Crack the Code: Unraveling the Mysteries of the Fredholm Integral Equation

Here is a persuasive advertisement for ghostwriting services specifically for the topic of solving the Fredholm integral equation:

**Unlock the Secrets of the Fredholm Integral Equation with Our Expert Ghostwriting Services**

Are you struggling to solve the Fredholm integral equation? Do you find yourself stuck in a sea of mathematical proofs and iteration schemes? Look no further! Our team of expert ghostwriters specializes in crafting high-quality, step-by-step solutions to this complex problem.

**Our Services:**

* Expertly crafted solutions to the Fredholm integral equation, complete with mathematical proofs and iteration schemes
* Customized to meet your specific needs and requirements
* Delivered on time, so you can meet your deadlines with confidence

**Why Choose Us?**

* Our team of experts has years of experience in mathematics and academic writing
* We use a rigorous quality control process to ensure accuracy and precision
* Our solutions are tailored to your specific needs, ensuring you receive a unique and effective solution

**Don’t Let the Fredholm Integral Equation Hold You Back**

Get the help you need to succeed in your mathematics course. Our ghostwriting services will give you the confidence and expertise you need to tackle even the most complex problems. Contact us today to learn more and take the first step towards academic success!

This advertisement aims to appeal to students who are struggling with the Fredholm integral equation by highlighting the benefits of using a ghostwriting service. The language is concise and easy to understand, and the services and advantages are clearly outlined. The call-to-action at the end encourages students to take the first step towards academic success.

(Hint: note that \(\widetilde{x}\) is a simple zero implies that \(f^{\prime}(x)\neq 0\) on some neighborhood \(U\) of \(\widetilde{x}\) where \(U\subset[a,b]\) and \(f^{\prime\prime}\) is bounded on \(U\)).

11. Show that the Fredholm integral equation \(y(x)=1+\lambda\int_{0}^{1}e^{x-t}y(t)\,dt\) has a solution for \(|\lambda|<\frac{1}{e}\). Set up an iteration scheme to show that

\[\lim_{n\to\infty}y_{n}(t)=1-\frac{\lambda}{1-\lambda}e^{x}\left(\frac{1}{e}- 1\right).\]

12. Let \(k(x,y)\) be a continuous function on the unit square \(0\leq x,y\leq 1\) satisfying \(|k(x,y)|<1\) for all \(x\) and \(y\). Show that there is a continuous function \(f(x)\) on \([0,1]\) such that one has

\[e^{x^{2}}=f(x)+\int_{0}^{1}k(x,y)f(y)dy.\]

Can there be more than one such function \(f\)?

### 2.3 Modes of Convergence

#### Pointwise and Uniform Convergence

Suppose that we are given a sequence of functions \(f_{1},f_{2},\dots\) all defined on a subset \(D\) of \(\mathbb{R}\). For each fixed point \(x_{0}\in D\) we form the numerical sequence \(\{f_{n}(x_{0})\}\). If \(\{f_{n}(x_{0})\}\) converges to \(f(x_{0})\) as sequence of real numbers for every \(x_{0}\in D\), we say \(\{f_{n}\}\)_converges pointwise_ to \(f\) and write \(f_{n}\to f\). We will soon see that desirable properties of the functions \(f_{n}\), such as continuity and integrability, need not be preserved under pointwise convergence. However, under a stronger type of convergence known as _uniform convergence_, the situation is much more positive.

**Definition 74**.: A sequence of functions \(\{f_{n}\}\) is _uniformly convergent_ to \(f\) on the set \(D\) if for each \(\epsilon>0\) there exists a number \(N\) such that \(|f_{n}(x)-f(x)|<\epsilon\) for all \(x\in D\) whenever \(n\geq N\). We use \(f_{n}\rightrightarrows f\) to describe uniform convergence (Figure 2.12).

Note that uniform convergence does depend on the set \(D\) being considered. Convergence could be uniform in one domain, but not on a larger domain. Clearly uniform convergence implies pointwise convergence; however, it imposes a stronger requirement on the number \(N\) so that \(N\) depends only on \(\epsilon\) and not on the particular point \(x_{0}\in D\).

**Example 81**.: The following are two examples of pointwise convergence.

* The functions \(f_{n}(x)=\dfrac{x}{n}\) converge pointwise to \(f(x)\equiv 0\). Clearly these functions are just lines with decreasing slopes, but the convergence is not uniform on \(\mathbb{R}\).
* The functions \(f_{n}(x)=x^{n}\) converge pointwise but not uniformly in the interval \([0,1]\) to the limit \(f(x)=0\) for \(x\in[0,1)\) and \(f(1)=1\). Note that each of the functions \(x^{n}\) is continuous; however, their limit \(f\) is not (Figure 2.13). Furthermore, for \(x>1\), \(\lim\limits_{n\to\infty}x^{n}\) does not exist and therefore for \(x>1\)\(\{f_{n}\}\) does not converge pointwise. However, if we restrict the domain to \([0,1]\), then \[\lim\limits_{n\to\infty}x^{n}=\begin{cases}0&\text{if}\ \ x\in[0,1)\\ 1&\text{if}\ \ x=1\end{cases}.\]So we have pointwise convergence to \(f\), however, \(f\) is not continuous. The following theorem illustrates that continuity is preserved under uniform convergence.

