domain
sequencelengths
0
3
difficulty
float64
1
9.5
problem
stringlengths
18
9.01k
solution
stringlengths
2
11.1k
answer
stringlengths
0
3.77k
source
stringclasses
67 values
[ "Mathematics -> Algebra -> Algebra -> Equations and Inequalities" ]
7
Find all solutions to $(m^2+n)(m + n^2)= (m - n)^3$ , where m and n are non-zero integers. Do it
Expanding both sides, \[m^3+mn+m^2n^2+n^3=m^3-3m^2n+3mn^2-n^3\] Note that $m^3$ can be canceled and as $n \neq 0$ , $n$ can be factored out. Writing this as a quadratic equation in $n$ : \[2n^2+(m^2-3m)n+(3m^2+m)=0\] . The discriminant $b^2-4ac$ equals \[(m^2-3m)^2-8(3m^2+m)\] \[=m^4-6m^3-15m^2-8m\] , which we want to be a perfect square. Miraculously, this factors as $m(m-8)(m+1)^2$ . This is square iff (if and only if) $m^2-8m$ is square or $m+1=0$ . It can be checked that the only nonzero $m$ that work are $-1, 8, 9$ . Finally, plugging this in and discarding extraneous roots gives all possible ordered pairs $(m, n)$ as \[\{(-1,-1),(8,-10),(9,-6),(9,-21)\}\] .
\[ \{(-1,-1), (8,-10), (9,-6), (9,-21)\} \]
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Triangulations", "Mathematics -> Algebra -> Intermediate Algebra -> Complex Numbers", "Mathematics -> Geometry -> Plane Geometry -> Angles" ]
8
Given circle $O$ with radius $R$, the inscribed triangle $ABC$ is an acute scalene triangle, where $AB$ is the largest side. $AH_A, BH_B,CH_C$ are heights on $BC,CA,AB$. Let $D$ be the symmetric point of $H_A$ with respect to $H_BH_C$, $E$ be the symmetric point of $H_B$ with respect to $H_AH_C$. $P$ is the intersection of $AD,BE$, $H$ is the orthocentre of $\triangle ABC$. Prove: $OP\cdot OH$ is fixed, and find this value in terms of $R$. (Edited)
Given a circle \( O \) with radius \( R \), and an inscribed acute scalene triangle \( ABC \) where \( AB \) is the largest side, let \( AH_A, BH_B, CH_C \) be the altitudes from \( A, B, C \) to \( BC, CA, AB \) respectively. Let \( D \) be the symmetric point of \( H_A \) with respect to \( H_BH_C \), and \( E \) be the symmetric point of \( H_B \) with respect to \( H_AH_C \). Let \( P \) be the intersection of \( AD \) and \( BE \), and \( H \) be the orthocenter of \( \triangle ABC \). We aim to prove that \( OP \cdot OH \) is fixed and find this value in terms of \( R \). To solve this, we use complex numbers and the properties of the orthocenter and the circumcircle. Let the circumcircle of \( \triangle ABC \) be the unit circle in the complex plane. The orthocenter \( H \) of \( \triangle ABC \) can be represented as \( h = a + b + c \), where \( a, b, c \) are the complex numbers corresponding to the vertices \( A, B, C \) respectively. The feet of the altitudes \( H_A, H_B, H_C \) can be expressed as: \[ h_a = \frac{1}{2} \left( a + b + c - \frac{bc}{a} \right), \] and similarly for \( h_b \) and \( h_c \). The point \( P \), which is the pole of \( H \) with respect to the circumcircle, is given by: \[ p = \frac{1}{\overline{h}} = \frac{abc}{ab + bc + ac}. \] Next, we compute the symmetric points \( D \) and \( E \). Let \( X \) be the foot of the perpendicular from \( H_A \) to \( H_BH_C \). Solving for \( X \) using the properties of perpendiculars in the complex plane, we find: \[ 2x = h_a + h_b + a^2 (\overline{h_b - h_a}), \] which simplifies to: \[ d = 2x - h_a = \frac{1}{2} \left( a + b + c - \frac{ac}{b} - \frac{ab}{c} + \frac{a^3}{bc} \right). \] We then show that \( D, A, \) and \( P \) are collinear by computing: \[ \frac{d - a}{\overline{d - a}} = \frac{a - p}{\overline{a - p}} = \frac{a^3 (a + b + c)}{ab + ac + bc}. \] Finally, since \( P \) is the pole of \( H \) with respect to the circumcircle, the product \( OP \cdot OH \) is given by: \[ OP \cdot OH = R^2. \] Thus, the value of \( OP \cdot OH \) is fixed and equals \( R^2 \). The answer is: \boxed{R^2}.
R^2
china_team_selection_test
[ "Mathematics -> Algebra -> Algebra -> Equations and Inequalities", "Mathematics -> Discrete Mathematics -> Combinatorics" ]
8
The $2010$ positive numbers $a_1, a_2, \ldots , a_{2010}$ satisfy the inequality $a_ia_j \le i+j$ for all distinct indices $i, j$ . Determine, with proof, the largest possible value of the product $a_1a_2\cdots a_{2010}$ .
