title
stringlengths
20
150
upvoted_answer
stringlengths
0
54.3k
Is there a simple example of a transitive vector field on the three sphere?
You can define a nowhere-zero vector field on any odd-dimension sphere. For instance, one is given by embedding the $(2n-1)$-sphere as the unit sphere in $\Bbb R^{2n}$, and at the point $x = (x_1, x_2, x_3,\ldots,x_{2n})\in S^{2n-1}$ define the tangent vector $$ v_x = (x_2, -x_1, x_4, -x_3, \ldots, x_{2n}, -x_{2n-1}) $$ This is readily seen to be orthogonal to $x$, and therefore tangent to the unit sphere at $x$. At the same time, it is never zero, because the origin of $\Bbb R^{2n}$ is not part of our sphere. On an even-dimension sphere, it can never be done, for the same reasons that it cannot be done on $S^2$ in particular. For instance, my proof here (stolen from theorem 2.28 in Hatcher, where my above example for odd-dimensional spheres can also be found) covers all of those cases simultaneously.
A family of polynomials whose roots all have moduli 1
This follows from surprisingly simple properties of the sine function, when we make the substitution $x=e^{it}, t\in[0,2\pi)$. We see that $$ x^{-(k+1)}f_k(x)=k\left(x^{k+1}-\frac1{x^{k+1}}\right)+(k+1)\left(x^k-\frac1{x^k}\right). $$ So with $x=e^{it}$ we see that $f(e^{it})=0$ if and only if $$ g_k(t)=k\sin((k+1)t)+(k+1)\sin(kt)=0. $$ It is easy to check that $f_k(x)$ has a simple zero at $x=1$ and a triple zero at $x=-1$ corresponding to a simple zero of $g_k(t)$ at $t=0$ and a triple zero at $t=\pi$. Let's look at the sign changes in the interval $[0,\pi]$ (the other half is symmetric because $g_k(t)$ is an odd function). Fix an integer $j$ in the range $0<j\le k$. At $t_j:=j\pi/(k+1)$ we have $$ \begin{aligned} g_k(t_j)&=k\sin((k+1)j\pi/(k+1))+(k+1)\sin(kj\pi/(k+1))\\ &=(k+1)\sin(kj\pi/(k+1)). \end{aligned} $$ For this range of $j$ we have $\lfloor\frac{kj}{k+1}\rfloor=j-1$. It follows that $t_j\in ((j-1)\pi,j\pi)$, implying that $g(t_j)$ has the same sign as $(-1)^{j+1}$ for all choices $j=1,2,\ldots,k$. Therefore $g_k(t)$ has a zero in all the intervals $(t_j,t_{j+1})$, $j=1,2,\ldots,k-1$, for a total of $k-1$ zeros in the interval $(0,\pi)$. Together with the $(k-1)$ mirror image zeros in the interval $t\in(\pi,2\pi)$ and the four known zeros (simple at $t=0$ and triple at $t=\pi$) we have accounted for $$ N=1+3+2(k-1)=2k+2=\deg f_k(x) $$ zeros of $f_k(x)$ on the unit circle. The claim follows. In the plot below $k=7$. The more complicated (blue) wave is $g_7(t)$. The orange simple wave is $8\sin 7t$. Its sign determines the sign of $g_7(t)$ at the points $t_1,\ldots t_7$. These are the points in $(0,\pi)$ where the two waves intersect, and are also tickmarked on the horizontal axis.
Converting Between Non-Decimal Bases?
You can use the same process you use for converting to/from base 10 but you will have to do arithmetic in another base. $17_{10}=32_5$, so $0.2E9B03_{17}=\frac{2\cdot 32^5+30\cdot 32^4+14\cdot 32^3+21\cdot 32^2+3}{32^6}_5$ Now multiply out the numerator and do long division in base $5$ to get the part right of the fraction point. For the whole numbers you do the same thing without the division.
Perspective of log-sum-exp as exponential cone
I will write up the solution for the sake of completeness and because the question seems to have appeared elsewhere http://ask.cvxr.com/t/convert-the-perspective-function-of-log-sum-exp-to-cvx/5232 The inequality $$t\geq x_0 f(x/x_0)$$ is $$t\geq x_0\cdot\log(\exp(x_1/x_0)+\cdots+\exp(x_n/x_0)).$$ Since $x_0>0$, dividing by it and then following the same procedure as for log-sum-exp in https://docs.mosek.com/modeling-cookbook/expo.html#log-sum-exp we arrive at the inequality $$1\geq \sum \exp\left(\frac{x_i-t}{x_0}\right)$$ which we can rewrite as $$x_0\geq \sum x_0\exp\left(\frac{x_i-t}{x_0}\right)$$ and eventually $$x_0\geq \sum z_i,\quad z_i\geq x_0\exp\left(\frac{x_i-t}{x_0}\right)$$ with exponential cones on the right-hand side.
Shortest distance from point and line
Calculate the intersection line: $$\begin{cases}x+y&+&z&=0\\x&-&z&=4\end{cases}\implies y=-4-2z\;,\;\;x=4+z$$ and the line is $\;(4,-4,0)+t(1,-2,1)\;,\;\;t\in\Bbb R\;$ , so the distance is $$\frac{||\;\left[(0,1,2)-(4,-4,0)\right]\times\left[(0,1,2)-(5,-6,1)\right]\;||}{||\;(4,-4,0)-(5,-6,1)\;||}=\frac{||\;(-4,5,2)\times(-5,7,1)\;||}{||\;(-1,2,-1)\;||}=$$ $$\frac{||\;(-9,-6,-3)\;||}{\sqrt6}=\frac{\sqrt{126}}{\sqrt6}=\sqrt{21}$$
Probability of selecting elements with replacement
If you're drawing with equal probability and subsequent draws are independent of each other, then the probability that you do not draw, for example, "1" in a single draw from $M$ elements is $\frac{M-1}{M}$. The probability that you do not draw any "1"s in $N$ draws is the product of the individual probabilities, i.e. $\frac{M-1}{M}\cdot\frac{M-1}{M}\cdot\ldots\cdot\frac{M-1}{M}=\left(\frac{M-1}{M}\right)^N = \left(1 - \frac{1}{M}\right)^N$. The probability that you draw at least 1 "1" in the $N$ draws is the complement of the probability that you draw none, i.e. it is $1 - \left(1 - \frac{1}{M}\right)^N$.
Sum of digits divisible by $27$
New answer instead of editing the - already accepted - answer because the argument is significantly different (and much simpler). Let $Q(x)$ denote the digit sum of $x$. Let $r\ge 1$. Then $n=10^r-1=\underbrace{99\ldots 9}_r$ is the smallest $n$ such that among any $n$ consecutive positive integers, at least one has digit sum a multiple of $9r$. That no smaller $n$ works, is immediately clear because in $1,2,3,\ldots, 10^r-2$, all digit sums are $>0$ and $<9r$. Remains to show that in any sequence of $n$ consecutive integers, a digit sum divisible by $9r$ occurs. This is well-known for $r=1$. For $r>1$, consider $n$ consecutive positive integers $$a,a+1,\ldots, a+n. $$ Among the first $9\cdot 10^{r-1}=n-(10^{r-1}-1)$ terms, one is a multiple of $9\cdot 10^{r-1}$. Say, $9\cdot 10^{r-1}\mid a+k=:b$ with $0\le k<9\cdot 10^{r-1}$. Then $Q(b)$ is a multiple of $9$, and as the lower $r-1$ digits of $b$ are zero, we have $Q(b+i)=Q(b)+Q(i)$ for $0\le i<10^{r-1}$ and hence $$Q( b+10^j-1)=Q(b)+9j,\qquad 0\le j\le r-1.$$ (Note that $k+10^{r-1}-1<10^r-1$, so these terms are really all in our given sequence). It follows that $9r$ divides one of these $Q(b+10^j-1)$.
$p \mid q^r+1$, Prove that either $2r\mid p-1$ or $p \mid q^2-1$
$(2r,p-1)$ is not always $2$, since $r$ is a prime and $p-1$ is even, if it is not $2$, it means $r| p-1$, so $2r|p-1$.
Does $\sum_{n=1}^{\infty} \frac1{2^{\sqrt{n}}}$ converge
The integral test tell us that $$\int_{1}^{\infty}\frac{1}{2^{\sqrt{t}}}dt\stackrel{u=\sqrt{t}}{=}2\int_{1}^{\infty}\frac{u}{2^{u}}du $$ $$\stackrel{IBP}{=}\frac{1}{\log\left(2\right)}+\frac{2}{\log\left(2\right)}\int_{1}^{\infty}2^{-u}du=\frac{1+\log\left(2\right)}{\log^{2}\left(2\right)}$$ so the series converges.
Alternative proof to $M$ maximal $\iff R/M$ is a field
Really long indeed! Suppose $M$ is maximal and that $J$ is an ideal of $S=R/M$. Suppose $J$ contains a nonzero element $r+M$. Then $r\notin M$, so $rR+M=R$ and so $1=rx+y$, with $x\in R$ and $y\in M$. Therefore $$ 1+M=rx+y+M=rx+M=(r+M)(x+M)\in J $$ so $J=S$. Suppose $S=R/M$ has only the trivial ideals. Suppose $I$ is an ideal of $R$ with $M\subsetneq I$. Let $r\in I$, $r\notin M$. Then $r+M$ is a nonzero ideal of $S$, hence the ideal it generates is $S$. Therefore $$ 1+M=(r+M)(x+M) $$ for some $x\in R$. This means $1=rx+y$, for some $y\in M$. Since $y\in M\subset I$, we conclude $1\in I$ and so $I=R$.
Meaning of Dot Products in Regards to Linear Algebra
The dot product gives you information about lengths and angles. The dot product of two vectors tells you the angle between them, via the equation $$u \cdot v = ||u|| ||v|| \cos \theta.$$ A consequence of this is that the dot product of a vector with itself tells you the length of the vector, via the equation $$v \cdot v = ||v||^2.$$ If you wanted to, you could actually define length and angle by means of the dot product, using the two equations above.
Integral as infimum of integrals
Not always. For example, if $\mu$ is a Dirac measure at a point $x \neq 0$.
Proving all solutions of an ODE to be periodic
If $\alpha = 0$, we get a linear ODE with constant coefficients whose solutions are $x(t) = A\cos(t) + B\sin(t)$ and so periodic. Assume that $\alpha \neq 0$. Performing the substitution $x'(t) = v(x(t))$, we have $$ x''(t) = v'(x(t))x'(t) = v'(x(t))v(x(t)) $$ and so $$ v'(x)v(x) + x - \sin(\alpha x) = 0 $$ or $$ v \, dv = (\sin(\alpha x) - x) \, dx. $$ Integrating, we get $$ \frac{v^2}{2} + \left( \frac{\cos (\alpha x)}{\alpha} + \frac{x^2}{2} \right) = C. $$ Renaming the constant $C$, we then have $$ F(x,v) = \alpha(x^2 + v^2) + 2\cos(\alpha x) = C $$ and we found a conserved quantity. What does this mean? Given a solution $x(t)$ with initial conditions $x(t_0) = x_0$ and $x'(t_0) = x'_0$, the solution must satisfy $$ F(x(t), v(x(t))) = F(x(t), x'(t)) \equiv F(x_0, x'_0) $$ and so $(x(t),x'(t))$ must lie on a fixed level set of $F$. The only critical point of $F$ is at the origin $(0,0)$ as the equation $$ (\nabla F)(x,v) = 2\alpha(x - \sin(\alpha x), v) = (0, 0) $$ has a unique solution $v = x = 0$ (here we use $0 < \alpha \leq 1$). This critical point is a minimum and all level sets are compact (we can even say that if $C > 2$ then $F^{-1}(C)$ is diffeomorphic to $S^1$) which implies that the solutions exist for all time and must be periodic.
