right|\epsilon,\quad y\in [\rho,\infty),\quad 0<r<\min(r_{1},r_{2}).
This, Theorem 11, and an easy induction argument imply (B).
(b) Substituting x=u\sqrt{y} yields
F(y)=\int_{0}^{\infty}\frac{dx}{x^{2}+y}=\frac{1}{\sqrt{y}}\int_{0}^{\infty} \frac{du}{u^{2}+1} =\frac{\pi}{2\sqrt{y}},\quad y>0. \tag{A}
Formally differentiating this yields yields
\begin{aligned} \int_{0}^{\infty}\frac{dx}{(x^{2}+y)^{n+1}} &=&\frac{\pi}{2n+1}1\cdot 3\cdots(2n-1)y^{-n-1/2} =\frac{\pi}{2^{2n+1}}\frac{(2n)!}{n!}y^{-n-1/2}\\ &=&\frac{\pi}{2^{2n+1}}\binom{2n}{n}y^{-n-1/2},\quad y>0.\end{aligned}
Theorem 11 implies that the formal differentiation is legitimate, since, if y\ge 0 and r>0, then
\int_{r}^{\infty}\frac{dx}{(x^{2}+y)^{n+1}}\le \int_{r}^{\infty}x^{-2n-2}dx=\frac{r^{-2n-1}}{(2n-1)};
hence, the improper integrals \displaystyle{\int_{0}^{\infty}\frac{dx}{(x^{2}+y)^{n+1}}}, n=0, 1, 2, … converge uniformly on [0,\infty).
Denote I_{n}(y)=\displaystyle{\int_{0}^{\infty}x^{2n+1}e^{-yx^{2}}\,dx}. Then
I_{0}(y)=\int_{0}^{\infty}xe^{-yx^{2}}= \frac{1}{2}\int_{0}^{\infty}2xe^{-yx^{2}}\,dx =-\frac{1}{2y}e^{-yx^{2}}\biggr|_{0}^{\infty}=\frac{1}{2y}.
Since
\int_{r}^{\infty}x^{2n+1}e^{-yx^{2}} \,dx\le \int_{r}^{\infty}x^{2n+1}e^{-\rho x^{2}} \,dx\text{\quad if\quad} 0<\rho\le r,
if n\ge 0, we can differentiate I_{n} formally with respect to y\in (0,\infty) to obtain
I_{n}(y)=(-1)^{n}I_{0}^{(n)}=\frac{n!}{2y^{n+1}}.
(d) Denote
\begin{aligned} I(y)&=&\int_{0}^{\infty}y^{x}\,dx =\int_{0}^{\infty}e^{x\log y}\,dx =\frac{1}{\log y}\int_{0}^{\infty}(\log y) y^{x}\,dx \\ &=&\frac{y^{x}}{\log y}\biggr|_{0}^{\infty}=-\frac{1}{\log y}\quad 0<y<1.\end{aligned}
Formally differentiating this yields I'(y)=\displaystyle{\int_{0}^{\infty}xy^{x-1}\,dx}. There are two improper integrals here: J_{1}(y)=\displaystyle{\int_{0}^{1}xy^{x-1}\,dx} and J_{2}(y)=\displaystyle{\int_{1}^{\infty}xy^{x-1}\,dx}. If r<1 then
\int_{0}^{r}xy^{x-1}\,dx=\frac{1}{y}\int_{0}^{r}xy^{x}\,dx \le \frac{1}{y}\int_{0}^{r}x\,dx=\frac{r^{2}}{2y}\le \frac{r^{2}}{2\rho_{1}}, \quad 0<\rho_{1}\le y\le 1.
Therefore J_{1}(y) converges uniformly on [\rho_{1},1]. If x>r>1 and \rho_{2}<1 then
\int_{r}^{\infty}xy^{x-1}\,dx<\int_{r}^{\infty}x\rho_{2}^{x-1}\,dx =\frac{1}{\rho_{2}}\int_{r}^{\infty}x\rho_{2}^{x}\,dx,
Since
\lim_{x\to\infty}\frac{1}{\rho_{2}}\int_{r}^{\infty}x\rho_{2}^{x}\,dx =0
Theorem 7 implies that J_{2}(y) converges uniformly on [0,\rho_{2}]. Therefore, if 0<\rho_{1}<\rho_{2}<1 then \displaystyle{\int_{0}^{\infty}xy^{x-1}} converges uniformly on [\rho_{1},\rho_{2}]. Now Theorem 11 implies that
\tag{A} I'(y)=\int_{0}^{\infty}xy^{x-1}\,dx,\quad 0<y<1.