The largest possible value is \[\prod_{i=1}^{1005}(4i-1) = 3\times 7 \times \ldots \times 4019.\] Proof No larger value is possible, since for each consecutive pair of elements: $(a_{2i-1},a_{2i}), 1\le i \le 1005$ , the product is at most $(2i-1) + 2i = 4i - 1$ , and so the product of all the pairs is at most: If we can demonstrate a sequence in which for all $1 \le i \le 1005$ the product $a_{2i-1}a_{2i} = 4i-1$ , and all the inequalities are satisfied, the above upper bound will be achieved and the proof complete. We will construct sequences of an arbitrarily large even length $2n \ge 4$ , in which: Given $a_1$ , from the equations $a_ia_{i+1} = 2i+1,\; 1\le i\le 2n-1$ , we obtain the whole sequence recursively: $a_1 = a_1,\; a_2 = 3/a_1,\; a_3 = 5/a_2 = 5a_1/3,\; a_4 = 7/a_3 = (3\cdot 7)/(5a_1) \ldots.$ And as a result: The same equations $a_ia_{i+1} = 2i+1$ can be used to compute the whole sequence from any other known term. We will often need to compare fractions in which the numerator and denominator are both positive, with fractions in which a positive term is added to both. Suppose $p, q, r$ are three positive real numbers, then: Returning to the problem in hand, for $i < j$ , $a_ia_j \le i+j \implies a_ia_{j+2} < i+j+2$ . If it were otherwise, we would have for some $i < j$ : so our assumption is impossible. Therefore, we need only verify inequalities with an index difference of $1$ or $2$ , as these imply the rest. Now, when the indices differ by $1$ we have ensured equality (and hence the desired inequalities) by construction. So, we only need to prove the inequalities for successive even index and successive odd index pairs, i.e. for every index $i > 2$ , prove $a_{i-2}a_i \le 2i-2$ . We now compare $a_ia_{i+2}/(2i+2)$ with $a_{i+2}a_{i+4}/(2i+6)$ . By our recurrence relations: So, for both odd and even index pairs, the strict inequality $a_ia_{i+2} < 2i+2$ follows from $a_{i+2}a_{i+4} \le 2i+6$ and we need only prove the inequalities $a_{2n-3}a_{2n-1} \le 4n-4$ and $a_{2n-2}a_{2n} \le 4n-2$ , the second of which holds (as an equality) by construction, so only the first remains. We have not yet used the equation $a_{2n-2}a_{2n} = 4n-2$ , with this we can solve for the last three terms (or equivalently their squares) and thus compute the whole sequence. From the equations: multiplying any two and dividing by the third, we get: from which, With the squares of the last four terms in hand, we can now verify the only non-redundant inequality: The inequality above follows because the numerator and denominator are both positive for $n > 1$ . This completes the construction and the proof of all the inequalities, which miraculously reduced to just one inequality for the last pair of odd indices. Additional observations If we choose a different first term, say $a_1' = M\cdot a_1$ , the sequence $a_i'$ will have the form: the same holds if we have a longer sequence, at every index of the shorter sequence, the longer sequence will be a constant multiple (for all the odd terms) or dividend (for all the even terms) of the corresponding term of shorter sequence. We observe that our solution is not unique, indeed for any $k>0$ , the same construction with $2n+2k$ terms, truncated to just the first $2n$ terms, yields a sequence $a'_i$ which also satisfies all the required conditions, but in this case $a'_{2n-2}a'{2n} < 4n-2$ . We could have constructed this alternative solution directly, by replacing the right hand side in the equation $a_{2n-2}a_{2n} = 4n-2$ with any smaller value for which we still get $a_{2n-3}a_{2n-1} \le 4n-4$ . In the modified construction, for some constant $M > 1$ , we have: and so: which satisfies the required inequality provided: The ratio $M_{\mathrm{max}}$ , between the largest and smallest possible value of $a_{2n-3}$ is in fact the ratio between the largest and smallest values of $a_1$ that yield a sequence that meets the conditions for at least $2n$ terms. In the $n=2$ case, the equation for $a_{2n-3}$ gives: $a_1^2 = \frac{21}{10}$ . We will next consider what happens to $a_1^2$ , and the sequence of squares in general, as $n$ increases. Let $A_{n,2i-1}, A_{n,2i}$ denote the $i^{\mathrm{th}}$ odd and even terms, respectively, of the unique sequence which satisfies our original equations and has $2n$ terms in total. Let $A_{n+1,2i-1}, A_{n+1,2i}$ be the odd and even terms of the solution with $2n+2$ terms. We already noted that there must exist a constant $M_n$ (that depends on $n$ , but not on $i$ ), such that: This constant is found explicitly by comparing the squares of the last term $A_{n,2n}$ of the solution of length $2n$ with the square of the third last term $A_{n+1,2n}$ of the solution of length $2n+2$ : Clearly $M_n > 1$ for all positive $n$ , and so for fixed $i$ , the odd index terms $A_{n,2i-1}$ strictly increase with $n$ , while the even index terms $A_{n,2i}$ decrease with $n$ . Therefore, for $n \ge 2$ , The product converges to a finite value even if taken infinitely far, and we can conclude (by a simple continuity argument) that there is a unique infinite positive sequence $A_\omega$ , in which $A_{\omega,i}A_{\omega,i+1} = 2i+1$ , that satisfies all the inequalities $A_{\omega,i}A_{\omega,j} < i+j,\; i \le j - 2$ . The square of the first term of the infinite sequence is: In summary, if we set $a_1 = \frac{\sqrt{\pi}}{\mathrm{AGM}(\sqrt{2}, 1)}$ , and then recursively set $a_{i+1} = (2i + 1)/a_i$ , we get an infinite sequence that, for all $n \ge 1$ , yields the maximum possible product $a_1a_2\cdots a_{2n}$ , subject to the conditions $a_ia_j \le i+j,\; 1 \le i < j \le 2n$ .
\[ \prod_{i=1}^{1005}(4i-1) = 3 \times 7 \times \ldots \times 4019 \]
usamo
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
5.5
Let $P_1P_2\ldots P_{24}$ be a regular $24$-sided polygon inscribed in a circle $\omega$ with circumference $24$. Determine the number of ways to choose sets of eight distinct vertices from these $24$ such that none of the arcs has length $3$ or $8$.
Let \( P_1P_2\ldots P_{24} \) be a regular 24-sided polygon inscribed in a circle \(\omega\) with circumference 24. We aim to determine the number of ways to choose sets of eight distinct vertices from these 24 such that none of the arcs has length 3 or 8. We generalize the problem by considering a regular polygon with \(3n\) vertices and selecting \(n\) vertices such that no two selected vertices are 3 or \(n\) apart. Label the vertices \(1, 2, \ldots, 3n\) and group them into sets of three: \(\{1, n+1, 2n+1\}\), \(\{2, n+2, 2n+2\}\), and so on until \(\{n, 2n, 3n\}\). Since we need to select \(n\) vertices, one from each group, the condition that no two vertices are \(n\) apart is automatically satisfied. Next, we need to ensure that no two selected vertices are 3 apart. Let \(a_n\) denote the number of ways to select \(n\) vertices with the given properties. Clearly, \(a_1 = 0\) because each vertex is three apart from itself. For \(a_2\), we manually compute that there are 6 valid sets. To find a general formula, we construct a recursion relation. Initially, there are \(3 \cdot 2^{n-1}\) ways to select \(n\) vertices, ignoring the condition that no two vertices can be 3 apart. However, this count overestimates the number of valid sets. The overcount is equal to the number of valid sets of \(n-1\) vertices, leading to the recursion relation: \[ a_n = 3 \cdot 2^{n-1} - a_{n-1}. \] To solve this, we derive a closed form. From the recursion relation, we get: \[ a_{n+1} = 3 \cdot 2^n - a_n. \] Subtracting the first equation from the second and simplifying, we obtain: \[ a_{n+1} = 3 \cdot 2^{n-1} + a_{n-1}. \] Further manipulation yields: \[ a_n - a_{n+1} = -2a_{n-1}. \] Rearranging and shifting indices, we find: \[ a_n = a_{n-1} + 2a_{n-2}. \] The characteristic polynomial of this recurrence relation has roots 2 and -1, giving us the general solution: \[ a_n = A \cdot 2^n + B(-1)^n. \] Using the initial conditions \(a_1 = 0\) and \(a_2 = 6\), we determine the constants \(A\) and \(B\): \[ A = 1, \quad B = 2. \] Thus, the closed form is: \[ a_n = 2^n + 2(-1)^n. \] For \(n = 8\), we have: \[ a_8 = 2^8 + 2(-1)^8 = 256 + 2 = 258. \] The answer is: \boxed{258}.