Question about the ring of polynomials bounded on a real variety
I wish I had thought more about this before I posted. Anyway, here is a counterexample. Let $f\in \mathbb{R}[x,y]$ be the polynomial $y(x^2+1)-1$. Let $I$ be the ideal generated by $f$. Let $V$ be the zero-set of $f$ in $\mathbb{R}^2$. Writing the equation $f=0$ in the form $$y=\frac{1}{1+x^2},$$ it is clear that the polynomial $y$ is bounded on $V$. Moreover $y$ is not an element of the ring $\mathbb{R}[\mathbb{R}+I]$ because $y$ is not constant as a function on $V$. We must verify that $I$ is prime and real radical. $I$ is prime: To check this, we note that the quotient ring $\mathbb{R}[x,y]/I$ is isomorphic to the ring $\mathbb{R}[x,1/(1+x^2)]$, which is an integral domain, being a subring of the field $\mathbb{R}(x)$. I is real radical: It is enough to find an order on the ring $\mathbb{R}[x,1/(1+x^2)]$. But the field $\mathbb{R}(x)$ can be ordered: For example think of the elements as functions on $\mathbb{R}$ ordered by eventual domination. It seems that the general problem of calculating the ring of bounded functions on a real variety is much much harder than I realized. If anyone has any references or thoughts on the matter please post.
Bilinear transformation and eigenvalues
We know that if a matrix $\;T\;$ is invertible, then $\;Tu=\alpha u\implies T^{-1}u=\alpha^{-1}u\;$ , so if $\;Bu=\mu u\;$ , then $\;(B\pm I)u=(\mu\pm 1)u\;$ , and from here: $$(B-I)^{-1}u=(\mu-1)^{-1}u\implies$$ so assuming the eigenvector is common we get $$\lambda u=Au=(B+I)(B-I)^{-1}u=(B+I)((\mu-1)^{-1}u)=(\mu-1)^{-1}(B+I)u=\frac{\mu+1}{\mu-1}u$$ Deduce now the final line of the proof.
Constructing a special point in quadrilateral
The ratio $\frac{AD}{BC}$ fixes the scale between the two triangles. So you are looking at a point $M$ such that $$\frac{AM}{BM} = \frac{DM}{CM} = \frac{AD}{BC}$$ The last of these fractions is given from your quadrilateral. The other two restrict $M$ to lie on a certain Apollonian circle. So I'd construct these two circles, and find $M$ as their point of intersection. Note that the triangles are not really similar since they have opposite orientation. Your question seems a bit vague in terms of orientation, hence I took the liberty of interpreting it like this. If you have more strict and different orientation requirements, please make them explicit.
Show that there is a constant $M$ such that for all $x,y \in X$ we have $|f(x) - f(y)| \leq M |x-y| + \epsilon$.
Your proof looks fine, but I think it is more illustrative to prove directly. Choose $\epsilon>0$. Since $X$ is compact, $f$ is uniformly continuous so there is some $\delta>0$ such that if $\|x-y\| < \delta$ then $\|f(x)-f(y)\| < \epsilon$. Now let $M= \max_{\|x-y\| \ge \delta} { \|f(x)-f(y)\| \over \|x-y\|}$. The $\max$ exists because $(x,y) \to { \|f(x)-f(y)\| \over \|x-y\|}$ is continuous on the compact set $\{(x,y) | x,y \in X, \|x-y\| \ge \delta \}$. Then $\|f(x)-f(y)\| \le M\|x-y\| + \epsilon$.
CFG for L = { (+,-)* | #(-) - #(+) ≤ 3 at every position of the word }
Hint: $$ \#(\texttt{-}) - \#(\texttt+) \leq 3 \Leftrightarrow \#(\texttt-) \leq 3 + \#(\texttt+) $$ Let's consider the following languages: (a) $ \#(\texttt-) = 3 + \#(\texttt+) $ (b) $ \#(\texttt-) < 3 + \#(\texttt+) $ If we call $A$ the grammar generated by (a) and $ B $ the grammar generated by (b), then your grammar is: $$ S \to A \mid B $$ A little help: $$ \begin{align*} A &\to K\texttt-K\texttt-K\texttt-K \\ K &\to K\texttt-K\texttt+K\mid K\texttt+K\texttt-K\mid\varepsilon \end{align*} $$ Try to think how you can construct the grammar $ B $ and you're done.
$(X,Y)$ normal, find $P(\mathrm{sign}(X) \neq \mathrm{sign}(Y)$
Hint: you can write $$ Y = \delta X+ \sqrt{1 - \delta^2} Z $$ for some $Z$ independent to $X,Y$, with normal distribution, centered and of variance 1.
Negating a statement
Your answer is "really correct". It is perhaps more usual in English to use "but" rather than "and" in a sentence as the one in your example. Formally (i.e., mathematically), there is no difference.
Singular solution of first order differential equation
$$y = px + p^3\tag1$$which is of Clairaut's form and hence the general solution is $$y=cx+c^3\qquad\text{, where $~c~$ is an arbitrary constant.}\tag2$$ Since the given equation is of Clairaut's form, $p-$discriminant and $c-$discriminant will be exactly same. Hence we will find the only $p-$discriminant. Differentiating $(1)$ by $~p~$, we have $$x+3p^2=0~.\tag3$$ which implies that the domain of the singular solution is $x≤0~$, there should be no complex values in a real problem. The differential equation $(1)$ can be written as, $$y=px+p^3=\frac13p(x+3p^2)+\frac23px=\frac23px\qquad\text{(using $(3)$)}$$ Squaring both side, $$y^2=\frac49p^2x^2= \frac49~x^2~\left(-\dfrac x3\right)=-\frac4{27}~x^3\qquad\text{(using $(3)$)}$$ $$\implies 4x^3+27y^2=0$$ which is the required singular solution. ${}$ thanks to Lutz Lehmann.
Is there any tri-angle ?
Take any triangle $\triangle ABC$ of area $100 \,\text{cm}^2$ with $B,C$ on a fixed line $\mathbf{L}\,$. Flatten the triangle by pushing vertex $A$ down towards $\mathbf{L}$ while sliding $B,C$ outwards along the line as to keep the area constant. In the limit, the distances from any point inside the triangle to either of the three sides tend to $0\,$, so all heights will eventually become $\le 1 \,\text{cm}$ for $\,A\,$ sufficiently close to $\,\mathbf{L}\,$. [ EDIT ]   For a concrete example of such a triangle, consider a triangle $\triangle ABC$ isosceles at $A$ with base $400 \,\text{cm}$ and height $\frac{1}{2} \,\text{cm}\,$, so that thea area is $\,\frac{1}{2} \cdot \frac{1}{2} \cdot 400 = 100 \,\text{cm}^2\,$. Then the equal sides will be $\sqrt{200^2+\left(\frac{1}{2}\right)^2} \,\text{cm}\,$ each, and the corresponding heights are $\,\frac{2 \,\cdot\, 100}{\sqrt{200^2+\left(\frac{1}{2}\right)^2}} \lt 1 \, \text{cm}\,$.
Entropy and Joint Probabilty Distributions for iid random variables
Below is an expanded version of the proof given in Elements Of Information Theory by T.M Cover (2nd Edition, pg 40 - Lemma 2.10.1) for the inequality in question. If $X_1$ and $X_2$ are iid (i.e. $X_1 \sim p(x)$ and $X_2 \sim p(x)$ ) \begin{align} P(X_1 = X_2) &= \sum_{x \in \mathcal{X}} P(X_1 = x , X_2 = x) \\ &= \sum_{x \in \mathcal{X}} P(X_1 = x) P(X_2 = x) \\ &= \sum_{x \in \mathcal{X}} p(x) p(x) \\ &= \sum_{x \in \mathcal{X}} p^2(x) \end{align} Using Jenson's Inequality we know that if a function, $f(y)$, is convex then $\mathbb{E}_{p(x)} \left[ f(y) \right] \geq f\left( \mathbb{E}_{p(x)}[y] \right)$, with equality if and only if $\textbf{y}$ is a constant. The function $f(y) = 2^{y}$ is convex and so if $z = \mathbb{E}_{p(x)} \left[\phi \right]$, then $$f(z) = f(\mathbb{E}_{p(x)} \left[\phi \right]) \leq \mathbb{E}_{p(x)} [f (\phi) ]$$ Where $$\mathbb{E}_{p(x)} [f (\phi) ] = \sum _{x \in \mathcal {X}} p(x) f(\phi)$$ Let $\phi = \log_2 p(x)$. Note that \begin{align} 2^{-H(X)} &= 2^{\mathbb{E}_{p(x)}[\log _2 p(x)]} \\ &= f(y) |_{y = \mathbb{E}_{p(x)}[\log _2 p(x)]} \\ &= f(y) |_{y = \mathbb{E}_{p(x)}[\phi]} \\ &= f(\mathbb{E}_{p(x)}[\phi]) \\ &\leq \mathbb{E}_{p(x)} [f (\phi) ] \\ &= \mathbb{E}_{p(x)} 2^{\phi} \\ &= \mathbb{E}_{p(x)} 2^{\log _2 p(x)} \\ &= \mathbb{E}_{p(x)} p(x) \\ &= \textstyle{\sum} p(x)p(x)\\ &= \textstyle{\sum} p^2 (x) \\ &= P(X_1 = X_2) \end{align} What this means is $P(X_1 = X_2) \geq 2^{-H(X)}$, with equality if and only if $\log p(x)$ is a constant. If $\log p(x)$ is constant then $p(x)$ is also constant and so the distribution in that case is uniform. $\square$
An integral from 0 to infinity
For the evaluation of $$\int_0^\infty \frac{\sin^{2n-1}x}{x}dx\quad\mbox{and}\quad \int_0^\infty \frac{\sin^{2n-1}x\cos x}{x}d x,$$ apply the Power-reduction formulae, $$\sin^{2n-1}x=\frac{1}{4^{n-1}}\sum_{k=0}^{n-1}(-1)^{k}\binom{2n-1}{n+k}\sin((2k+1)x),$$ note that $$\sin((2k+1)x)\cos x=\frac{1}{2}\left(\sin((2k+2)x)+\sin(2kx)\right),$$ and use the fact that for $a>0$, $$\int_{0}^{+\infty}\frac{\sin(ax)}{x}\,dx=\frac{\pi}{2}.$$ Hence $$\int_0^\infty \frac{\sin^{2n-1}x}{x}dx= \frac{1}{4^{n-1}}\sum_{k=0}^{n-1}(-1)^{k}\binom{2n-1}{n+k}\int_0^\infty\frac{\sin((2k+1)x)}{x}dx=\frac{n\pi}{4^n(2n-1)}\binom{2n}{n}$$ and $$\int_0^\infty \frac{\sin^{2n-1}x\cos x}{x}dx=\frac{\pi}{2\cdot4^n(2n-1)}\binom{2n}{n}.$$
Summing a series of integrals
This really isn't so bad. $$\int_0^i\frac{i}{(i+x^2)^{3/2}}dx=\frac{i}{\sqrt{i(i+1)}}.$$ So you're summing: $$\sum_{i=m}^n\frac{i}{i+1}.$$ The latter sum unfortunately doesn't have an explicit form, unless you're willing to use digamma functions.
Prove that all the five sequences converge to the same point $P \in \mathbb{R}^3$.