**Theorem 90** (Interchange of limit operations).: _Suppose that \(f_{n}\) for \(n=1,2,\dots\), are defined and continuous on \(D\subset\mathbb{R}\) such that \(f_{n}\rightrightarrows f\) (uniformly) as \(n\to\infty\). Then \(f\) is continuous on \(D\)._

Proof

: Let \(\epsilon>0\) be given and let \(x_{0}\) be any point in \(D\). From

\[|f(x)-f(x_{0})|=|f(x)-f_{n}(x)+f_{n}(x)-f_{n}(x_{0})+f_{n}(x_{0})-f(x_{0})|,\]

the inequality

\[|f(x)-f(x_{0})|\leq|f(x)-f_{n}(x)|+|f_{n}(x)-f_{n}(x_{0})|+|f_{n}(x_{0})-f(x_{0})|\]

holds for all \(x\in D\). Because \(f_{n}\rightrightarrows f\), there exists a positive integer \(N\) depending only on \(\epsilon\) such that \(|f(x)-f_{n}(x)|\leq\frac{\epsilon}{3}\) and \(|f_{n}(x_{0})-f(x_{0})|\leq\frac{\epsilon}{3}\) for all \(x\in D\) whenever \(n\geq N\). Choose \(n\geq N\) and use the fact that \(f_{n}\) is continuous at \(x_{0}\) to conclude \(|f_{n}(x)-f_{n}(x_{0})|\leq\frac{\epsilon}{3}\) holds. Therefore

\[|f(x)-f(x_{0})|\leq\frac{\epsilon}{3}+\frac{\epsilon}{3}+\frac{\epsilon}{3}=\epsilon\]

holds for all \(x\in D\) with \(|x-x_{0}|<\delta\). Thus \(f\) is continuous at \(x_{0}\).

Note that under conditions of the above theorem, when \(\{x_{m}\}\) is any convergent sequence in \(D\) whose limit \(x\) is in \(D\), then by the continuity of \(f_{n}\)’s we have \(\lim\limits_{m\to\infty}f_{n}(x_{m})=f_{n}(x)\), therefore

\[\lim\limits_{n\to\infty}(\lim\limits_{m\to\infty}f_{n}(x_{m}))=\lim\limits_{ m\to\infty}(\lim\limits_{n\to\infty}f_{n}(x_{m})).\]

Thus we see that limit operations can be interchanged.

Figure 2.13: Limit function is not continuous

**Definition 75**.: Let \(C[a,b]\) denote the space of all continuous real-valued functions defined on \([a,b]\). For \(f\in C[a,b]\), define

\[||f||=\sup\{|f(x)|:\ x\in[a,b]\}.\]

We say \(f_{n}\) converges to \(f\) in \(C[a,b]\) if and only if \(||f_{n}-f||\to 0\).

**Theorem 91** (Uniform convergence in the \(\sup\) norm).: \(\{f_{n}\}\) _converges uniformly to \(f\) on \([a,b]\) if and only if \(\{f_{n}\}\) converges to \(f\) in \(C[a,b]\)._

Proof

: Follows directly from definitions.

### Test for Uniform Convergence

**Theorem 92** (Test for Uniform Convergence of Sequences).: _A sequence of functions with domain \(D\) converges uniformly to a function \(f\) if there is a real-valued sequence \(\{a_{n}\}\) such that \(a_{n}\to 0\) and_

\[|f_{n}(x)-f(x)|\leq|a_{n}|\]

_for all \(x\in D\) and for all \(n\in N\)._

Proof

: We know that \(a_{n}\to 0\). So, given \(\epsilon>0\) there exists \(n\in\mathbb{N}\) such that if \(n>N\), then \(|a_{n}|<\epsilon\). This \(N\) is independent of \(x\in D\) and so

\[|f_{n}(x)-f(x)|\leq|a_{n}|<\epsilon\]

for all \(x\in D\) whenever \(n>N\). This implies that \(\{f_{n}\}\) converges uniformly.

**Theorem 93**.: _Let \(\{f_{n}\}\) be a sequence of functions defined on \(A\). Then \(f_{n}\rightrightarrows f\) uniformly on \(B\subset A\)\(\Leftrightarrow\) the sequence of numbers_

\[d_{n}:=\sup\{|f_{n}(x)-f(x)|:x\in B,n\in\mathbb{N}\}\]

_converges to \(0\)._

Proof

: \(\Rightarrow\) Suppose that \(f_{n}\rightrightarrows f\) uniformly. Then, given \(\epsilon>0\) there exists \(N\in\mathbb{N}\) such that \(|f_{n}(x)-f(x)|<\epsilon\) for all \(n>N\) and for all \(x\in B\). So, for \(n>N\), \(d_{n}<\epsilon\). Therefore, \(d_{n}\to 0\).

\(\Leftarrow\) Suppose \(d_{n}\to 0\). Then we know that for sufficiently large \(n\) and for all \(x\in B\),

\[|f_{n}(x)-f(x)|

So \(f_{n}\rightrightarrows f\) uniformly.

**Remark 47**.: Note that Theorems 91, 92, and 93 are several interpretations of the fact that

\[f_{n}\rightrightarrows f\quad\text{uniformly}\quad\text{ if and only if}\quad||f_{n}-f||\to 0.\]

For example for Theorem 92 we can just take \(a_{n}=||f_{n}-f||\).

**Example 82**.: Let \(f_{n}(x)=\dfrac{nx^{2}}{1+nx}\) and \(A=[0,1]\). We claim that \(\{f_{n}(x)\}\) converges uniformly.

Note that \(\lim_{n\to\infty}f_{n}(x)=\lim_{n\to\infty}\dfrac{nx^{2}}{1+nx}=x.\) So, we have

\[|f_{n}(x)-f(x)|=\left|\dfrac{nx^{2}}{1+nx}-x\right|=\left|\dfrac{x}{1+nx} \right|.\]

mathematical notation

评论

发表回复

您的电子邮箱地址不会被公开。 必填项已用 * 标注