258
china_national_olympiad
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations" ]
7
$P(x)$ is a polynomial of degree $3n$ such that \begin{eqnarray*} P(0) = P(3) = \cdots &=& P(3n) = 2, \\ P(1) = P(4) = \cdots &=& P(3n-2) = 1, \\ P(2) = P(5) = \cdots &=& P(3n-1) = 0, \quad\text{ and }\\ && P(3n+1) = 730.\end{eqnarray*} Determine $n$ .
By Lagrange Interpolation Formula $f(x) = 2\sum_{p=0}^{n}\left ( \prod_{0\leq r\neq3p\leq 3n}^{{}}\frac{x-r}{3p-r} \right )+ \sum_{p=1}^{n}\left ( \prod_{0\leq r\neq3p-2\leq 3n}^{{}} \frac{x-r}{3p-2-r}\right )$ and hence $f(3n+1) = 2\sum_{p=0}^{n}\left ( \prod_{0\leq r\neq3p\leq 3n}^{{}}\frac{3n+1-r}{3p-r} \right )+ \sum_{p=1}^{n}\left ( \prod_{0\leq r\neq3p-2\leq 3n}^{{}} \frac{3n+1-r}{3p-2-r}\right )$ after some calculations we get $f(3n+1) =\left ( \binom{3n+1}{0}- \binom{3n+1}{3}+\binom{3n+1}{6}- ... \right )\left ( 2.(-1)^{3n}-1 \right )+1$ Given $f(3n+1)= 730$ so we have to find $n$ such that $\left ( \binom{3n+1}{0}- \binom{3n+1}{3}+\binom{3n+1}{6}- ... \right )\left ( 2.(-1)^{3n}-1 \right )= 729$ Lemma: If $p$ is even $\binom{p}{0}- \binom{p}{3}+ \binom{p}{6}- \cdots = \frac{2^{p+1}sin^{p}\left ( \frac{\pi}{3} \right )(i)^{p}\left ( cos\left ( \frac{p\pi}{3} \right ) \right )}{3}$ and if $p$ is odd $\binom{p}{0}- \binom{p}{3}+ \binom{p}{6}- \cdots = \frac{-2^{p+1}sin^{p}\left ( \frac{\pi}{3} \right )(i)^{p+1}\left ( sin\left ( \frac{p\pi}{3} \right ) \right )}{3}$ $i$ is $\sqrt{-1}$ Using above lemmas we do not get any solution when $n$ is odd, but when $n$ is even $3n+1=13$ satisfies the required condition, hence $n=4$
\[ n = 4 \]
usamo
[ "Mathematics -> Number Theory -> Factorization" ]
7
Find all integers $n \geq 3$ such that the following property holds: if we list the divisors of $n !$ in increasing order as $1=d_1<d_2<\cdots<d_k=n!$ , then we have \[d_2-d_1 \leq d_3-d_2 \leq \cdots \leq d_k-d_{k-1} .\] Contents 1 Solution (Explanation of Video) 2 Solution 2 3 Video Solution 4 See Also
We claim only $n = 3$ and $n = 4$ are the only two solutions. First, it is clear that both solutions work. Next, we claim that $n < 5$ . For $n \geq 5$ , let $x$ be the smallest $x$ such that $x+1$ is not a factor of $n!$ . Let the smallest factor larger than $x$ be $x+k$ . Now we consider $\frac{n!}{x-1}$ , $\frac{n!}{x}$ and $\frac{n!}{x+k}$ . Since $\frac{n!}{x-1} > \frac{n!}{x} > \frac{n!}{x+k}$ , if $n$ were to satisfy the conditions, then $\frac{n!}{x-1}-\frac{n!}{x} \geq \frac{n!}{x} - \frac{n!}{x+k}$ . However, note that this is not true for $x \geq 5$ and $k > 1$ . Note that the inequality is equivalent to showing $\frac{1}{x(x-1)} \geq \frac{k}{x(x+k)}$ , which simplifies to $x+k \geq kx-k$ , or $\frac{x}{x-2} \geq k \geq 2$ . This implies $x \geq 2x-4$ , $x \leq 4$ , a contradiction, since the set of numbers $\{1, 2, 3, 4, 5\}$ are all factors of $n!$ , and the value of $x$ must exist. Hence, no solutions for $n \geq 5$ .
The integers \( n \geq 3 \) that satisfy the given property are \( n = 3 \) and \( n = 4 \).
usamo
[ "Mathematics -> Applied Mathematics -> Statistics -> Probability -> Counting Methods -> Other" ]
6
Let $n$ be a nonnegative integer. Determine the number of ways that one can choose $(n+1)^2$ sets $S_{i,j}\subseteq\{1,2,\ldots,2n\}$ , for integers $i,j$ with $0\leq i,j\leq n$ , such that: 1. for all $0\leq i,j\leq n$ , the set $S_{i,j}$ has $i+j$ elements; and 2. $S_{i,j}\subseteq S_{k,l}$ whenever $0\leq i\leq k\leq n$ and $0\leq j\leq l\leq n$ . Contents 1 Solution 1 2 Solution 2 2.1 Lemma 2.2 Filling in the rest of the grid 2.3 Finishing off 3 See also
Note that there are $(2n)!$ ways to choose $S_{1, 0}, S_{2, 0}... S_{n, 0}, S_{n, 1}, S_{n, 2}... S{n, n}$ , because there are $2n$ ways to choose which number $S_{1, 0}$ is, $2n-1$ ways to choose which number to append to make $S_{2, 0}$ , $2n-2$ ways to choose which number to append to make $S_{3, 0}$ ... After that, note that $S_{n-1, 1}$ contains the $n-1$ in $S_{n-1. 0}$ and 1 other element chosen from the 2 elements in $S_{n, 1}$ not in $S_{n-1, 0}$ so there are 2 ways for $S_{n-1, 1}$ . By the same logic there are 2 ways for $S_{n-1, 2}$ as well so $2^n$ total ways for all $S_{n-1, j}$ , so doing the same thing $n-1$ more times yields a final answer of $(2n)!\cdot 2^{n^2}$ . -Stormersyle
\[ (2n)! \cdot 2^{n^2} \]
usajmo
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
6
What is the largest number of towns that can meet the following criteria. Each pair is directly linked by just one of air, bus or train. At least one pair is linked by air, at least one pair by bus and at least one pair by train. No town has an air link, a bus link and a train link. No three towns, $A, B, C$ are such that the links between $AB, AC$ and $BC$ are all air, all bus or all train.