Partial answer (The estimate below are a bit hard to make precise. Feel free to edit/criticize/downvote): Let $O = \frac{1}{5}(A_n+B_n+C_n+D_n+E_n)$ be the center of mass of the five points $A_n, B_n, C_n , D_n, E_n$. Note that $O$ is really independent of $n$ as the next five points are midpoints of the previous five. By a translation of the initial data, we assume $O$ is the origin. Let $I_n = \max\{|A_n|, |B_n|, |C_n|, |D_n|, |E_n|\}$. It's obiouvs that $I_n\ge I_{n+1}$. To show your claim, it suffices to show that $I_n \to 0$. (Thus $P = O$) One is tempting to show that $I_{n+1} <(1-\epsilon) I_n$ for some $\epsilon$ small independent of $n$. But this cannot be done in general. Indeed we do have Claim: $I_{n+6} < (1-\epsilon) I_n$ for some $\epsilon>0$ small, for all $n$. To show this, WLOG let $I_n = |A_n|$. If $\left|\frac{1}{2}(A_n+ B_n) \right| \le (1-\epsilon)I_n$, we go on to the next step (in the next step there would be at most 4 elemnet with $|\cdot|\ge (1-\epsilon)I_n$). If not, then we can choose $\epsilon$ small so that $$\left|\frac{1}{2}(A_n+ B_n) \right| \ge (1-\epsilon)|A_n|$$ only if $\cos \theta_{A_nB_n} >1-\delta$ and $|B_n| > (1-\delta)|A_n|$, where $\theta_{A_nB_n}$ is the angle between $A_n$ and $B_n$. ($\delta$ here is small, depending on $\epsilon$). Now as $O$ is the center of mass of the five points, $C_n$ will not be closed to $B_n$, and $E_n$ not to $A_n$. We choose $\epsilon>0$ small so that $$\left| \frac{1}{2}(A_n +E_n) \right|, \left|\frac{1}{2}(B_n+ C_n) \right| \le (1-2\epsilon)|A_n|$$ Keep this in mind, we go to the next $(n+1)$ step. Now we have $|A_{n+1}| \ge (1-\epsilon)|I_n|$, $|B_{n+1}|, |E_{n+1}|\le (1-2\epsilon)|I_n|$. So $|A_{n+2}| , |E_{n+2}| \le (1-\epsilon) I_n$. In partucular in the $n+2$ step, at most three elements have $|\cdot | \ge (1-\epsilon) I_n$. Then it is obvious that in the $n+6$ step, we have $$I_{n+6} < (1-\epsilon)I_n$$ (Remark I suspect one can show $I_{n+2} <(1-\epsilon)I_n$)
Opposite effective classes in a Grothendieck group
Here is an example. Let $k$ be a field and let $\mathcal{A}$ be the category of morphisms $f:V\to W$ of $k$-vector spaces such that $\ker f$ and $\operatorname{coker} f$ are both finite-dimensional. By the snake lemma, $\mathcal{A}$ is an abelian subcategory of the category of all morphisms of $k$-vector spaces, and in particular it is an abelian category. Now let $A$ be the object $0\to k$, let $B$ be the object $k\to 0$, and let $C$ be the object $k\stackrel{1}\to k$. Note that there is an infinite direct sum of copies of $C$ in $\mathcal{A}$, so $[C]=0$ in $K_0(\mathcal{A})$. Since there is a short exact sequence $0\to A\to C\to B\to 0$, this means $[A]=-[B]$. However, by the snake lemma, there is a homomorphism $\chi:K_0(\mathcal{A})\to\mathbb{Z}$ which sends $f:V\to W$ to $\dim\ker(f)-\dim\operatorname{coker}(f)$. In particular, $\chi([A])=-1$, so $[A]\neq 0$. (In fact, it is not difficult to see that $\chi$ is an isomorphism.)
Difference between function, derivative and second derivative?
Not really; those examples are unrealistically simple. One of the most useful techniques is to match up zeroes of one graph $-$ places where it crosses the $x$-axis $-$ with ‘flat spots’ $-$ horizontal tangents $-$ of another. If graph $A$ has a zero everywhere that graph $B$ has a horizontal tangent, there’s a good chance that $A$ represents the derivative of the function whose graph is $B$. Of course you should then check further: is $A$ positive $-$ above the $x$-axis $-$ where $B$ is increasing, and negative where $B$ is decreasing? If so you’ve almost certainly found a match: either $A$ is the graph of $f\;'$ and $B$ that of $f$, or $A$ is the graph of $f\;''$ and $B$ that of $f\;'$. Once you’ve matched up one pair like this, the third graph is generally pretty easy to identify. In the example just described, in which $A$ shows the derivative of the $B$ function, there are just two possibilities: either $C$ shows the derivative of the $A$ function, or $B$ shows the derivative of the $C$ function. It shouldn’t be hard to choose between the two by using the same ideas that I mentioned in the first paragraph.
Natural question about harmonic functions
I don't think this is true. Consider, for example, the open cube $Q=(0,1)\times (0,1)$. Define on the boundary of $Q$ a function $v$ by the following: $v(x)=1$ for $x\in Q_1=\{(z,1):\ z\in [0,1]\}$ and $v(x)=0$ on the rest. Consider the problem $$\tag{1} \left\{ \begin{array}{rl} \Delta u=0 &\mbox{ if $x\in Q$} \\ u=v &\mbox{ if $x\in \partial Q$} \end{array} \right. $$ Problem $(1)$ has a unique solution $u$, which is harmonic in $Q$. Now, for each $n=1,2,...$, define a sequence of functions $v_n\in H^{1/2}(\partial Q)\cap C(\partial Q)$ by the following: $v_n(x)\le v_{n+1}(x)$, $v_n(x)=1$ for $x\in \{(z,1):\ z\in [1/2^{n+1},1-1/2^{n+1}]\}$, $0\le v(x)\le 1$ for $x\in \{(z,1):\ z\in [1/2^{n+2},1/2^{n+1}]\cup[1-1/2^{n+1},1-1/2^{n+2}]\}$, $v_n(x)=0$ for $x\in \{(z,1):\ [0,1/2^{n+2}]\cup [1-1/2^{n+2},1]\}$. For each $n$ there is a unique solution $u_n$ to the problem $$ \left\{ \begin{array}{rl} \Delta u=0 &\mbox{ if $x\in Q$} \\ u=v_n &\mbox{ if $x\in \partial Q$} \end{array} \right. $$ The sequence $u_n$ is harmonic in $Q$ and is continous in $\partial Q$. By the maximum principle $u_n(x)\le u_{n+1}(x)$ for all $n$ and $u_n(x)\to u(x)$ for all $x\in \overline{Q}$. Ps: It is possible that the above sequence $u_n$ is not continuous on the boundary, because $Q$ is not regular, however, this can be easily fixed, by considering regular sets, and applying the same reasoning as here, for example, one could consider regularize $Q$ on the corners, or one could consider a circle, and define $v$ to be different whether you take points in the upper half part or in the lower half part of the circle.
How to compute $\lim _{x\to 0}\frac{(1+x)^{1\over x}-e}{x}$ without using a series expansion?
L'Hospital's Rule saves the day. $$\begin{align} \lim_{x\to 0}\frac{(1+x)^{1/x}-e}{x}&=\lim_{x\to 0}\left(\frac{(1+x)^{1/x}}{x+1}\frac{x-(1+x)\log (1+x)}{x^2}\right)\\\\ &=e\lim_{x\to 0}\left(\frac{-\log (1+x)}{2x}\right)\\\\ &=-\frac{e}{2} \end{align}$$
Integral test error approximation
You want $\sum_{n = N+1}^{\infty} \frac{1}{n^4} < .005$. We have the integral comparison $$\int_{N+1}^{\infty} \frac{1}{x^4}< \sum_{n = N+1}^{\infty} \frac{1}{n^4}< \int_{N}^{\infty} \frac{1}{x^4}$$ or $$\frac{1}{3(N+1)^3} < \sum_{n = N+1}^{\infty} \frac{1}{n^4} < \frac{1}{3N^3}$$ The equation $\frac{1}{3N^3} = 0.005$ has the solution $N = 4.05...$ and we see that $$\frac{1}{3 \cdot 5^3} < \sum_{n = 5}^{\infty} \frac{1}{n^4} < \frac{1}{3\cdot 4^3}$$ or $$ 0.00266 < \sum_{n = 5}^{\infty} \frac{1}{n^4} < 0.00520$$ Still that does not guarantee that the remainder starting with the fifth term is less than $0.005$ but we are close. Let's get a better approximation: $$\sum_{n = 5}^{\infty} \frac{1}{n^4} = \frac{1}{5^4} + \sum_{n = 6}^{\infty} \frac{1}{n^4}< \frac{1}{5^4} + \frac{1}{3\cdot 5^3} = 0.0016+ 0.00266... = 0.00426... < 0.005$$ Thus the first such $N$ is $4$ ( anything larger OK too). $\bf{Added:}$ Thanks to the observation of @Claude Leibovici: we check $$\frac{\pi^4}{90} - ( 1+ \frac{1}{2^4} + \frac{1}{3^4} + \frac{1}{4^4}) =0.00357...$$ while $$\frac{\pi^4}{90} - ( 1+ \frac{1}{2^4} + \frac{1}{3^4} ) = 0.00747...$$
Is it true that there is a exact sequence of form $0 \to K \to A^m \to P \to O$ where $K$ is finitely generated?
If $P$ is generated by $p_1,\ldots, p_m$, we obtain a homomorphism $f\colon A^m\to P$ by sending the $(a_1,\ldots, a_m)$ to $a_1p_1+\ldots + a_mp_m$. This homomorphism is onto, giving rise to a short exact sequence $$0\to K\to A^m\to P\to 0 $$ As $P$ is projective, we can pull the identity $P\to P$ to a homomorphism $g\colon P\to A^m$ such that $f\circ g=\operatorname{id}_P$. The sequence is said to split. A sequence split on the right also splits on the left (this holds in more general context): Given $x\in A^n$, we see that $f(x-g(f(x)))=0$, hence $x\mapsto x-g(f(x))$ is a map $A^m\to K$, and is in fact an epimorphism.
Find a basis and dimension for the vector space $V=\bigl\{f(x)\in\mathbb{R}[x];\ \deg f<5,\ f''(0)=f(1)=f(-1)=0\bigr\}.$
Hint: Write $$ f(x)=a_0+a_1x+\cdots+a_4x^4. $$ Find the relations among the $a_i$'s using $$f''(0)=f(1)=f(-1)=0.$$ For instance, what can you tell by $f(1)=0$? Once you have the relations (precisely, linear equations) for these $a_i$'s, you would get a subspace of $\mathbb{R}^5$, the dimension of which is what you are looking for.
Calculate $P(1 \le X \le 2)$
Hint: $P(a \leq X \leq b) = F(b) - lim_{n\rightarrow\infty}F(a-\frac{1}{n})$.
Compactness implies Continuity?
$f(x) = 1$ if $x$ is rational, otherwise $f(x)=0.$ So $f(K)$ is either one or two points.
Boundary layer problem
This is not a boundary layer perturbation problem. The proposed $ϵ=100x^2$ is not a small term around $x=1$. One reduced equation is $$ y'=y^2-2y+1=(y-1)^2 $$ with solution $$ y\equiv 1\text{ or }y=1-\frac{1}{x-c}. $$ Now one can add perturbation terms either to $y$ or to $(y-1)^{-1}$
Probability - Biased coin — Betting game
What you have stated is a version of the classical gamblers ruin problem. There is a good article on that at wikipedia: https://en.wikipedia.org/wiki/Gambler's_ruin Since you put all your wealth at stake in the first bet, it is clear that probability of ruin is larger than $1-p$. (That is the probability of loosing everything on round one, and you can loose it later also).