However, since I(y)=-\displaystyle{\frac{1}{\log y}}, we know that I'(y)=\displaystyle{\frac{1}{y(\log y)^{2}}}. This and (A) imply that \displaystyle{\int_{0}^{\infty}xy^{x}\,dx}=\displaystyle{\frac{1}{(\log x)^{2}}}.
20. Here F(y)=\displaystyle{\int_{0}^{\infty} e^{-x^{2}}\cos 2xy\,dx}, so Theorem 11 implies that
F'(y)=-2\int_{0}^{\infty}xe^{-x^{2}}\sin 2xy\,dx, \tag{A}
since the integral on the right converges uniformly on (-\infty,\infty), by Theorem 6.
Integration by parts yields
\begin{aligned} F(y)&=& =\frac{1}{2y}\int_{0}^{\infty}e^{-x^{2}}(2y\cos 2xy)\,dx\\ &=&\frac{1}{2y}\left(e^{-x^{2}}\sin 2xy\,dx\biggr|_{0}^{\infty} +2\int_{0}^{\infty}xe^{-x^{2}}\sin 2xy\,dx\right) \\ &=&\frac{1}{y} \int_{0}^{\infty}xe^{-x^{2}}\sin 2xy\,dx =-\frac{1}{2y} F'(y).\end{aligned}
From this and (A), F'(y)+2yF(y)=0, so \displaystyle{\frac{F'(y)}{F(y)}}=-2y, \log F(y)=-y^{2}+\log F(0), and F(y)=F(0)e^{-y^{2}}. Since F(0)=\displaystyle{\int_{0}^{\infty}e^{-x^{2}}\,dx}=\displaystyle{\frac{\sqrt{\pi}}{2}} (Example 12), F(y)=\displaystyle{\frac{\sqrt{\pi}}{2}}e^{-y^{2}}.
Here F(y)=\displaystyle{\int_{0}^{\infty} e^{-x^{2}}\sin 2xy\,dx}, so Theorem 11 implies that
F'(y)=2\int_{0}^{\infty}xe^{-x^{2}}\cos 2xy\,dx, \tag{A}
since the integral on the right converges uniformly on (-\infty,\infty). Integrating this by parts yields
\begin{aligned} F'(y) &=&-e^{-x^{2}}\cos 2xy\biggr|_{0}^{\infty}- 2y\int_{0}^{\infty} e^{-x^{2}}\sin 2xy\,dx \\ &=&1-2y F(y),\end{aligned}
so F'(y)+2yF(y)=1, e^{y^{2}}F'(y)+2e^{y^{2}}yF(y)=e^{y^{2}}, and \displaystyle{\left(e^{y^{2}}F(y)\right)'=e^{y^{2}}}. Therefore, since F(0)=0, F(y)=\displaystyle{e^{-y^{2}}\int_{0}^{y}e^{u^{2}}\,du}.
22. Theorems 6 and 11 imply that S and C are n times continuously differentiable on (-\infty,\infty) if \displaystyle{\int_{-\infty}^{\infty}|x^{n}f(x)|\,dx<\infty}.
23. We will show first that
C_{k}(y)=\int_{a}^{\infty} x^{k}f(x) \cos xy \,dx \text{\: and\:} S_{k}(y)=\int_{a}^{\infty}x^{k}f(x)\sin xy\,dx,\: 0\le k\le n,
converge uniformly on U_{\rho}=(-\infty,-\rho]\cup[\rho,\infty) if \rho>0. Note that if \lim_{x\to\infty} x^{n}f(x)=0, then \lim_{x\to\infty} x^{k}f(x)=0, k=0, 1, 2,…n. If 0\le k\le n, then
\tag{A} \int_{r}^{r_{1}}x^{k}f(x)\cos xy\,dx= \frac{1}{y}\left[x^{k}f(x)\sin xy\biggr|_{r}^{r_{1}}- \int_{r}^{r_{1}}(x^{k}f(x))'\sin xy\,dx\right].