Assume $AB$ , $AC$ , and $AD$ are all rail. None of $BC$ , $CD$ , or $CD$ can be rail, as those cities would form a rail triangle with $A$ . If $BC$ is bus, then $BD$ is bus as well, as otherwise $B$ has all three types. However, $CD$ cannot be rail (as $\triangle ACD$ would be a rail triangle), bus (as $BCD$ would be a bus triangle), or ferry (as $C$ and $D$ would have all three types). Therefore, no city can have three connections of the same type. Assume there are 5 towns - $A$ , $B$ , $C$ , $D$ , and $E$ . Two connections from $A$ must be of one type, and two of another; otherwise there would be at least three connections of the same type from $A$ , which has been shown to be impossible. Let $AB$ and $AC$ be rail connections, and $AD$ and $AE$ be bus. Assume $CD$ is air. $BC$ cannot be rail ( $\triangle ABC$ would be a rail triangle) or bus ( $C$ would have all three types), so $BC$ must be air. $DE$ cannot be bus ( $\triangle ADE$ would be a bus triangle) or rail ( $D$ would have all three types), so $DE$ must be air. $BE$ cannot be rail ( $E$ would have all three types) or bus ( $B$ would have all three types), so $BE$ must be air. However, $BD$ cannot be rail ( $D$ would have all three types), bus ( $B$ would have all three types), or air ( $D$ would have three air connections). Therefore, the assumption that $CD$ is air is false. $CD$ can equally be rail or bus; assume it is bus. $BC$ cannot be rail ( $\triangle ABC$ would be a rail triangle) or air ( $C$ would have all three types), so $BC$ must be bus. $BD$ cannot be air ( $B$ would have all three types) or bus ( $D$ would have three bus connections), so $BD$ must be rail. $DE$ cannot be air ( $D$ would have all three types) or bus ( $D$ would have three bus connections), so $DE$ must be rail. $CE$ cannot be air ( $C$ would have all three types) or bus ( $E$ would have three bus connections), so $CE$ must be rail. The only connection remaining is $BE$ , which cannot be orange as both $B$ and $D$ would have all three types, but this means there are no air connections. Therefore, it is impossible with five (or more) towns. A four-town mapping is possible: $AB$ , $BC$ , $CD$ , and $DA$ are connected by bus. $AC$ is connected by rail. $BD$ is connected by air. Therefore, the maximum number of towns is $4$ .
The maximum number of towns is \(4\).
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Polygons" ]
5
Let $n>3$ be a positive integer. Equilateral triangle ABC is divided into $n^2$ smaller congruent equilateral triangles (with sides parallel to its sides). Let $m$ be the number of rhombuses that contain two small equilateral triangles and $d$ the number of rhombuses that contain eight small equilateral triangles. Find the difference $m-d$ in terms of $n$ .
First we will show that the side lengths of the small triangles are $\tfrac{1}{n}$ of the original length. Then we can count the two rhombuses. Lemma: Small Triangle is Length of Original Triangle Let the side length of the triangle be $x$ , so the total area is $\tfrac{x^2 \sqrt{3}}{4}$ . Since the big triangle is divided into $n^2$ congruent equilateral triangles, the area of each smaller equilateral triangle is \[\frac{x^2 \sqrt{3}}{4n^2}\] \[\frac{\left( \frac{x}{n} \right)^2 \sqrt{3}}{4}.\] Thus, the length of each smaller triangle is $\tfrac{x}{n}. \blacktriangleright$ Since the lengths of each smaller triangle is $\tfrac{x}{n}$ and are parallel to the big triangle’s sides, the triangles are arranged like the below diagram. In diagram, $n=4$ . Now we can count the rhombuses. After dividing the triangles, for each line that is not in the border of the big equilateral triangle, there are two smaller equilateral triangles. The longest row in the triangle has $n-1$ lines and the shortest row has $1$ line. Thus, there are $\tfrac{n(n-1)}{2}$ lines parallel to one side, resulting in a total of $\tfrac{3n(n-1)}{2}$ lines or rhombuses with 2 equilateral triangles. For a rhombus to have eight triangles, the center line must have two lines and the rows above and below must have one line. We can count the number of segments made of two lines that are not in the last row in the big triangle. The second longest row in the triangle has $n-2$ lines, while the second shortest row has $2$ lines. Thus, there are $\tfrac{(n-2)(n-3)}{2}$ valid segments from two lines parallel to one side, resulting in a total of $\tfrac{3(n-2)(n-3)}{2}$ rhombuses with 8 equilateral triangles. Because $m = \tfrac{3n(n-1)}{2}$ and $d = \tfrac{3(n-2)(n-3)}{2}$ , \begin{align*} m-n &= \frac{3n(n-1)}{2} - \frac{3(n-2)(n-3)}{2} \\ &= \frac{3n^2 - 3n}{2} - \frac{3n^2 - 15n + 18}{2} \\ &= \frac{12n - 18}{2} \\ &= \boxed{6n - 9}. \end{align*}
\[ 6n - 9 \]
jbmo
[ "Mathematics -> Geometry -> Solid Geometry -> 3D Shapes" ]
7
$P$ , $A$ , $B$ , $C$ , and $D$ are five distinct points in space such that $\angle APB = \angle BPC = \angle CPD = \angle DPA = \theta$ , where $\theta$ is a given acute angle. Determine the greatest and least values of $\angle APC + \angle BPD$ .
Greatest value is achieved when all the points are as close as possible to all being on a plane. Since $\theta < \frac{\pi}{2}$ , then $\angle APC + \angle BPD < \pi$ Smallest value is achieved when point P is above and the remaining points are as close as possible to colinear when $\theta > 0$ , then $\angle APC + \angle BPD > 0$ and the inequality for this problem is: $0 < \angle APC + \angle BPD < \pi$ ~Tomas Diaz. [email protected] Alternate solutions are always welcome. If you have a different, elegant solution to this problem, please add it to this page.
\[ 0 < \angle APC + \angle BPD < \pi \]
usamo
[ "Mathematics -> Algebra -> Algebra -> Algebraic Expressions", "Mathematics -> Number Theory -> Prime Numbers" ]
8
Let $\mathbb{N}$ be the set of positive integers. A function $f:\mathbb{N}\to\mathbb{N}$ satisfies the equation \[\underbrace{f(f(\ldots f}_{f(n)\text{ times}}(n)\ldots))=\frac{n^2}{f(f(n))}\] for all positive integers $n$ . Given this information, determine all possible values of $f(1000)$ .