If G is a finite group and $x \in G$, there is an integer n $\geq 1$ such that $x^n = e$
Suppose that $x \in G$. If $x = e$ then set $n = 1$. Suppose that $x \neq e$. Since $G$ is finite the sequence $x^{n}$ for $n \in \mathbb{N}$ has duplicates. Suppose that $x^{m} = x^{n}$ for distinct $m$ and $n$ with $m &lt; n$. Then $x^{n - m} = e$.
How to evaluate $\int _{-\infty }^{\infty }\!{\frac {\cos \left( x \right) }{{x}^{4}+1}}{dx}$
This can be done using residues, with the function $f(z) = \exp(i z)/(z^4 + 1)$ and a contour that goes along the real axis and returns on a circular arc in the upper half plane.
What Is Exponentiation?
My chief understanding of the exponential and the logarithm come from Spivak's wonderful book Calculus. He devotes a chapter to the definitions of both. Think of exponentiation as some abstract operation $f_a$ ($a$ is just some index, but you'll see why it's there) that takes a natural number $n$ and spits out a new number $f_a(n)$. You should think of $f_a(n) = a^n$. To match our usual notion of exponentiation, we want it to satisfy a few rules, most importantly $f_a(n+m) = f_a(n)f_a(m)$. Like how $a^{n+m} = a^na^m$. Now, we can extend this operation to the negative integers using this rule: take $f_a(-n)$ to be $1/f_a(n)$. then $f_a(0) = f_a(n-n) = f_a(n)f_a(-n) = 1$, like how $a^0=1$. Then we can extend the operation to the rational numbers, by taking $f_a(n/m) = \sqrt[m]{f_a(n)}$. Like how $a^{n/m} = \sqrt[m]{a^n}$. Now, from here we can look to extend $f_a$ to the real numbers. This takes more work than what's happened up to now. The idea is that we want $f_a$ to satisfy the basic property of exponentiation: $f_a(x+y)=f_a(x)f_a(y)$. This way we know it agrees with usual exponentiation for natural numbers, integers, and rational numbers. But there are a million ways to extend $f_a$ while preserving this property, so how do we choose? Answer: Require $f_a$ to be continuous. This way, we also have a way to evaluate $f_a(x)$ for any real number $x$: take a sequence of rational numbers $x_n$ converging to $x$, then $f_a(x)$ is $\lim_{n\to\infty} f_a(x_n)$. This seems like a pretty reasonable property to require! Now, actually constructing a function that does this is hard. It turns out it's easier to define its inverse function, the logarithm $\log(z)$, which is the area under the curve $y=1/x$ from $1$ to $z$ for $0&lt;z&lt;\infty$. Once you've defined the logarithm, you can define its inverse $\exp(z) = e^z$. You can then prove that it has all the properties of the exponential that we wanted, namely continuity and $\exp(x+y)=\exp(x)\exp(y)$. From here you can change the base of the exponential: $a^x = (e^{\log a})^x = e^{x\log a}$. To conclude: the real exponential function $\exp$ is defined (in fact uniquely) to be a continuous function $\mathbb{R}\to\mathbb{R}$ satisfying the identity $\exp(x+y)=\exp(x)\exp(y)$ for all real $x$ and $y$. One way to interpret it for real numbers is as a limit of exponentiating by rational approximations. Its inverse, the logarithm, can similarly be justified. Finally, de Moivre's formula $e^{ix} = \cos(x)+i\sin(x)$ is what happens when you take the Taylor series expansion of $e^x$ and formally use it as its definition in the complex plane. This is more removed from intuition; it's really a bit of formal mathematical symbol-pushing.
Finding Unions and intersections given two probabilities.
Because $\Pr(A\cap B)$ is in this case equal to $\Pr(A)\Pr(B)$, the events $A$ and $B$ are independent, so the answer you got is correct. However, you would not necessarily be viewed as having written a correct solution. If for example $\Pr(A\cap B)=0.1$, multiplying $0.8$ by $0.6$ would yield the wrong answer. I suggest using an argument like this. We have $$\Pr(A^c\cap B)+\Pr(A\cap B)=\Pr(B).$$ We know two of the terms, so we can find the third, which is $0.6-0.12$. Your other two arguments are correct.
Generating function of derangements
You know that $$e^z=\sum_{n\ge 0}\frac1{n!}z^n$$ and that $$D(z)=\sum_{n\ge 0}\frac{d(n)}{n!}z^n\;.$$ Now take the Cauchy product: $$D(z)e^z=\left(\sum_{n\ge 0}\frac{d(n)}{n!}z^n\right)\left(\sum_{n\ge 0}\frac1{n!}z^n\right)=\sum_{n\ge 0}c_nz^n\;,$$ where $$c_n=\sum_{k=0}^n\left(\frac{d(k)}{k!}\cdot\frac1{(n-k)!}\right)=\sum_{k=0}^n\frac{d(k)}{k!(n-k)!}=\frac1{n!}\sum_{k=0}^n\frac{n!}{k!(n-k)!}d(k)\;.$$ What is another notation for $\dfrac{n!}{k!(n-k)!}$? Now use the identity that you were given and one of the first power series that you learned to express $D(z)e^z$ as a rather simple function of $z$, and solve for $D(z)$.
Can we add Lagrangian-like multipliers for a joint constraint $g(x, y)=0$ in a min-max problem?
When solving this kind of problem $$ \min_x \max_y f(x, y) \ \ \text{ s.t. } \ g(x, y) = 0 $$ we are searching for the $f_g(x,y)$ saddle points and those points are handled normally in the Lagrange Multipliers procedures so $$ L(x,y,\mu) = f(x,y)+\mu g(x,y) $$ and the stationary points are determined by $$ \nabla L = 0 $$ regarding only saddle points.
Would stating that $(-\infty, +\infty)= \mathbb{R}$ be correct?
Yeah it is correct. The first notation means all real numbers larger than $-\infty$ and less than $+\infty$, and the second notation means all real numbers, but since any real number is larger than $-\infty$ and less than $+\infty$, it excludes no real numbers.
Vector Space of Polynomial Functions
Suppose that $p_0, p_1, ..., p_m$ are polynomials in $P_m(F)$ such that $p_i(2)=0$ for each $i$. Prove that the set of vectors $p_0, ..., p_m$ is linearly dependent. If each $p_i(t)$ evaluates to $0$ at $t = 2$, why can you conclude that the $p_i$ must therefore be linearly dependent?: For intuition: Try to construct linearly independent polynomials, say for $P_2(F)$, such that for each, $p_i(2) = 0$; and see why these polynomials cannot, in fact, be linearly independent. Suppose there are $m+1$ linearly independent polynomials such that $p_i(2) = 0 \forall p_i$. Then $p_0 = 0$. Hence, e.g., 1 would necessarily be linearly independent of the $m+1$ $p_i$, also in $P_m(F)$. That would give a basis of $P_m(F)$ that has $m+2$ linear independent vectors. Impossible!
Differentiation about the square root of x and y
Implicitly: You'll want to differentiate the function to $x$ $$\frac{\sqrt x}{dx} + \frac{\sqrt y}{dx} = \frac{16}{dx}$$ $$\frac{1}{2\sqrt{x}} + \frac{\frac{dy}{dx}}{2\sqrt{y}} = 0$$ $$\frac{dy}{dx}=-\frac{\sqrt{y}}{\sqrt{x}}=\frac{dy}{dx}=\frac{\sqrt{x}-16}{\sqrt{x}}$$
Property of integral and integrator. (NET June 2011)
1 is definitely false. Think of a sequence of rectangles with decreasing width and increasing height 2 is false. The "typewriter sequence" is the standard counterexample. 3 is false, by the same example as 1. 4 is true (assuming that when you say each function is uniformly bounded, that means that there is a uniform bound for all the functions). This is the bounded convergence theorem.
How can I actually solve this kind of partial differential equations?
You have found one particular solution, and since the PDE is linear any other solution will differ from that one by a solution of the homogeneous equation: $$ x \frac{\partial z}{\partial x}+t \frac{\partial z}{\partial t}+y \frac{\partial z}{\partial y}=0 . $$ This equation says that the directional derivative of $z$ in the radial direction is zero; in other words, $z$ is constant on each ray from the origin. So the general solution of your PDE is $$ z(x,y,t) = \frac{xyt}{3} + g(x,y,t) $$ where $g$ is a function which depends only on the direction of the vector $(x,y,t)$, not on its magnitude. (Expressed in spherical coordinates, $g$ would depend only on the angles $\theta$ and $\phi$, not on the radial variable $r$.) There are some issues to consider if you want a smooth function at the origin ($g$ must be constant in that case), but this is to be expected since the left-hand side of your PDE is zero there.
Determining the relevant set/field for least upper bounds
The fact that $\mathbb{R}$ and $\mathbb{Q}$ are fields doesn't come into play here; what matters is the set we are looking for a supremum in. Given an ordered set $T$, and a subset $S$, the supremum of $S$ in $T$ is the least element of $T$ that is greater than or equal to every element of $S$. That is, $x=\sup(S)$ must satisfy $s\leq x$ for all $s\in S$ $x\leq y$ for all $y\in T$ such that ($s\leq y$ for all $s\in S$) Now, depending on what $T$ is, an element satisfying these properties may or may not exist in $T$. If we consider $S_2$ as a subset of $\mathbb{R}$, there is a supremum for $S_2$: it is $\sqrt{2}$. Either you have not mentioned it, or the book leaves it implicit, but the claim is that $S_2$ has no supremum when we are thinking of it as living in $\mathbb{Q}$, i.e. $S=S_2$ and $T=\mathbb{Q}$. Suppose a rational number $r$ has the property that $s\leq r$ for all $s\in S_2$. If we had $r^2&lt;2$, then $t=2-r^2&gt;0$, and for any $t&gt;0$ there is some $n\in\mathbb{N}$ such that both $\frac{t}{2}&gt;\frac{1}{n^2}&gt;0$ and $\frac{t}{2}&gt;\frac{2}{n}&gt;0$. Then we have $$(r+\tfrac{1}{n})^2=r^2+\tfrac{2}{n}+\tfrac{1}{n^2}&lt;r^2+t=2,$$ so $r$ cannot be a supremum for $S_2$ in $\mathbb{Q}$ because $r+\frac{1}{n}\in S_2$ but $r+\frac{1}{n}\not\leq r$. Thus, any rational number $r$ has the property that $s\leq r$ for all $s\in S_2$ must also have the property that $r^2\geq 2$. We know that equality is impossible, as $\sqrt{2}$ is irrational, so we must have that $2&lt;r^2$. But by a similar argument, there is some $n\in\mathbb{N}$ such that $2&lt;(r-\tfrac{1}{n})^2$, so that $r-\frac{1}{n}$ also has the property that $s\leq r-\frac{1}{n}$ for all $s\in S_2$; but $r-\frac{1}{n}&lt;r$, so $r$ cannot be a supremum for $S_2$. Thus, we have shown that there is actually no supremum for $S_2$, when it is considered as a subset of $\mathbb{Q}$; for any element of $\mathbb{Q}$ that is greater than everything in $S_2$, we can find a strictly smaller element of $\mathbb{Q}$ with that same property. In fact, this argument actually shows that there is no supremum for $S_2$ in any dense subset of $\mathbb{R}$ that is missing $\sqrt{2}$.
Asymptotic behavior of an integrable function
No, you are right there is a counterexample. Join the following dots: $$\left\{\left(n-\frac{1}{2^n},0\right),(n,1),\left(n+\frac{1}{2^n},0\right)\;:\;n\in\mathbb{N}^+\right\}.$$ and you will obtain the graph of a piecewise linear function with the desired property. P.S. Here is an unbounded version: $$\left\{\left(n-\frac{1}{4^n},0\right),(n,2^n),\left(n+\frac{1}{4^n},0\right)\;:\;n\in\mathbb{N}^+\right\}. $$
Prove $4^n-1$ is divisible by $3$, for all $n\in\Bbb N$?