Henceforth k is fixed. Our assumptions imply that if \rho>0 and \epsilon>0 then there is an r_{0}\in [a,\infty) such that
\int_{r_{0}}^{\infty}|(x^{k}f(x))'|\,dx<\rho\epsilon \text{\quad and \quad} |x^{k}f(x)|<\rho\epsilon,\quad x\ge r_{0}.
Therefore (A) implies that
\left|\int_{r}^{r_{1}}x^{k}f(x)\cos xy\,dx\right|<3\epsilon,\quad r_{0}\le r<r_{1},\: y\in (-\infty,-\rho]\cup[\rho,\infty).
Now Theorem 4 implies that C_{0}, C_{1},…, C_{k} converge uniformly on (-\infty,-\rho]\cup[\rho,\infty). Since every y\ne0 is in such an interval, Theorem 11 now implies that that if y\ne 0 then
C^{(k)}(y)=\int_{a}^{\infty}x^{k}f(x)\sin xy\,dx,\quad 0\le k\le n.
A similar argument applies to S, S',…S^{(n)}.
Let I(y;r,r_{1})=\displaystyle{\int_{r}^{r_{1}}\frac{1}{x}\sin\frac{y}{x}\,dx}. Assume for the moment that y\ge 0. Substituting u=y/x yields
I(y;r,r_{1})=\int_{y/r_{1}}^{y/r}\left(\frac{u}{y}\right)\sin u \left(-\frac{y}{u^{2}}\right)\,du = \int_{y/r_{1}}^{y/r}\frac{\sin u}{u}\,du.
Therefore, since \displaystyle{\left|\frac{\sin u}{u}\right|}\le 1 for all u, |I(y;r,r_{1})|\le y/r,1\le r\le r_{1}. In fact, since I(-y;r,r_{1})=-I(y;r,r_{1}), we can write |I(y;r,r_{1})\le |y|/r, 1\le r\le r_{1}. Therefore, Theorem 4 implies that \displaystyle{\int_{1}^{\infty}\frac{1}{x}\sin\frac{y}{x}\,dx} converges uniformly on every finite interval.
Now denote F_{r}(y)=\displaystyle{\int_{1}^{r}\cos\frac{y}{x}\,dx}. substituting u=y/x yields F_{r}(y)=y\displaystyle{\int_{y/r}^{y}\frac{\cos u}{u^{2}}\,du}, so \lim_{r\to\infty}F_{r}(y)=\infty for all y\ge 0. Since F_{r}(-y)=F_{r}(y), it follows that \lim_{r\to\infty}F_{r}(y)=\infty for all y, so the answer to the question is “no.”
Let P_{n} be the induction assumption
F^{(n)}(s)=(-1)^{n}\int_{0}^{\infty}e^{-sx}x^{n}f(x)\,dx,\quad s>s_{0},
which is true by the definition of F for n=0. If P_{n} is true, then Theorems 11 and 13 imply that
\begin{aligned} F^{(n+1)}(s)&=&(-1)^{n}\frac{d}{ds} \int_{0}^{\infty}e^{-sx}x^{n}f(x)\,dx=(-1)^{n} \int_{0}^{\infty}\frac{d}{ds}\left(e^{-sx}x^{n}f(x)\right)\,dx\\ &=&(-1)^{n+1}\int_{0}^{\infty}e^{-sx}x^{n+1}f(x)\,dx,\end{aligned}
so P_{n} implies P_{n+1}, which completes the induction proof.
Let G(x)=\displaystyle{\int_{0}^{x}e^{-s_{0}t}f(t)\,dt}. If s>s_{0} then
\tag{A} F(s)=\int_{0}^{\infty}e^{-sx}f(x)\,dx =\int_{0}^{\infty}e^{-(s-s_{0})x}G'(x)\,dx =(s-s_{0})\int_{0}^{\infty}e^{-(s-s_{0})x}G(x)\,dx
(integration by parts). Since \displaystyle{(s-s_{0})\int_{0}^{\infty}e^{-(s-s_{0})x}\,dx=1}, (A) implies that
\tag{B} F(s)-F(s_{0})=(s-s_{0})\int_{0}^{\infty}e^{-(s-s_{0})x}(G(x)-F(s_{0}))\,dx.