Let $f^r(x)$ denote the result when $f$ is applied to $f^{r-1}(x)$ , where $f^1(x)=f(x)$ . $\hfill \break \hfill \break$ If $f(p)=f(q)$ , then $f^2(p)=f^2(q)$ and $f^{f(p)}(p)=f^{f(q)}(q)$ $\implies p^2=f^2(p)\cdot f^{f(p)}(p)=f^2(q)\cdot f^{f(q)}(q)=q^2$ $\implies p=\pm q$ $\implies p=q$ since $p,q>0$ . Therefore, $f$ is injective. It follows that $f^r$ is also injective. Lemma 1: If $f^r(b)=a$ and $f(a)=a$ , then $b=a$ . Proof: $f^r(b)=a=f^r(a)$ which implies $b=a$ by injectivity of $f^r$ . Lemma 2: If $f^2(m)=f^{f(m)}(m)=m$ , and $m$ is odd, then $f(m)=m$ . Proof: Let $f(m)=k$ . Since $f^2(m)=m$ , $f(k)=m$ . So, $f^2(k)=k$ . $\newline f^2(k)\cdot f^{f(k)}(k)=k^2$ . Since $k\neq0$ , $f^{f(k)}(k)=k$ $\implies f^m(k)=k$ $\implies f^{gcd(m, 2)}(k)=k$ $\implies f(k)=k$ This proves Lemma 2. I claim that $f(m)=m$ for all odd $m$ . Otherwise, let $m$ be the least counterexample. Since $f^2(m)\cdot f^{f(m)}(m)=m^2$ , either $(1) f^2(m)=k<m$ , contradicted by Lemma 1 since $k$ is odd and $f^2(k)=k$ . $(2) f^{f(m)}(m)=k<m$ , also contradicted by Lemma 1 by similar logic. $(3) f^2(m)=m$ and $f^{f(m)}(m)=m$ , which implies that $f(m)=m$ by Lemma 2. This proves the claim. By injectivity, $f(1000)$ is not odd. I will prove that $f(1000)$ can be any even number, $x$ . Let $f(1000)=x, f(x)=1000$ , and $f(k)=k$ for all other $k$ . If $n$ is equal to neither $1000$ nor $x$ , then $f^2(n)\cdot f^{f(n)}(n)=n\cdot n=n^2$ . This satisfies the given property. If $n$ is equal to $1000$ or $x$ , then $f^2(n)\cdot f^{f(n)}(n)=n\cdot n=n^2$ since $f(n)$ is even and $f^2(n)=n$ . This satisfies the given property. The problems on this page are copyrighted by the Mathematical Association of America 's American Mathematics Competitions .
The possible values of \( f(1000) \) are all even numbers.
usamo
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
5
Find, with proof, the maximum positive integer \(k\) for which it is possible to color \(6k\) cells of a \(6 \times 6\) grid such that, for any choice of three distinct rows \(R_{1}, R_{2}, R_{3}\) and three distinct columns \(C_{1}, C_{2}, C_{3}\), there exists an uncolored cell \(c\) and integers \(1 \leq i, j \leq 3\) so that \(c\) lies in \(R_{i}\) and \(C_{j}\).
The answer is \(k=4\). This can be obtained with the following construction: [grid image]. It now suffices to show that \(k=5\) and \(k=6\) are not attainable. The case \(k=6\) is clear. Assume for sake of contradiction that the \(k=5\) is attainable. Let \(r_{1}, r_{2}, r_{3}\) be the rows of three distinct uncolored cells, and let \(c_{1}, c_{2}, c_{3}\) be the columns of the other three uncolored cells. Then we can choose \(R_{1}, R_{2}, R_{3}\) from \(\{1,2,3,4,5,6\} \backslash\left\{r_{1}, r_{2}, r_{3}\right\}\) and \(C_{1}, C_{2}, C_{3}\) from \(\{1,2,3,4,5,6\} \backslash\left\{c_{1}, c_{2}, c_{3}\right\}\) to obtain a contradiction.
\[ k = 4 \]
HMMT_2
[ "Mathematics -> Algebra -> Abstract Algebra -> Other", "Mathematics -> Algebra -> Equations and Inequalities -> Other" ]
6
Find all functions $f:\mathbb{R}\rightarrow\mathbb{R}$ that satisfy \[f(x^2-y)+2yf(x)=f(f(x))+f(y)\] for all $x,y\in\mathbb{R}$ .
Plugging in $y$ as $0:$ \begin{equation} f(x^2)=f(f(x))+f(0) \text{ } (1) \end{equation} Plugging in $x, y$ as $0:$ \[f(0)=f(f(0))+f(0)\] or \[f(f(0))=0\] Plugging in $x$ as $0:$ \[f(-y)+2yf(0)=f(f(0))+f(y),\] but since $f(f(0))=0,$ \begin{equation} f(-y)+2yf(0)=f(y) \text{ } (2) \end{equation} Plugging in $y^2$ instead of $y$ in the given equation: \[f(x^2-y^2)+2y^2f(x)=f(f(x))+f(y^2)\] Replacing $y$ and $x$ : \[f(y^2-x^2)+2x^2f(y)=f(f(y))+f(x^2)\] The difference would be: \begin{equation} f(x^2-y^2)-f(y^2-x^2)+2y^2f(x)-2x^2f(y)=f(f(x))-f(x^2)-f(f(y))-f(y^2) \text{ } (3) \end{equation} The right-hand side would be $f(0)-f(0)=0$ by $(1).$ Also, \[f(x^2-y^2)-f(y^2-x^2)=2(x^2-y^2)f(0)\] by $(2)$ So, $(3)$ is reduced to: \[2(x^2-y^2)f(0)+2y^2f(x)-2x^2f(y)=0\] Regrouping and dividing by 2: \[y^2(f(x)-f(0))=x^2(f(y)-f(0))\] \[\frac{f(x)-f(0)}{x^2}=\frac{f(y)-f(0)}{y^2}\] Because this holds for all x and y, $\frac{f(x)-f(0)}{x^2}$ is a constant. So, $f(x)=cx^2+f(0)$ . This function must be even, so $f(y)-f(-y)=0$ . So, along with $(2)$ , $2yf(0)=0$ for all $y$ , so $f(0)=0$ , and $f(x)=cx^2$ . Plugging in $cx^2$ for $f(x)$ in the original equation, we get: \[c(x^4-2x^2y+y^2)+2cx^2y=c^3x^4+cy^2\] \[c(x^4+y^2)=c(c^2x^4+y^2)\] So, $c=0$ or $c^2=1.$ All of these solutions work, so the solutions are $f(x)=-x^2, 0, x^2$ . -codemaster11
\[ f(x) = -x^2, \quad f(x) = 0, \quad f(x) = x^2 \]
usajmo
[ "Mathematics -> Number Theory -> Congruences" ]
8
Does there exists a positive irrational number ${x},$ such that there are at most finite positive integers ${n},$ satisfy that for any integer $1\leq k\leq n,$ $\{kx\}\geq\frac 1{n+1}?$
To determine whether there exists a positive irrational number \( x \) such that there are at most finitely many positive integers \( n \) satisfying the condition that for any integer \( 1 \leq k \leq n \), \( \{kx\} \geq \frac{1}{n+1} \), we proceed as follows: Assume for contradiction that there exists such an \( x \). This would imply that there exists a positive integer \( N \) such that for all \( n > N \), the inequality \( \{nx\} > \frac{1}{n+1} \) holds. However, by Dirichlet's approximation theorem, for any irrational number \( x \) and any positive integer \( n \), there exists an integer \( k \) such that \( 1 \leq k \leq n \) and \( \{kx\} < \frac{1}{n+1} \). This contradicts our assumption. Therefore, no such positive irrational number \( x \) exists. The answer is: \boxed{\text{No}}.