Note that $$4^n-1\equiv(1)^n-1\equiv 0 \mod 3$$
Pigeon holes principle
HINT: Notice that $257=2^8+1$, so you might guess that there will be $2^8$ pigeonholes. You have three atomic sentences. They can have $2^3=8$ different combinations of truth values. How many different truth tables can you build from these $8$ combinations of truth values? Now notice that if $p_1\iff p_2$ is a tautology, then $p_1$ and $p_2$ have the same truth table. (Why?)
linear algebra, show B is a basis for $\mathbb{R}^3$
To show it is a basis, pick a vector ${\bf x}=(x,y,z)$ in the space. What you want to show is that the system $$a_1(1,0,1)+a_2(1,1,2)+a_3(1,2,4)=\bf{x}$$ has always a solution ${\bf v}=(a_1,a_2,a_3)$ regardless of the choice of ${\bf x}\in \Bbb R^3$. Another way would be to show they are linearly independent, for $\dim \Bbb R^3=3$, so three linearly independent vectors will span $\Bbb R^3$. To this end, show that they span the origin only in the trivial way, that is: $$a_1(1,0,1)+a_2(1,1,2)+a_3(1,2,4)=0$$ only when $a_1,a_2,a_3=0$. For the second, just apply the previous solution with the special case ${\bf x}=(3,-1,3)$. NOTE As julien has commented, it suffices to show either that the vectors are linearly independent or that they span $\Bbb R^3$. Why is this so? If the vectors are not linearly independent, they will not span $\Bbb R^3$, and if the vectors are linearly independent, they will span $\Bbb R^3$. This means the two statements are equivalent, and this roots from the fact $\Bbb R^3$ has dimension $3$.
If $c \in \mathbb{R}, c<0, A \subseteq \mathbb{R}$ prove that $\sup(cA) = c \cdot \inf(A)$
Your attempt at a proof is, at the least, insufficient. I suggest you to use less symbols and to spell out what you need to prove. Let $r=\inf(A)$; you want to prove $cr\ge x$, for all $x\in cA$ for all $\varepsilon&gt;0$, there exists $y\in cA$ such that $y&gt;cr-\varepsilon$. First condition. Since $r=\inf(A)$, we know that $r\le a$, for all $a\in A$; if $x\in cA$, then $x=ca$, for some $a\in A$; from $r\le a$ and $c&lt;0$, it follows $cr\ge ca=x$. Second condition. There exists $b\in A$ such that $r-\dfrac{\varepsilon}{c}&gt;b$, because $-\varepsilon/c&gt;0$. Then $$ c\left(r-\frac{\varepsilon}{c}\right)&lt;cb $$ which is the same as $$ cr-\varepsilon &lt; cb $$ and we can take $y=cb\in cA$.
What is the limit behavior of this random sum?
Let $\phi$ be the characteristic function of $X_1$. Then $\phi_n$, the characteristic function of $S_n$, is given by $$\phi_n(t)=\prod_{j=1}^n\phi\left(\frac tj\right),$$ so $(S_n,n\geqslant 1)$ converges in distribution if and only if for each $t\in\mathbb R$, the product $\prod_{j=1}^\infty\phi\left(\frac tj\right)$ is convergent (in the usual sense or $\phi(t/j)=0$ for some $j$). There are case where the product is convergent, and when the product is divergent. For example, consider random variables with characteristic function $\phi(t):=e^{-|t|^\alpha}$, where $0\lt\alpha\lt 2$. Then the product is convergent if and only if $\alpha\gt 1$. If $S_n$ converges in distribution to $0$ then $|\phi(t)|=1$ for each $t$.
How to proof that the $\mathbb{Z}$-span of weights of a faithful $L$-modul contains the root lattic?
$V$ being a faithful ${\mathfrak g}$-module means that the structure homomorphism $\rho: {\mathfrak g}\to{\mathfrak g}{\mathfrak l}(V)$ is injective. By definition, $\rho$ is a homomorphism of Lie algebras, but we may also view it as a morphism of ${\mathfrak g}$-modules if we equip ${\mathfrak g}$ with the adjoint action and ${\mathfrak g}{\mathfrak l}(V)$ with the ${\mathfrak g}$-action given by $X.- := [\rho(X),-]$. Now your claim follows from the following observations: Firstly, by the injectivity of ${\mathfrak g}\to {\mathfrak g}{\mathfrak l}(V)$ any root of ${\mathfrak g}$ (i.e. any weight of the adjoint representation of ${\mathfrak g}$) is a weight of ${\mathfrak g}{\mathfrak l}(V)$. Secondly, ${\mathfrak g}{\mathfrak l}(V)\cong V\otimes_{\mathbb k} V^{\ast}$ as ${\mathfrak g}$-modules, and weights add upon tensoring. Can you fill in the details yourself?
On the image of an embedding of $\mathbb{R}$ in $\mathbb{R}^3$
$g_1(x)=x_2-x_1^2$ and $g_2(x)=x_3-x_1^3$ do that.
Find the area included between an arc of cycloid $x=a(\theta - \sin (\theta))$, $y=a(1-\cos (\theta))$ and its base.
There are different ways to use integration to compute an area. What you seem to have overlooked is that you are given a parametric equation for the cycloid, namely $$\begin{align*} x(\theta) &amp;= a (\theta - \sin \theta) \\ y(\theta) &amp;= a (1 - \cos \theta) \end{align*}$$ which, for a fixed constant $a$, relates the $(x,y)$ coordinate of a point on the cycloid to a parameter $\theta$. Therefore, the expression for the area $$\int y \, dx$$ that you cite is not an integral with respect to $x$, but actually with respect to $\theta$. Written out completely, it actually means $$A = \int_{\theta = \theta_0}^{\theta_1} y(\theta) \, \frac{d}{d\theta}[x(\theta)] \, d\theta.$$ In your case the interval of integration over half a period is $\theta \in [0, \pi]$ as the text states.
$X$ a normed space. $(x_n) \subset X$ s.t. $|f(x_n)| < M_f$ for all $n$ and $\forall f \in X^{\ast}$ then there is $C$ s.t. $\|x_n\| \le C$
This is just the Banach-Steinhaus Theorem applied to the dual space $X^*$ which is always a Banach space, even if $X$ is only a normed space. For $x \in X$ let $\hat x \in X^{**} = (X^*)^*$ be defined via $\hat x(f) = f(x)$ for all $f \in X^*$. Then, for all $n \in \mathbb{N}$ and $f \in X^*$ we have $$ |\hat{x_n}(f)| = |f(x_n)| \leq M_f, $$ i.e. the family of operators $(\hat{x_n})_{n \in \mathbb{N}}$ is pointwise bounded. Banach-Steinhaus gives you that the family is uniformly bounded, i.e. there exists some $C &gt; 0$ such that $$ \|\hat{x_n}\| \leq C, $$ for all $n \in \mathbb{N}$. But $X$ is isometrically isomorphic to a subspace of its Bi-dual, i.e. $\|\hat{x_n}\| = \|x_n\|$.
Linear algebra with a transformation i think
If $P$ is of the Jordan Canonical Form, then it is easy to verify that $P^2=P$ implies (2). In addition, you can verify that $P^2=P$ if and only if $P$ is a diagonal matrix with eigenvalues in $\{0,1\}$.
What is a monotonic increasing function $[0,1] \to [0,1]$?
As André Nicolas wrote in the comments: There are many choices, for example $x^\alpha$ for suitable $\alpha$, which could be constant but need not be. For your example, $\alpha=1/2$ will give a sort of OK fit.
The greatest integer function
For $x\in [2,6)$, $x\to 3x^2$ is an increasing function and it attains the values in $[3\cdot 2^2,3\cdot 6^2)=[12,108)$. For any integer $k\in [12,107]$, let $$I_k=\{x\in [2,6): k\leq 3x^2&lt;k+1\}=[\sqrt{k/3},\sqrt{(k+1)/3}).$$ Note that if $x\in I_k$ then $\lfloor 3x^2\rfloor=k$ which means that $\lfloor 3x^2\rfloor$ is constant on each interval $I_k$. Moreover $|I_k|=\sqrt{(k+1)/3}-\sqrt{k/3}$ where $|I_k|$ is the length of $I_k$. Hence by the definition of integral, $$\int_{2}^6 \lfloor 3x^2\rfloor dx=\sum_{k=12}^{107} k|I_k| =\frac{1}{\sqrt{3}}\sum_{k=12}^{107} k(\sqrt{k+1}-\sqrt{k}) \approx 206.005$$ where $|I_k|$ is the length of the interval $I_k$. P.S. For general information about the floor function take a look HERE.
Properties of Variance for Random Variables
Variances always add, so the answer should be $16+81=97$.
Will there be a surjective homomorphism...?
Hint: the two parts are equivalent. If such a homomorphism exists, then its kernel is a subgroup of $\mathbb{Q}$ of index $2$. But the homomorphism does not exist. For choose some element $ x\in \mathbb{Q}$ that gets mapped to $1$. What does $\frac{x}{2}$ get mapped to?
2011 IMC Section A Problem 3
So we have $$...=1!^23!^2...2009!^2\cdot 2011!^2 \cdot 2\cdot 4\cdot 6...\cdot 2010\cdot 2012$$ $$=1!^23!^2...2009!^2\cdot 2011!^2 \cdot 2^{1006}\cdot \color{red}{1006!}$$
Show that the last component of an eigenvector of a tridiagonal hermitian matrix is not 0.
Let $A$ be a Hermitian tridiagonal matrix of order $n$, with $a_{i, i + 1} \ne 0$, $i = 1, \ldots, n - 1$. Write $A = \begin{bmatrix} A^{(1)} &amp; y\\y^* &amp; a\end{bmatrix}$, as you have defined, and suppose that $x = \begin{bmatrix}x^{(1)} \\ 0\end{bmatrix}$ is an eigenvector. Then $$Ax = \lambda x \implies \begin{bmatrix} A^{(1)} &amp; y\\y^* &amp; a\end{bmatrix} \begin{bmatrix}x^{(1)} \\ 0\end{bmatrix} = \lambda \begin{bmatrix}x^{(1)} \\ 0\end{bmatrix} \implies \begin{array}{l} A^{(1)} x^{(1)} = \lambda x^{(1)}\\ y^*x^{(1)} = 0\end{array}.$$ Since $A$ is tridiagonal with $a_{n,n+1} \ne 0$, $y^* = \begin{bmatrix}0 &amp; \cdots &amp; 0 &amp; a^*_{n,n+1}\end{bmatrix}$, so that $y^*x^{(1)} = 0 \implies x^{(1)}_{n-1} = 0$. Now, $A^{(1)}$ is a Hermitian tridigonal matrix of order $n - 1$, with $a^{(1)}_{i, i+1} \ne 0$, $i = 1, \ldots, n - 2$, and having $x^{(1)} = \begin{bmatrix}x^{(2)} \\ 0 \end{bmatrix}$ as an eigenvector. Thus, by induction, we find that every component of $x$ is zero, which contradicts the assumption that $x$ is an eigenvector. Thus, any eigenvector of $A$ must have a non-zer0 $n$th component.