Now suppose \epsilon>0. Since F(s_{0})=\displaystyle{\int_{0}^{\infty}e^{-s_{0}t} f(t)\,dt}=\lim_{t\to\infty}G(x), there is an r such that |G(x)-F(s_{0})|<\epsilon if x\ge r; hence, from (B), then
\begin{aligned} |F(s)-F(s_{0})|&\le& (s-s_{0})\int_{0}^{r}e^{-(s-s_{0})x} |G(x)-F(s_{0})|+\epsilon(s-s_{0})\int_{r}^{\infty} e^{-(s-s_{0})x}\,dx\\ &<& (s-s_{0})\int_{0}^{r}e^{-(s-s_{0})x} |G(x)-F(s_{0})|+\epsilon.\end{aligned}
Since r is fixed, we can let s\to s_{0}^{+} to conclude that \limsup_{s\to s_{0}+}|F(s)-F(s_{0})|\le \epsilon, which implies that \lim_{s\to S_{0}+}F(s)=F(s_{0}).
If s\ge s_{1}>s_{0} then
|e^{-sx}f(x)|= |e^{-(s-s_{0})x}e^{s_{0}x}f(x)|\le M e^{-(s-s_{0})x} \le M e^{-(s_{1}-s_{0})x}.
Since
\int_{0}^{\infty}Me^{-(s_{1}-s_{0})x}\,dx=\frac{M}{s_{1}-s_{0}}<\infty,
Theorem 6 implies the stated conclusion.
In Theorem 13 we assumed only that \int_{0}^{x}e^{-s_{0}u}f(u)\,du is bounded; here we are assuming that \int_{0}^{\infty}e^{-s_{0}u}f(u)\,du is convergent.
Let
G(x)=\int_{x}^{\infty}e^{-s_{0}t}f(t)\,dt \text{\quad and\quad} H(x)=\sup\left\{|G(t)|\, \big|\, t\ge x\right\}.
Then
\tag{A} |G(x)|\le H(x)\text{\quad and \quad} \lim_{x\to\infty}G(x)=\lim_{x\to\infty}H(x)=0,
since \int_{0}^{\infty}e^{-s_{0}x}f(x)\,dx converges. Since f is continuous on [0,\infty), G'(x)=-e^{-s_{0}x}f(x). Integration by parts yields
\begin{aligned} \int_{r}^{\infty}e^{-sx}f(x)\,dx&=& \int_{r}^{\infty}e^{-(s-s_{0})x}(e^{-s_{0}x}f(x))\,dx =-\int_{0}^{\infty}e^{-(s-s_{0})x}G'(x)\,dx\\ &=&-e^{-(s-s_{0})x}G(x)\biggr|_{r}^{\infty} +(s-s_{0})\int_{r}^{\infty}e^{-(s-s_{0})x}G(x)\,dx\\ &=&e^{-(s-s_{0})r}G(r)+(s-s_{0})\int_{r}^{\infty}e^{-(s-s_{0})x}G(x)\,dx,\quad s\ge s_{0}.\end{aligned}
Therefore
\begin{aligned} \left|\int_{r}^{\infty}e^{-sx}f(x)\,dx\right|&\le& |G(r)|e^{-(s-s_{0})r}+H(r)(s-s_{0})\int_{r}^{\infty}e^{-(s-s_{0})x}\,dx\\ &=&(G(r)+H(r))e^{-(s-s_{0})r}\le 2H(r)e^{-(s-s_{0})}, \quad s\ge s_{0},\end{aligned}
so (A) implies that F(s) converges uniformly on [s_{0},\infty).