\text{No}
china_team_selection_test
[ "Mathematics -> Geometry -> Plane Geometry -> Polygons" ]
6.5
Determine the smallest positive real number $ k$ with the following property. Let $ ABCD$ be a convex quadrilateral, and let points $ A_1$, $ B_1$, $ C_1$, and $ D_1$ lie on sides $ AB$, $ BC$, $ CD$, and $ DA$, respectively. Consider the areas of triangles $ AA_1D_1$, $ BB_1A_1$, $ CC_1B_1$ and $ DD_1C_1$; let $ S$ be the sum of the two smallest ones, and let $ S_1$ be the area of quadrilateral $ A_1B_1C_1D_1$. Then we always have $ kS_1\ge S$. [i]Author: Zuming Feng and Oleg Golberg, USA[/i]
To determine the smallest positive real number \( k \) such that for any convex quadrilateral \( ABCD \) with points \( A_1 \), \( B_1 \), \( C_1 \), and \( D_1 \) on sides \( AB \), \( BC \), \( CD \), and \( DA \) respectively, the inequality \( kS_1 \ge S \) holds, where \( S \) is the sum of the areas of the two smallest triangles among \( \triangle AA_1D_1 \), \( \triangle BB_1A_1 \), \( \triangle CC_1B_1 \), and \( \triangle DD_1C_1 \), and \( S_1 \) is the area of quadrilateral \( A_1B_1C_1D_1 \), we proceed as follows: We need to show that \( k = 1 \) is the smallest such number. Consider the case where the points \( A_1 \), \( B_1 \), \( C_1 \), and \( D_1 \) are chosen such that the quadrilateral \( A_1B_1C_1D_1 \) is very close to a medial configuration. In this configuration, the areas of the triangles \( \triangle AA_1D_1 \), \( \triangle BB_1A_1 \), \( \triangle CC_1B_1 \), and \( \triangle DD_1C_1 \) can be made arbitrarily small compared to the area of \( A_1B_1C_1D_1 \). By examining degenerate cases and applying geometric transformations, it can be shown that the ratio \( \frac{S}{S_1} \) can approach 1. Therefore, we have \( S_1 \ge S \), which implies \( k = 1 \) is the smallest possible value that satisfies the inequality \( kS_1 \ge S \) for all configurations of the quadrilateral \( ABCD \) and points \( A_1 \), \( B_1 \), \( C_1 \), and \( D_1 \). Thus, the smallest positive real number \( k \) with the given property is: \[ \boxed{1} \]
1
usa_team_selection_test
[ "Mathematics -> Algebra -> Algebra -> Algebraic Expressions", "Mathematics -> Algebra -> Algebra -> Polynomial Operations" ]
5
Let $\frac{x^2+y^2}{x^2-y^2} + \frac{x^2-y^2}{x^2+y^2} = k$ . Compute the following expression in terms of $k$ : \[E(x,y) = \frac{x^8 + y^8}{x^8-y^8} - \frac{ x^8-y^8}{x^8+y^8}.\]
To start, we add the two fractions and simplify. \begin{align*} k &= \frac{(x^2+y^2)^2 + (x^2-y^2)^2}{x^4-y^4} \\ &= \frac{2x^4 + 2y^4}{x^4 - y^4}. \end{align*} Dividing both sides by two yields \[\frac{k}{2} = \frac{x^4 + y^4}{x^4 - y^4}.\] That means \begin{align*} \frac{x^4 + y^4}{x^4 - y^4} + \frac{x^4 - y^4}{x^4 + y^4} &= \frac{k}{2} + \frac{2}{k} \\ \frac{(x^4 + y^4)^2 + (x^4 - y^4)^2}{x^8 - y^8} &= \frac{k^2 + 4}{2k} \\ \frac{2x^8 + 2y^8}{x^8 - y^8} &= \frac{k^2 + 4}{2k}. \end{align*} Dividing both sides by two yields \[\frac{x^8 + y^8}{x^8 - y^8} = \frac{k^2 + 4}{4k}.\] That means \begin{align*} \frac{x^8 + y^8}{x^8 - y^8} - \frac{x^8 - y^8}{x^8 + y^8} &= \frac{k^2 + 4}{4k} - \frac{4k}{k^2 + 4} \\ &= \frac{k^4 + 8k^2 + 16 - 16k^2}{4k(k^2 + 4)} \\ &= \frac{k^4 - 8k^2 + 16}{4k(k^2 + 4)} \\ &= \boxed{\frac{(k^2 - 4)^2}{4k(k^2 + 4)}}. \end{align*}
\[ \boxed{\frac{(k^2 - 4)^2}{4k(k^2 + 4)}} \]
jbmo
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
5
Let $A_{n}=\{a_{1}, a_{2}, a_{3}, \ldots, a_{n}, b\}$, for $n \geq 3$, and let $C_{n}$ be the 2-configuration consisting of \( \{a_{i}, a_{i+1}\} \) for all \( 1 \leq i \leq n-1, \{a_{1}, a_{n}\} \), and \( \{a_{i}, b\} \) for \( 1 \leq i \leq n \). Let $S_{e}(n)$ be the number of subsets of $C_{n}$ that are consistent of order $e$. Find $S_{e}(101)$ for $e=1,2$, and 3.
For convenience, we assume the \( a_{i} \) are indexed modulo 101, so that \( a_{i+1}=a_{1} \) when \( a_{i}=a_{101} \). In any consistent subset of \( C_{101} \) of order 1, \( b \) must be paired with exactly one \( a_{i} \), say \( a_{1} \). Then, \( a_{2} \) cannot be paired with \( a_{1} \), so it must be paired with \( a_{3} \), and likewise we find we use the pairs \( \{a_{4}, a_{5}\}, \{a_{6}, a_{7}\}, \ldots, \{a_{100}, a_{101}\} \) - and this does give us a consistent subset of order 1. Similarly, pairing \( b \) with any other \( a_{i} \) would give us a unique extension to a consistent configuration of order 1. Thus, we have one such 2-configuration for each \( i \), giving \( S_{1}(101)=101 \) altogether. In a consistent subset of order 2, \( b \) must be paired with two other elements. Suppose one of them is \( a_{i} \). Then \( a_{i} \) is also paired with either \( a_{i-1} \) or \( a_{i+1} \), say \( a_{i+1} \). But then \( a_{i-1} \) needs to be paired up with two other elements, and \( a_{i} \) is not available, so it must be paired with \( a_{i-2} \) and \( b \). Now \( b \) has its two pairs determined, so nothing else can be paired with \( b \). Thus, for \( j \neq i-1, i \), we have that \( a_{j} \) must be paired with \( a_{j-1} \) and \( a_{j+1} \). So our subset must be of the form \( \{\{b, a_{i}\}, \{a_{i}, a_{i+1}\}, \{a_{i+1}, a_{i+2}\}, \ldots, \{a_{101}, a_{1}\}, \ldots, \{a_{i-2}, a_{i-1}\}, \{a_{i-1}, b\}\} \) for some \( i \). On the other hand, for any \( i=1, \ldots, 101 \), this gives a subset meeting our requirements. So, we have 101 possibilities, and \( S_{2}(101)=101 \). Finally, in a consistent subset of order 3, each \( a_{i} \) must be paired with \( a_{i-1}, a_{i+1} \), and \( b \). But then \( b \) occurs in 101 pairs, not just 3, so we have a contradiction. Thus, no such subset exists, so \( S_{3}(101)=0 \).