Hessian matrix of the mahalanobis distance wrt the Cholesky decomposition of a covariance matrix
Let $u=x-\mu$, $f:\Lambda\rightarrow\Lambda\Lambda^T$, $g:\Sigma\rightarrow u^T\Sigma^{-1}u$ and $h=g\circ f$. Then $Dg_{\Sigma}: H\in S_k\rightarrow -u^T\Sigma^{-1}H\Sigma^{-1}u$ and $Df_{\Lambda}:K\in T_k\rightarrow K\Lambda^T+\Lambda K^T$, where $S_k$ is the set of symmetric matrices and $T_k$ is the set of lower triangular matrices. Thus $Dh_{\Lambda}:K\rightarrow -u^T\Sigma^{-1}(K\Lambda^T+\Lambda K^T)\Sigma^{-1}u=-2tr(\Lambda^T\Sigma^{-1}uu^T\Sigma^{-1}K)=-2&lt;\Sigma^{-1}uu^T\Sigma^{-1}\Lambda,K&gt;$ where $&lt;Y,Z&gt;$ is the real scalar product $tr(Y^TZ)$. Let $\nabla (h)_{\Lambda}=-2\Sigma^{-1}uu^T\Sigma^{-1}\Lambda$; we'll say that the previous function is the gradient of $h$ because, for every $i\geq j$, $\dfrac{\partial h}{\partial \Lambda_{ij}}=\nabla(h)_{i,j}$; the strictly upper part of $\nabla (h)$ is useless. The second derivative is the following symmetric bilinear form: $D^2h_{\Lambda}:(K,L)\in T_k\times T_k\rightarrow 2u^T\Sigma^{-1}(L\Lambda^T+\Lambda L^T)\Sigma^{-1}(K\Lambda^T+\Lambda K^T)\Sigma^{-1}u-u^T\Sigma^{-1}(KL^T+LK^T)\Sigma^{-1}u$. For example, if $p\geq q,r\geq s$, then $\dfrac{\partial^2 h}{\partial \Lambda_{pq}\partial\Lambda_{rs}}=D^2h_{\Lambda}(E_{pq},E_{rs})$. Note that the Hessian of $h$ is a complicated tensor .
solving laplace equation with separation of variables
The function $u=0$ cannot be a solution of the problem because of the boundary conditions (which are not all zero on the boundary). In the first boundary conditions the first $u(\pi, y)=0$ must be replaced by $u(x,\pi)=0$ in order to have zero boundary conditions for $y$. After coorecting this and applying the separation of variables we get: $$u(x,y)=\sum_{n=1}^{\infty}\big(c_{n}\sinh(nx)+d_n\cosh(nx)\big)\sin(ny).$$ Now using the othogonality of the functions $\sin(ny)$ and BC's for $x$ we will determine $c_n$ and $d_n$. $$u(0, y)=0\Longrightarrow d_n=0.$$ $$u(\pi, y)=\sin y\cos y=\frac{1}{2}\sin(2y)=\sum_{n=1}^{\infty}c_n\sinh(n\pi)\sin(ny).$$ Orthogonality of $\sin(ny)$ implies that $$\int_0^\pi\sin(2y)\sin(ny)dy=\pi c_n\sinh(n\pi).$$ In this case we should be more careful, because the integral on the LHS is not zero for all $n$. It is zero (by othogonality) for all $n$ except for $n=2$. When $n=2$, $$\int_0^\pi\sin^2(2y)dy=\frac{\pi}{2}$$ and hence $c_2=\displaystyle\frac{1}{2\sinh(2\pi)}$. Therefore, $$u(x,y)=\frac{\sinh(2x)\sin (2y)}{2\sinh(2\pi)}.$$
$ \int_{0}^{88} \sin \sqrt{x} $ dx, $n=4$ . Use the Midpoint Rule with the given value of $n$ to approximate the integral.
The default assumption in calculus is that angles, unless otherwise specified, are measured in radians. If that is the case, put your calculator in radian mode amd calculate $$\frac{88}{4}\left(\sin(\sqrt{11})+\sin(\sqrt{33})+\sin(\sqrt{55})+\sin(\sqrt{77})\right).$$ To $5$ decimal places, I get $18.12003$. Rounded to say $4$ decimal places this is $18.1200$. Remark: Weird problem, particularly using the default assumption that we are using radians. With $n=4$, one cannot expect a numerical method such as Trapezoidal Rule, Midpoint Rule, or Simpson's Rule, to produce even a crude estimate of the integral.
Find a 2nd independent solution of the given equation (given one solution)?
Write the differential equation in the form $ y''+P(x)y'+Q(x)y=0$ Since $x&gt;1 \implies x \neq 1$ so $$y''-\frac x {x-1}y'+\frac y {x-1}=0$$ $$ \text {1)} \left (\frac {y_2} {y_1}\right )'=\frac {W(y_1,y_2)} {y_1^2}$$ $$\frac {y_2'y_1-y_1'y_2}{y_1^2}=\frac {W(y_1,y_2)} {y_1^2}$$ $$\frac {y_2'e^x-e^xy_2}{e^{2x}}=\frac {W(e^x,y_2)} {e^{2x}}$$ $${y_2'e^x-e^xy_2}= {W(e^x,y_2)}$$ $${y_2'e^x-e^xy_2}=\left| \pmatrix{e^x &amp; y_2\\ e^x &amp; y'_2} \right|$$ $${y_2'e^x-e^xy_2}= e^xy'_2-e^xy_2$$ 2)Then use Abel's identity ... with $P(x)=-\dfrac x{x-1}$ $$W=W_0e^{-\int P(x)dx}=W_0e^{\int \frac x {x-1}}=Ke^{\int 1+\frac 1 {x-1}}=Ke^xe^{ln|x-1|}=Ke^x(x-1) \text { since x&gt;1}$$ $$W= e^xy'_2-e^xy_2 \implies y'_2-y_2=K(x-1)$$ I let you finish...solve for $y_2$ the equation.It's a first order linear equation. Another way to integrate this equation... $$(x-1)y''-xy'+y=0$$ $$(x-1)y''-xy'\color{red}{+y'-y'}+y=0$$ $$(x-1)y''\color{blue}{-xy'+y'}+\color{green}{y-y'}=0$$ $$(x-1)y''\color{blue}{-(x-1)y'}+y-y'=0$$ $$(x-1)(y''-y')-(y'-y)=0$$ Substitute $z=y'-y$ $$(x-1)z'-z=0$$ Integrate.. $$\int \frac {dz}{z}=\int \frac {dx}{x-1}$$ $$z=K_1(x-1)$$ $$y'-y=K_1(x-1)$$ $e^{-x}$ as inetgrating factor $$(ye^{-x})'=e^{-x}K_1(x-1)$$ $$ye^{-x}=\int e^{-x}K_1(x-1)dx+K_2$$ $$y(x)=K_1e^{x}\int e^{-x}(x-1)dx+K_2e^x$$ $$\boxed{y(x)=K_1x+K_2e^x}$$
Finding the derivative of an integral function
Since $$y = c e^{-x} + e^{-x} \int_0^x \frac{\tan t}{t} \, dt,\tag1$$ as you correctly show $$y' = -ce^{-x} + e^{-x} \frac{\tan x}{x} - e^{-x} \int_0^x \frac{\tan t}{t} \, dt.\tag2$$ Now if we rearrange (1) we have $$e^{-x} \int_0^x \frac{\tan t}{t} \, dt = y - c e^{-x},$$ so substituting this result into (2) we have $$y' = -c e^{-x} + e^{-x} \frac{\tan x}{x} - (y - ce^{-x}),$$ or $$y' + y = \frac{e^{-x} \tan x}{x},$$ a first-order differential equation that is free from any integral sign and has (1) as a solution.
A linear code must be able to send $256$ different messages s.t. it corrects one error. What is the least possible length of such a code?
Welcome to MSE! I'd consider the Hamming bound, $$q^n \geq |C|\sum_{i=0}^t {n\choose i}(q-1)^i,$$ where $|C|=2^8$, $q=2$, $d=2t+1=3$, i.e., $t=1$. Search for the smallest block length $n\geq 8$ such that this inequality is satisfied.
Finding the volume of a cone which is partially filled with water
Let $r_1,r_2$ be the radii of the upper circular water surface when the cone is base-down/base-up respectively, $r$ the cone's radius and $h$ the cone's overall height. Then we have, by similarity relations, $$\frac{h-h_{w_1}}{r_1}=\frac hr=\frac{h_{w_2}}{r_2}$$ $$\frac{r(h-h_{w_1})}h=r_1,\frac{rh_{w_2}}h=r_2\tag1$$ Since the volume of water is unchanged, $$\frac\pi3(r^2+rr_1+r_1^2)h_{w_1}=\frac\pi3r_2^2h_{w_2}$$ $$(r^2+rr_1+r_1^2)h_{w_1}=r_2^2h_{w_2}$$ Substituting the relations in $(1)$, $$\left(r^2+r\left(\frac{r(h-h_{w_1})}h\right)+\left(\frac{r(h-h_{w_1})}h\right)^2\right)h_{w_1}=\left(\frac{rh_{w_2}}h\right)^2h_{w_2}$$ Dividing by $r^2$ and then multiplying by $h^2$, we get a quadratic in $h$: $$3h_{w_1}h^2-3h_{w_1}^2h+h_{w_1}^3-h_{w_2}^3=0$$ Thus we can solve for $h$. If we are given only $h_{w_1}$ and $h_{w_2}$, then we are stuck; we cannot find $r,r_1,r_2$ even though they are now in known proportions. For example, two cones with $h=3$ and $r=1,2$ filled so that $h_{w_1}=1$ in both cases will have the same $h_{w_2}$. If we are also given $r=h$, then the volume is simply $\frac\pi3h^3$ using the $h$ we have calculated. (Knowing that $r=h=a$ and the value of $a$ makes the problem trivial.)
Proving that $\frac{ \cos (A+B)}{ \cos A-\cos B}=-\cot \frac{A-B}{2} \cot \frac{A+B}{2}$
First, note that $\dfrac{\cos(A + B)}{\cos A - \cos B} \neq -\cot \left( \dfrac{A-B}{2} \right) \cot \left(\dfrac{A+B}{2} \right)$ In particular, if we use something like $A = \pi/6, B = 2\pi/6$, then the left is $0$ as $\cos(\pi/2) = 0$ and the right side is a product of two nonzero things. I suspect instead that you would like to prove: $$\dfrac{\cos A + \cos B}{\cos A - \cos B} = -\cot \left( \dfrac{A-B}{2} \right) \cot \left(\dfrac{A+B}{2} \right)$$ HINTS And this looks to me like an exercise in the sum-to-product and product-to-sum trigonometric identities (wiki reference). In fact, if you just apply these identities to the top and the bottom, you'll get the result.
If $M$ can be immersed in $\mathbb{R}^k\times(S^1)^l$ with codimension $1$, can it be immersed in $\mathbb{R}^{k+l}$ with codimension $1$?
Note that $\mathbb{R}\times S^1$ embeds in $\mathbb{R}^2$ via the map $(x, z) \mapsto e^xz$. Repeated application of this map gives an embedding $\varphi : \mathbb{R}^k\times(S^1)^l \to \mathbb{R}^{k+l}$. If $f : X \to \mathbb{R}^k\times(S^1)^l$ is an immersion, then $\varphi\circ f : X \to \mathbb{R}^{k+l}$ is also an immersion (because $\varphi$ is an embedding). Moreover, the codimension of the immersion $f$ is equal to the codimension of the immersion $\varphi\circ f$ (because $\dim \mathbb{R}^k\times(S^1)^l = \dim \mathbb{R}^{k+l}$).