From Theorem 13, F(s)=\displaystyle{\int_{0}^{\infty}e^{-sx}f(x)\, dx} converges for all s>s_{0}. Denote G(x)=\displaystyle{\int_{0}^{x}e^{-s_{0}t}f(t)\,dt}, x\ge 0. Then
\begin{aligned} F(s)&=&\int_{0}^{\infty}e^{-(s-s_{0})x}e^{-s_{0}x}f(x)\,dx= \int_{0}^{\infty}e^{-(s-s_{0})x}G'(x)\,dx \\ &=&(s-s_{0})\int_{0}^{\infty}e^{-(s-s_{0})x}G(x)\,dx\end{aligned}
(integration by parts). Since \displaystyle{(s-s_{0})\int_{0}^{\infty}e^{-(s-s_{0})x}\,dx}=1,
F(s)-F(s_{0})=\int_{0}^{\infty}e^{-(s-s_{0})x}(G(x)-F(s_{0}))\,dx
If \epsilon>0 there is an R such that |G(x)-F(s_{0})|<\epsilon if x\ge R. Therefore, if s>s_{0} then
\begin{aligned} |F(s)-F(s_{0})|&<& (s-s_{0})\int_{0}^{R}e^{-(s-s_{0})x}|G(x)-F(s_{0})|\,dx+\epsilon\\ &<&(s-s_{0})\int_{0}^{R}|G(x)-F(s_{0})|\,dx+\epsilon.\end{aligned}
Hence \limsup_{s\to s_{0}+}|F(s)-F(s_{0})|\le \epsilon. Since \epsilon is arbitrary, this implies that
\lim_{s\to s_{0}+}|F(s)-F(s_{0})|=0.
The assumptions of Exercise 28 imply that \int_{r}^{\infty}e^{-s_{0}x}f(x)\,dx converges for every r>0. Since
\int_{r}^{\infty}e^{-s_{0}x}f(x)\,dx=\int_{0}^{\infty}e^{-s(r+x)}f(x+r)\,dx =e^{-sr}\int_{0}^{\infty}e^{-sx}f(x+r)\,dx,
we can apply the result of Exercise 30 with f(x) replaced by f(x+r), to conclude that
\begin{aligned} \lim_{s\to s_{0}+}\int_{r}^{\infty}e^{-sx}f(x)\,dx&=& e^{-s_{0}r}\int_{0}^{\infty}e^{-s_{0}x}f(x+r)\,dx\\ &=&\int_{0}^{\infty}e^{-s_{0}(x+r)}f(x+r)\,dx\\ &=&\int_{r}^{\infty}e^{-s_{0}x}f(x)\,dx.\end{aligned}
If G(x)=\displaystyle{\int_{0}^{x}e^{-s_{0}t}f(t)\,dt}, then |G(x)|\le M on [0,\infty) for some M. If \epsilon>0, there is an r>0 such that
\tag{A} \int_{0}^{r}e^{-s_{0}x}|f(x)|\,dx <\epsilon.
If s>s_{0}, then
\begin{aligned} \int_{r}^{\infty}e^{-sx}f(x)\,dx&=&\int_{r}^{\infty}e^{-(s-s_{0})x}G'(x)\,dx \\ &=&e^{-(s-s_{0})x}G(x)\biggr|_{r}^{\infty} +(s-s_{0})\int_{r}^{\infty}G(x)e^{-(s-s_{0})x}\,dx\\ &=&-e^{-(s-s_{0})r}G(r) +(s-s_{0})\int_{r}^{\infty}G(x)e^{-(s-s_{0})x}\,dx.\end{aligned}
Therefore, since |G(x)|\le M,
\begin{aligned} \left|\int_{r}^{\infty}e^{-sx}f(x)\,dx\right| &\le&Me^{-(s-s_{0})r}+M(s-s_{0})\int_{r}^{\infty}e^{-(s-s_{0})x}\,dx\\ &=&M\left(e^{-(s-s_{0})r}-e^{-(s-s_{0})x}\biggr|_{r}^{\infty}\right) =2Me^{-(s-s_{0})r}.\end{aligned}
This and (A) imply that
\left|\int_{0}^{\infty}e^{-sx}f(x)\,dx\right|\le \epsilon+2Me^{-(s-s_{0})r}.
Therefore,
\limsup_{s\to\infty} \left|\int_{0}^{\infty}e^{-sx}f(x)\,dx\right|\le \epsilon.