\[ S_{1}(101) = 101, \quad S_{2}(101) = 101, \quad S_{3}(101) = 0 \]
HMMT_2
[ "Mathematics -> Number Theory -> Other", "Mathematics -> Algebra -> Other" ]
7
The geometric mean of any set of $m$ non-negative numbers is the $m$ -th root of their product. $\quad (\text{i})\quad$ For which positive integers $n$ is there a finite set $S_n$ of $n$ distinct positive integers such that the geometric mean of any subset of $S_n$ is an integer? $\quad (\text{ii})\quad$ Is there an infinite set $S$ of distinct positive integers such that the geometric mean of any finite subset of $S$ is an integer?
a) We claim that for any numbers $p_1$ , $p_2$ , ... $p_n$ , $p_1^{n!}, p_2^{n!}, ... p_n^{n!}$ will satisfy the condition, which holds for any number $n$ . Since $\sqrt[n] ab = \sqrt[n] a * \sqrt[n] b$ , we can separate each geometric mean into the product of parts, where each part is the $k$ th root of each member of the subset and the subset has $k$ members. Assume our subset has $k$ members. Then, we know that the $k$ th root of each of these members is an integer (namely $p^{n!/k}$ ), because $k \leq n$ and thus $k | n!$ . Since each root is an integer, the geometric mean will also be an integer. b) If we define $q$ as an arbitrarily large number, and $x$ and $y$ as numbers in set $S$ , we know that ${\sqrt[q]{\frac{x}{y}}}$ is irrational for large enough $q$ , meaning that it cannot be expressed as the fraction of two integers. However, both the geometric mean of the set of $x$ and $q-1$ other arbitrary numbers in $S$ and the set of $y$ and the same other $q-1$ numbers are integers, so since the other numbers cancel out, the geometric means divided, or ${\sqrt[q]{\frac{x}{y}}}$ , must be rational. This is a contradiction, so no such infinite $S$ is possible. -aops111 (first solution dont bully me)
\[ \text{(i)} \quad \text{For all positive integers } n. \] \[ \text{(ii)} \quad \text{No, there is no such infinite set } S. \]
usamo
[ "Mathematics -> Algebra -> Algebra -> Equations and Inequalities", "Mathematics -> Algebra -> Intermediate Algebra -> Other" ]
7
Find all real numbers $x,y,z\geq 1$ satisfying \[\min(\sqrt{x+xyz},\sqrt{y+xyz},\sqrt{z+xyz})=\sqrt{x-1}+\sqrt{y-1}+\sqrt{z-1}.\]
The key Lemma is: \[\sqrt{a-1}+\sqrt{b-1} \le \sqrt{ab}\] for all $a,b \ge 1$ . Equality holds when $(a-1)(b-1)=1$ . This is proven easily. \[\sqrt{a-1}+\sqrt{b-1} = \sqrt{a-1}\sqrt{1}+\sqrt{1}\sqrt{b-1} \le \sqrt{(a-1+1)(b-1+1)} = \sqrt{ab}\] by Cauchy. Equality then holds when $a-1 =\frac{1}{b-1} \implies (a-1)(b-1) = 1$ . Now assume that $x = \min(x,y,z)$ . Now note that, by the Lemma, \[\sqrt{x-1}+\sqrt{y-1}+\sqrt{z-1} \le \sqrt{x-1} + \sqrt{yz} \le \sqrt{x(yz+1)} = \sqrt{xyz+x}\] . So equality must hold in order for the condition in the problem statement to be met. So $(y-1)(z-1) = 1$ and $(x-1)(yz) = 1$ . If we let $z = c$ , then we can easily compute that $y = \frac{c}{c-1}, x = \frac{c^2+c-1}{c^2}$ . Now it remains to check that $x \le y, z$ . But by easy computations, $x = \frac{c^2+c-1}{c^2} \le c = z \Longleftrightarrow (c^2-1)(c-1) \ge 0$ , which is obvious. Also $x = \frac{c^2+c-1}{c^2} \le \frac{c}{c-1} = y \Longleftrightarrow 2c \ge 1$ , which is obvious, since $c \ge 1$ . So all solutions are of the form $\boxed{\left(\frac{c^2+c-1}{c^2}, \frac{c}{c-1}, c\right)}$ , and all permutations for $c > 1$ . Remark: An alternative proof of the key Lemma is the following: By AM-GM, \[(ab-a-b+1)+1 = (a-1)(b-1) + 1 \ge 2\sqrt{(a-1)(b-1)}\] \[ab\ge (a-1)+(b-1)+2\sqrt{(a-1)(b-1)}\] . Now taking the square root of both sides gives the desired. Equality holds when $(a-1)(b-1) = 1$ .