How do I prove a sequence is Cauchy
As an easier example of how to apply the definition of a Cauchy sequence, define the sequence $\{\frac{1}{n}\}$. Given any $\epsilon&gt;0$, you would like to find an $N$ such that for any $n,m&gt;N$, $\left|\frac{1}{n}-\frac{1}{m}\right|&lt;\epsilon$. This would certainly be the case if $\frac{1}{n},\frac{1}{m}&lt;\frac{\epsilon}{2}$ since $\left|\frac{1}{n}-\frac{1}{m}\right|\leq\left|\frac{1}{n}\right|+\left|\frac{1}{m}\right|$. Therefore you need to force $n,m&gt;\frac{2}{\epsilon}$. Conveniently enough, you are allowed to choose whichever $N$ you like, so choosing $N&gt;\frac{2}{\epsilon}$ will do the trick. The above constitutes the work you do beforehand, now the proof. Claim: The sequence $\{\frac{1}{n}\}$ is Cauchy. Proof: Let $\epsilon&gt;0$ be given and let $N&gt;\frac{2}{\epsilon}$. Then for any $n,m&gt;N$, one has $0&lt;\frac{1}{n},\frac{1}{m}&lt;\frac{\epsilon}{2}$. Therefore, $\epsilon&gt;\frac{1}{n}+\frac{1}{m}=\left|\frac{1}{n}\right|+\left|\frac{1}{m}\right|\geq\left|\frac{1}{n}-\frac{1}{m}\right|$. Thus, the sequence is Cauchy as was to be shown.
Find the number of sides of a polygon where it satisfies the relation $\frac{1}{A_1A_2} =\frac{1}{A_1A_3} + \frac{1}{A_1A_4}$
Note that if the polygon $A_1A_2 \ldots A_{n}$ is a regular polygon, then the quadrilateral $A_1A_2A_3A_4$ is a cyclic quadrilateral. Furthermore, even $A_1A_3A_4A_5$ is a cyclic quadrilateral. We will shorten this to $ABCDE$ for convenience of avoiding subscripts. We use the famous Ptolemy's theorem: In a cyclic quadrilateral $ABCD$, $AB\cdot CD + BC \cdot AD = AC \cdot BD$. Let : $AB=BC=CD=DE=a$, $AC = BD = CE = b$, $AD = BE =c$ ,and $AE=d$. Apply Ptolemy to $ABCD$: $a^2 + ac = b^2(1)$. Next, apply Ptolemy to $ACDE$ : $ab+ad =bc(2)$. Apply Ptolemy to $ABDE$ : $a^2 + bd=c^2(3)$. Furthermore, we are given that $\dfrac{1}{a} = \dfrac{1}{b} + \dfrac{1}{c}(4)$. Solve for $d$ from $2$: $$ d = \frac{bc-ab}{a} $$ Substitute this value of $d$ in $3$: $$ a^2 + b\left(\frac{bc-ab}{a}\right) = c^2 \implies a^2 + \left(\frac{b^2c-ab^2}{a}\right) = c^2 $$ now, use $4$: $$ a^2 + (b^2c-ab^2)\left(\frac{1}{b} + \frac{1}{c}\right) = c^2 $$ On simplification: $$ a^2 + bc = c^2 $$ But, by $2$, $$ bc=bd \implies c=d $$ Hence, the shape has exactly two distinct diagonal lengths, hence there are four diagonals from $A$ to other vertices, two of length $b$ and two of length $c$. That means that the number of vertices is $4+3=7$, where the $3$ is for $A$ and it's adjacent vertices. Hence, the shape is that of a regular heptagon. If the shape is not regular, nothing definitive can be said about the polygon. By the way, if this would have been a regular polygon such that $A_1A_2 + A_1A_3 = A_1A_6$, then the regular polygon is a nonagon (again, Ptolemy's theorem is useful).
Is a forest a tuple, a set or both?
Wiki: "A forest is an undirected graph in which any two vertices are connected by at most one path" The important thing is that this forbids cycles. Now, assume V1=V2={A,B,C} and V3={}. Furthermore, let E1={AB,BC} and E2={AC,BC} (E3={}, of course). In both definitions, you produce the cycle AB, BC, CA, i.e. not a forest. However, note that this depends how exactly you interpret (i) as a graph (this is a bit of an unusual notation... thinking of it, defining a graph as a set with 3 elements seems weird...I strongly advise against it).
Proof of LCM property.
Yes: $a\mid ak\implies a\mid\operatorname{lcm}(ak,cm)$. For the same reason, $c\mid\operatorname{lcm}(ak,cm)$. Therefore, $\operatorname{lcm}(a,c)\mid\operatorname{lcm}(ak.cm)$.
discrete version of Laplacian
Here is a low-brow answer. If you accept that the discrete version of the second derivative of a function $f$ is $f(x+1) - 2f(x) + f(x-1)$, then the discrete Laplacian, say in two dimensions, is $\Delta f(x, y) = \frac{f(x+1, y) + f(x-1, y) + f(x,y+1) + f(x,y-1)}{4} - f(x,y)$. But the first term is just the transition probabilities of a random walk on $\mathbb{Z}^2$ where one moves to each of the four horizontally or vertically adjacent neighbors with equal probability. A similar statement is true in $n$ dimensions. More generally one can define a discrete Laplacian on any graph which mimics the usual Laplacian. In fact there is an entire textbook by Doyle and Snell dedicated to working out potential theory on graphs.
Questions regarding mutual independence of events
Yes it is possible. See https://en.wikipedia.org/wiki/Pairwise_independence for a standard example. Yes it is possible, let $A_1 = \emptyset$ and the statement holds regardless of the rest of the $A_i$, which do not have to be independent. No this does not imply mutual independence, or even pairwise independence. If $A_1$ and $A_2$ are disjoint, they cannot be independent unless at least one of them has probability zero. Intuitively, if I told you that $A_1$ occurred, then you know for certain that $A_2$ did not occur.
What series does the function $\frac{1}{(1-ax)^r}$ generate?
The series is $$S=(1-ax)^{-r}=\sum_{k=0}^{\infty}(-1)^k {-r \choose k} (ax)^k=\sum_{k=0}^{\infty} {r+k-1 \choose r - 1} (ax)^k.$$ Which is valid for $|x|&lt;a^{-1}.$
If $2AB = (BA)^2+I_n$ then $1$ is an eigenvalue for $AB$
The condition implies that $BA$ commutes with $AB$ and hence they are simultaneously triangularisable over $\mathbb C$. Let $AB$ and $BA$ be simultaneously triangularised. Since $AB$ and $BA$ in general have identical spectra, if $\lambda_1,\ldots,\lambda_n$ are the eigenvalues of $BA$ along its diagonal, then the entries along the diagonal of $AB$ are $\lambda_{\sigma(1)},\ldots,\lambda_{\sigma(n)}$ for some permutation $\sigma$. So, the given condition implies that $f(\lambda_i)=\lambda_{\sigma(i)}$ where $f:z\mapsto (z^2+1)/2$. In other words, $f$ is a bijection among the eigenvalues of $BA$. As $f$ maps real numbers to real numbers, it must also be a bijection among the real eigenvalues of $BA$. Since $BA$ is a real matrix of odd dimension, real eigenvalues do exist. Now, as $f(x)\ge x$ on $\mathbb R$, $f$ must map the largest real eigenvalue of $BA$ to itself. Solving $f(x)=x$, we see that this eigenvalue is $1$. Edit. As the OP points out in his comment, actually we only need to consider the largest real eigenvalue. Let $(\lambda,x)$ be an eigenpair of $BA$, where $\lambda$ is the largest real eigenvalue of $BA$. Then $(\frac{\lambda^2+1}2,x)$ is an eigenpair of $AB$. However, since $AB$ and $BA$ have identical spectra and $\frac{\lambda^2+1}2\ge \lambda$, we must have $\frac{\lambda^2+1}2=\lambda$ and hence $\lambda=1$.
Complex series expansion of a particular form using formula
Yes this should be fine. It's helpful to note that $$\frac1{z-b}-\frac1{z-a}={a-b\over(z-a)(z-b)}.$$ The identity now follows by dividing both sides by $a-b$ and expanding using geometric series.
Does a set $A \subseteq [0,1]$ exist such that $A$ is homeomorphic to $[0,1] \setminus A$?
Such subsets do exist. The interval $[0,1]$ is homeomorphic to the extended real line $X=[-\infty, \infty]$ (with the standard topology). Now, let $A\subset X$ be the union of $\infty$ with the collection of intervals $[2n, 2n+1)$, $n\in {\mathbb Z}$. The set $A$ is homeomorphic to $B$, which is the union of $\infty$ and the collection of intervals $(2n, 2n+1]$: Send $\infty$ to itself and $[2n, 2n+1)\to (2n, 2n+1]$, $\forall n$ via linear maps. Composing this homeomorphism with the map $x\mapsto -x+1$, we get a homeomorphism $A\to X\setminus A$. qed Edit: Here is a proof that for such an example ($A$ homeomorphic to $A^{\mathrm c}=[0,1]\setminus A$) the set $A$ has to consist of infinitely many components. Suppose that $A$ is a finite union of intervals (I allow open, half-open, closed and degenerate intervals). For each interval $I$ define its "modified Euler characteristic" $\chi^c(I)$ as the number of vertices (end-points which belong to $I$) minus the number of edges (which is 1 if $I$ is nondegenerate and $0$ if $I$ is a singleton). Thus, for $I=[a,b]$, we get $\chi^c(I)=1$, while for $I=(0,1)$, we get $\chi^c(I)=-1$; we also have $\chi^c((0,1])=0$. Now, extend $\chi^c$ to finite unions of intervals in the obvious fashion. For compact subsets with finitely many components, $\chi^c=\chi$, the usual Euler characterstic. One can (easily) show that $\chi^c$ is additive: $$ \chi^c(\bigsqcup_{i=1}^n I_i)=\sum_{i=1}^n \chi^c(I_i) $$ (this is false for the usual Euler characteristic!) and is invariant under homeomorphisms. Now, if $A\subset [0,1]$ is a finite union of intervals and $A^{\mathrm c}$ is homeomorphic to $A$, then $$ 2\chi^c(A)=\chi^c(A)+ \chi^c(A^{\mathrm c})=\chi^c([0,1])=1 $$ which is absurd. The same works for the interval $(0,1)$. The same argument works in higher dimensions, but you have to modify what "finite number of components" means. Instead, assume that $A$ is "semialgebraic", i.e. is given by a finite system of inequalities of the type $p_i(x)&gt;0$, $p_j(x)\ge 0$, where $p$'s are polynomials of several variables. The key is that $\chi^c$ of the closed $n$-dimensional disk is $1$ and that $\chi^c$ is again additive. The modified Euler characteristic can be regarded as the "right" Euler characteristic for semialgebraic sets; it can be defined as the alternating sum of ranks of homology groups for the Delfs' homology theory ("homology with closed support", not to be confused with Borel-Moore!). This interpretation explains why $\chi^c$ is a topological invariant (this is no longer obvious with the 2nd definition below). A more direct definition of $\chi^c$ is to consider "incomplete simplicial complexes" triangulating semialgebraic sets, i.e. generalized simplicial complexes where where simplices might be missing some faces (like the interval $[0,1)$ is missing the vertex $1$) and then use the standard alternating sum of the face numbers, as I did above in the 1-dimensional case.
Two points with “identical” local geometry on Riemann Manifolds
I would always be careful with the word "identical" in mathematics. In this situation, what I would say is that the local geometries of $s$ and $t$ are isometric is there exist open subsets $U \subset S$ and $V \subset T$, and an isometry $f : U \to V$, such that $s \in U$, $t \in V$, and $f(s)=t$. In case you don't the definition, to say that $f$ is an isometry means that $f$ is a diffeomorphism, and for every $x \in U$ and $y=f(x) \in V$ the derivative map $D_x f : T_x U \to T_y V$ is a linear isomorphism that preserves the inner products on those spaces (given by the Riemannian metrics), i.e. for each $v,w \in T_x U$ we have $\langle v,w \rangle = \langle D_x(v),D_x(w) \rangle$.