Since \epsilon is arbitrary, this implies that
\limsup_{s\to\infty}\int_{0}^{\infty}e^{-sx}f(x)\,dx=0,
From Exercise 18(d), \displaystyle{\int_{0}^{\infty}\frac{e^{-ax}-e^{-bx}}{x}\sin x\,dx} =\tan^{-1}b-\tan^{-1}a. From Exercise 30, letting b\to\infty yields
\int_{0}^{\infty}e^{-ax}\frac{\sin x}{x}\,dx= \frac{\pi}{2}-\tan^{-1}a, \text{\quad so \bf{(b)}\quad} \int_{0}^{\infty}\frac{\sin x}{x}\,dx=\frac{\pi}{2}.
Integrating by parts yields
\tag{A} \int_{0}^{r}e^{-sx}f'(x)\,dt =e^{-sr}f(r)-f(0) +\int_{0}^{r}se^{-sx}f(x)\,dx.
Suppose s\ge s_{1}>s_{0}. Since |f(x)|\le Me^{s_{0}x}, e^{-sr}|f(r)|\le Me^{-(s_{1}-s_{0})r}. Therefore e^{-sr}f(r)=0 converges uniformly to zero on [s_{1},\infty). Since
\begin{aligned} \left|\int_{r}^{\infty}se^{-sx}f(x)\,dx \right|&\le& M|s|\int_{r}^{\infty}e^{-(s-s_{0})x}\,dx\le \frac{M|s|e^{-(s_{1}-s_{0})r}}{s-s_{0}}\\ &\le&\frac{M(s-s_{0}+|s_{0}|)e^{-(s_{1}-s_{0})r}}{s-s_{0}} \le M\left(1+\frac{|s_{0}|}{s_{1}-s_{0}}\right)e^{-(s_{1}-s_{0})r},\end{aligned}
it follows that \displaystyle{\int_{r}^{\infty}se^{-sx}f(x)\,dx} converges to zero uniformly on [s_{1},\infty). Since this implies that \displaystyle{\int_{0}^{r}se^{-sx}f(x)\,dx} converges uniformly on [s_{1},\infty), (A) implies that G(s) converges uniformly on [s_{1},\infty).
(b) In this case let f'(x)=xe^{x^{2}}\sin e^{x^{2}}, so f(x)=-\displaystyle{\frac{1}{2}}\cos e^{x^{2}}. Since |\cos e^{x^{2}}|\le 1 for all x, the hypotheses stated in (a) hold with s_{0}=0. Therefore G(s) converges uniformly on [\rho,\infty) if \rho>0.
33. We will first show that \displaystyle{\int_{0}^{\infty}e^{-s_{0}x}\frac{f(x)}{x}\,dx} converges. Denote G(x)=\displaystyle{\int_{0}^{x}e^{-s_{0}t}f(t)\,dt}. Since F(s_{0}) is convergent; say |G(x)|\le M, 0\le x<\infty. If 0<r<r_{1} then
\int_{r}^{r_{1}}e^{-s_{0}x}\frac{f(x)}{x}\,dx= \int_{r}^{r_{1}}\frac{G'(x)}{x}\,dx =\frac{G(r)}{r}-\frac{G(r_{1})}{r_{1}}-\int_{r}^{r_{1}}\frac{G(x)}{x^{2}}\,dx.
Therefore
\left|\int_{r}^{r_{1}}e^{-s_{0}x}\frac{f(x)}{x}\,dx\right|\le \frac{3M}{\rho},\quad \rho<r<r_{1}
so Theorem 2 implies that H(s)=\displaystyle{\int_{0}^{\infty}e^{-st}\frac{f(x)}{x}\,dx} converge when s=s_{0}. Therefore Exercise 27 implies that it converges uniformly on [s_{0},\infty), Therefore Theorem 10 implies that
\begin{aligned} \int_{s_{0}}^{s}F(u)\,du &=&\int_{s_{0}}^{s}\left(\int_{0}^{\infty} e^{-ux}f(x)\,dx\right)\,du =\int_{0}^{\infty}\left(\int_{s_{0}}^{s}e^{-ux}\,du\right)f(x)\,dx\\ &=&\int_{0}^{\infty}\left(e^{-s_{0}x}-e^{-sx}\right)\frac{f(x)}{x}\,dx\end{aligned}
From Exercise 30 (with f(x) replaced by f(x)/x), \displaystyle{\lim_{s\to\infty}\int_{0}^{\infty}e^{-sx}\frac{f(x)}{x}\,dx}=0, which implies the stated conclusion.