What is $\lim_{n\to \infty} {\sqrt[n+1]{(n+1)!} \over \sqrt[n]{n!}}$
Note: $$\frac{\sqrt[n+1]{n!}}{\sqrt[n]{n!}} &lt;\frac{\sqrt[n+1]{(n+1)!}}{\sqrt[n]{n!}}&lt;\frac{\sqrt[n+1]{(n+1)!}}{\sqrt[n+1]{n!}}.$$ Taking limit: $$\lim_{n\to\infty} (n!)^{\frac{1}{n(n+1)}}\le L \le \lim_{n\to\infty} (n+1)^{\frac{1}{n+1}} \Rightarrow$$ $$1\le L \le 1 \Rightarrow $$ $$L=1.$$
Express unit sphere as countable union of great circles?
you may reduce the dimension of your question by considering the intersections of your great circles with a suitably chosen equator. so can a circle be made up of a countable set of points?
Proving a criteria for a sequence of functions on $[0,1]$ converging to a function non uniformly
This is false. Let $f(x) = 0$, and: $$ f_n(x) = \begin{cases} x^n, &amp;\text{if $x \in [0,1)$} \\ 0, &amp;\text{otherwise} \end{cases} \\ $$ We have $f_n(x) \to f(x)$ pointwise, as $\lim\limits_{n \to \infty} x^n = 0$ $\forall x \in [0,1)$, but not uniformly as $\sup\limits_{x \in [0,1]} |f_n(x) - f(x)| = 1$ $\forall n$. However, for any sequence $x_n \to x$ we must have $f(x_n) = 0 \to 0 = f(x)$.
Evaluating and proving $\lim\limits_{x\to\infty}\frac{\sin x}x$
You know that $-\dfrac{1}{x} \leq \dfrac{\sin(x)}{x} \leq \dfrac{1}{x}$ Now let $x \to \infty$ and apply the squeeze theorem.
Window perimeter optimization
Start by identifying the constraint and what you're trying to optimize/maximize: Constraint: $A = 2xy+\frac{\pi x^2}{2} = 10$ Maximize: $P = 2y+2x+\pi x$ From here, solve the constraint for $y$ so that we can make the maximization equation a function of $x$: $$10 = 2xy+\frac{\pi x^2}{2} \\ 2xy = 10-\frac{\pi x^2}{2} \\ y = \frac{5}{x}-\frac{\pi x}{4}$$ Plug this into the maximization equation: $$2\left(\frac{5}{x}-\frac{\pi x}{4}\right)+2x+\pi x \longrightarrow \frac{1}{2} (4+\pi ) x+\frac{10}{x}$$ Then take the derivative of this equation: $$\frac{d}{dx}P=\frac{4+\pi }{2}-\frac{10}{x^2}$$ Set it equal to $0$ to find critical points: $$x = \pm 2 \sqrt{\frac{5}{4+\pi}}$$ Plug it back into the equation for the perimeter: $$P\left(2 \sqrt{\frac{5}{4+\pi}}\right) = 2 \sqrt{5 (4+\pi )} \approx 11.9512$$
Prob. 2 (d), Sec. 27 in Munkres' TOPOLOGY, 2nd ed: Open supersets and ϵ-neighborhoods of compact set
Suppose by contradiction that no $U(A,\epsilon)$ is a subset of $U.$ For $n\in \mathbb N$ take $a_n\in A$ and $p_n\in B_d(a_n,1/n)$ \ $U.$ Since A is compact we can take a subsequence $(a_{n_i})_i$ of the sequence $(a_n)_n$ with $(a_{n_i})_i$ converging to some $a\in A.$ Now take $m\in \mathbb N$ such that $B_d(a,1/m)\subset U.$ Take any $n_i&gt;2m$ such that $d(a,a_{n_i})&lt;1/2m .$ Then $$d(a,p_{n_i})\leq d(a,a_{n_i})+d(a_{n_i},p_{n_i})&lt;1/2m+1/n_i&lt;1/m .$$ This implies $p_{n_i}\in B_d(a,1/m)\subset U,$ contrary to $p_{n_i}\not \in U.$
A Bound for the Error of the Numerical Approximation of a the Integral of a Continuous Function
This is where using randomness can help. Let $X_i\sim U(0,t)$ be independent and let $$I_n=\frac t n\sum_{i=1}^n f(X_i)$$ Then $I_n$ converges to $I=\int_0^tf(x)\,dx$ a.s. and in $L^2$: $$\sigma^2_n=E((I_n-I)^2)=\frac {\int_0^tf^2(x)\,dx-I^2}n\to 0$$ regardless of how hairy your function is. To get the simplest bound it's enough to know $M=\max_{x\in[0,t]} |f(x)|$ for then $\sigma_n^2\le\frac {M^2t}n$.
Sum of certain binomial coefficients
Hint: Each term is just $\dbinom{q+k}{k} = \dbinom{q+k}q$. Then just hockey stick identity.
Ramification of $5$ in $\mathbb{Q}( \sqrt[5]{n})$
I will show that there is only one prime lying over $5$ in $\mathcal{O}_K$. I will use the following proposition, which may be found in Neukirch's Algebraic Number Theory (prop. 8.2): Proposition: Suppose the extension $L/K$ is generated by the zero $\alpha$ of the irreducible polynomial $f\in K\left[ X\right]$. Then the valuations $w_1,\dots,w_r$ of $L$ extending a valuation $v$ of $K$ correspond $1-1$ to the irreducible factors $f_1,\dots,f_r$ in the decomposition of $f$ over the completion $K_v$. Using this proposition, it is enough to show that $X^5-n$ has a unique irreducible factor in $\mathbf{Q}_5$. If $n=0$, this is evident. Otherwise, one may show that $X^5-n$ is irreducible. First note that $X^n-5$ has no quadratic factor in $\mathbf{Q}_5$: indeed, let $\alpha\in \overline{\mathbf{Q}_5}$ be a root of such a quadratic factor $Q$ so that $Q$ has roots $\alpha$ and $\alpha\zeta$ for some nontrivial fifth root of unity $\zeta$; then $\alpha^2\zeta$ and $\alpha(1+\zeta)$ are in $\mathbf{Q}_5$ (these are the coefficients of $Q$); from this, one deduces that $\alpha^2$ is in $\mathbf{Q}_5$, hence that $\zeta$ is in $\mathbf{Q}_5$, which leads to a contradiction. The same kind of reasoning shows that $X^5-n$ does not split, that it has no cubic factor with two linear factors nor any factor of degree $4$ (in all those cases, the contradiction is always the same: $\mathbf{Q}_5$ would contain a fifth primitve root of unity). Thus, $X^5-n$ has a unique irreducible factor in $\mathbf{Q}_5$ which means that $5$ is totally ramified in $\mathcal{O}_K$.
Limit with trigonometric function
$$\text{Let }\displaystyle y=\lim_{x\to1}\left(\frac{x^2+1}{2x}\right)^{\cot^2{\pi x}} \implies\ln y=\lim_{x\to1} \frac{\cos^2{\pi x}}{\sin^2\pi x}\cdot \ln\left(\frac{x^2+1}{2x}\right)$$ Setting $x-1=h,$ $$\ln y=\lim_{h\to0} \frac{\cos^2{\pi(h+1)}}{\sin^2\pi(h+1)}\cdot \ln\left(\frac{(h+1)^2+1}{2(h+1)}\right)$$ Now as $\displaystyle\frac{(h+1)^2+1}{2(h+1)}=1+\frac{h^2}{2(h+1)},$ $\displaystyle \cos^2{\pi(h+1)}=\{\cos(\pi+\pi\cdot h)\}^2=\{-\cos(\pi\cdot h)\}^2=\cos^2(\pi\cdot h)$ Similarly, $\displaystyle \sin^2{\pi(h+1)}=\sin^2(\pi\cdot h)$ using $\displaystyle\lim_{u\to0}\frac{\ln(1+u)}u=1$ and $\displaystyle\lim_{v\to0}\frac{\sin v}v=1,$ $\displaystyle\ln y$ $$=\left(\lim_{h\to0}\cos(\pi\cdot h)\right)^2\lim_{h\to0}\frac{\ln\left(1+\frac{h^2}{2(h+1)}\right)}{\frac{h^2}{2(h+1)}}\cdot\frac1{\left(\lim_{h\to0}\frac{\sin(\pi\cdot h)}{\pi \cdot h}\right)^2}\cdot\frac1{\lim_{h\to0}2(h+1)\pi^2}$$ $$\implies \ln y=1^2\cdot1\cdot\frac1{1^2}\cdot\frac1{2\pi^2}$$ $$\implies \ln y=e^{\frac1{2\pi^2}} $$
Covariance, covariance operator, and covariance function
The first definition is a special case of the second. Rather than a Hilbert space H, let's look at a Banach space $X$, and distinguish it from its dual space $X^*$. Every continuous linear functional $\varphi \in X^*$ is a random variable $\varphi : X \to \mathbb R$, so it makes sense to take expectations and covariances. We define the expectation of a functional by $\mathbb E[\varphi] = \int_X \varphi[x] \, \mathrm d \mathbf P(x)$, and the covariance of two functionals to be $$\operatorname{cov}[\psi|\varphi] = \int_X \big( \psi[x] - \mathbb E[\psi]\big) \big( \varphi[x] - \mathbb E[\varphi]\big) \, \mathrm d \mathbf P(x).$$ Now, consider a probability measure $\mathbf P$ on a Hilbert space $X = H$. By the Riesz representation theorem, we know that the dual space $H^*$ is isomorphic to $H$, and all the functionals are of the form $\varphi_h[x] := \langle h, x \rangle$. The mean can be represented by a single element $m \in H$, which is called the "Pettis integral" of $\mathbf P$. This element satisfies the property that $\mathbb E[\varphi_h] = \varphi_h[m] = \langle h, m \rangle$ for all $h \in H$. Consequently, $$\operatorname{cov}[\varphi_h|\varphi_{h'}] = \int_X \big\langle h, x - m \big\rangle \big\langle h', x-m \big\rangle \, \mathrm d \mathbf P(x).$$ This is the formula you were looking for. It's just a special case of the usual covariance formula, specialized to the setting where the random variables of interest are continuous linear functionals, and the probability space is a Hilbert space.
Lebesgue integrable implies zero almost everywhere
Hint: If $b\in L^\infty(\mathbb R),$ then there exists a sequence $g_n$ of bounded continuous functions on $\mathbb R$ such that $\|g_n\|_\infty\le \|b\|_\infty$ for all $n,$ and $g_n \to b$ pointwise a.e.
What is the explicit formula for the general term of the sequence?
Hint: $$ A_n = \begin{bmatrix} a_n \\ a_{n-1} \end{bmatrix} = \begin{bmatrix} 5a_{n-1} - 6a_{n-2} \\ a_{n-1} \end{bmatrix} = \begin{bmatrix} 5 &amp; -6 \\ 1 &amp; 0 \end{bmatrix} \begin{bmatrix} a_{n-1}\\ a_{n-2} \end{bmatrix} = \begin{bmatrix} 5 &amp; -6 \\ 1 &amp; 0 \end{bmatrix} A_{n-1}. $$
How to find tangential accel?
You have $\theta=0.3t^2\implies \dot{\theta}=0.6t\implies \ddot{\theta}=0.6$ So for C, the tangential acceleration is $$r\ddot{\theta}=15\times0.6=9$$ And for D, the radial acceleration is $$r\dot{\theta}^2=15\times3.6^2=194.4$$