\[ \boxed{\left(\frac{c^2+c-1}{c^2}, \frac{c}{c-1}, c\right)} \]
usamo
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations" ]
6
Find, with proof, all nonconstant polynomials $P(x)$ with real coefficients such that, for all nonzero real numbers $z$ with $P(z) \neq 0$ and $P\left(\frac{1}{z}\right) \neq 0$, we have $$\frac{1}{P(z)}+\frac{1}{P\left(\frac{1}{z}\right)}=z+\frac{1}{z}$$
Answer: $\quad P(x)=\frac{x\left(x^{4 k+2}+1\right)}{x^{2}+1}$ or $P(x)=\frac{x\left(1-x^{4 k}\right)}{x^{2}+1}$ Solution: It is straightforward to plug in and verify the above answers. Hence, we focus on showing that these are all possible solutions. The key claim is the following. Claim: If $r \neq 0$ is a root of $P(z)$ with multiplicity $n$, then $1 / r$ is also a root of $P(z)$ with multiplicity $n$. Proof 1 (Elementary). Let $n^{\prime}$ be the multiplicity of $1 / r$. It suffices to show that $n \leq n^{\prime}$ because we can apply the same assertion on $1 / r$ to obtain that $n^{\prime} \leq n$. To that end, suppose that $(z-r)^{n}$ divides $P(z)$. From the equation, we have $$z^{N}\left[P\left(\frac{1}{z}\right)+P(z)\right]=z^{N}\left[\left(z+\frac{1}{z}\right) P(z) P\left(\frac{1}{z}\right)\right]$$ where $N \gg \operatorname{deg} P+1$ to guarantee that both sides are polynomial. Notice that the factor $z^{N} P(z)$ and the right-hand side is divisible by $(z-r)^{n}$, so $(z-r)^{n}$ must also divide $z^{N} P\left(\frac{1}{z}\right)$. This means that there exists a polynomial $Q(z)$ such that $z^{N} P\left(\frac{1}{z}\right)=(z-r)^{n} Q(z)$. Replacing $z$ with $\frac{1}{z}$, we get $$\frac{P(z)}{z^{N}}=\left(\frac{1}{z}-r\right)^{n} Q\left(\frac{1}{z}\right) \Longrightarrow P(z)=z^{N-n}(1-r z)^{n} Q\left(\frac{1}{z}\right)$$ implying that $P(z)$ is divisible by $(z-1 / r)^{n}$. Proof 2 (Complex Analysis). Here is more advanced proof of the main claim. View both sides of the equations as meromorphic functions in the complex plane. Then, a root $r$ with multiplicity $n$ of $P(z)$ is a pole of $\frac{1}{P(z)}$ of order $n$. Since the right-hand side is analytic around $r$, it follows that the other term $\frac{1}{P(1 / z)}$ has a pole at $r$ with order $n$ as well. By replacing $z$ with $1 / z$, we find that $\frac{1}{P(z)}$ has a pole at $1 / r$ of order $n$. This finishes the claim. The claim implies that there exists an integer $k$ and a constant $\epsilon$ such that $$P(z)=\epsilon z^{k} P\left(\frac{1}{z}\right)$$ By replacing $z$ with $1 / z$, we get that $$z^{k} P\left(\frac{1}{z}\right)=\epsilon P(z)$$ Therefore, $\epsilon= \pm 1$. Moreover, using the main equation, we get that $$\frac{1}{P(z)}+\frac{\epsilon z^{k}}{P(z)}=z+\frac{1}{z} \Longrightarrow P(z)=\frac{z\left(1+\epsilon z^{k}\right)}{1+z^{2}}$$ This is a polynomial if and only if $(\epsilon=1$ and $k \equiv 2(\bmod 4))$ or $(\epsilon=-1$ and $k \equiv 0(\bmod 4))$, so we are done.
\[ P(x) = \frac{x\left(x^{4k+2}+1\right)}{x^{2}+1} \quad \text{or} \quad P(x) = \frac{x\left(1-x^{4k}\right)}{x^{2}+1} \]
HMMT_2
[ "Mathematics -> Number Theory -> Prime Numbers" ]
5
Find all positive integers $n\geq 1$ such that $n^2+3^n$ is the square of an integer.
After rearranging we get: $(k-n)(k+n) = 3^n$ Let $k-n = 3^a, k+n = 3^{n-a}$ we get: $2n = 3^a(3^{n-2a} - 1)$ or, $(2n/(3^a)) + 1 = 3^{n-2a}$ Now, it is clear from above that $3^a$ divides $n$ . so, $n \geq 3^a$ If $n = 3^a, n - 2a = 3^a - 2a \geq 1$ so $RHS \geq 3$ But $LHS = 3$ If $n > 3^a$ then $RHS$ increases exponentially compared to $LHS$ so $n$ cannot be $> 3^a$ . Thus $n = 3^a$ . Substituting value of $n$ above we get: $3 = 3^{3^a - 2a}$ or $3^a - 2a = 1$ this results in only $a = 0$ or $a = 1$ Thus $n = 1$ or $3$ . ~Kris17
The positive integers \( n \geq 1 \) such that \( n^2 + 3^n \) is the square of an integer are \( n = 1 \) and \( n = 3 \).
jbmo
[ "Mathematics -> Geometry -> Plane Geometry -> Angles" ]
6
The feet of the angle bisectors of $\Delta ABC$ form a right-angled triangle. If the right-angle is at $X$ , where $AX$ is the bisector of $\angle A$ , find all possible values for $\angle A$ .
This problem needs a solution. If you have a solution for it, please help us out by adding it .
The problem provided does not contain a solution. Therefore, no final answer can be extracted.
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Polygons" ]
5.25
For a convex quadrilateral $P$, let $D$ denote the sum of the lengths of its diagonals and let $S$ denote its perimeter. Determine, with proof, all possible values of $\frac{S}{D}$.
Suppose we have a convex quadrilateral $A B C D$ with diagonals $A C$ and $B D$ intersecting at $E$. To prove the lower bound, note that by the triangle inequality, $A B+B C>A C$ and $A D+D C>A C$, so $S=A B+B C+A D+D C>2 A C$. Similarly, $S>2 B D$, so $2 S>2 A C+2 B D=2 D$ gives $S>D$. To prove the upper bound, note that again by the triangle inequality, $A E+E B>A B, C E+B E>B C, A E+E D>A D, C E+E D>C D$. Adding these yields $2(A E+E C+B E+E D)>A B+B C+A D+C D=S$. Now since $A B C D$ is convex, $E$ is inside the quadrilateral, so $A E+E C=A C$ and $B E+E D=B D$. Thus $2(A C+B D)=D>S$. To achieve every real value in this range, first consider a square $A B C D$. This has $\frac{S}{D}=\sqrt{2}$. Suppose now that we have a rectangle with $A B=C D=1$ and $B C=A D=x$, where $0<x \leq 1$. As $x$ approaches 0, $\frac{S}{D}$ gets arbitrarily close to 1, so by intermediate value theorem, we hit every value $\frac{S}{D} \in(1, \sqrt{2}]$. To achieve the other values, we let $A B=B C=C D=D A=1$ and let $\theta=m \angle A B E$ vary from $45^{\circ}$ down to $0^{\circ}$ (i.e. a rhombus that gets thinner). This means $A C=2 \sin \theta$ and $B D=2 \cos \theta$. We have $S=4$ and $D=2(\sin \theta+\cos \theta)$. When $\theta=45^{\circ}, \frac{S}{D}=\sqrt{2}$, and when $\theta=0^{\circ}, \frac{S}{D}=2$. Thus by the intermediate value theorem, we are able to choose $\theta$ to obtain any value in the range $[\sqrt{2}, 2)$. Putting this construction together with the strict upper and lower bounds, we find that all possible values of $\frac{S}{D}$ are all real values in the open interval $(1,2)$.
The possible values of $\frac{S}{D}$ for a convex quadrilateral are all real values in the open interval $(1, 2)$.
HMMT_2
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
8
Let $S$ be the set of $10$-tuples of non-negative integers that have sum $2019$. For any tuple in $S$, if one of the numbers in the tuple is $\geq 9$, then we can subtract $9$ from it, and add $1$ to the remaining numbers in the tuple. Call thus one operation. If for $A,B\in S$ we can get from $A$ to $B$ in finitely many operations, then denote $A\rightarrow B$. (1) Find the smallest integer $k$, such that if the minimum number in $A,B\in S$ respectively are both $\geq k$, then $A\rightarrow B$ implies $B\rightarrow A$. (2) For the $k$ obtained in (1), how many tuples can we pick from $S$, such that any two of these tuples $A,B$ that are distinct, $A\not\rightarrow B$.
### Part 1: We need to find the smallest integer \( k \) such that if the minimum number in \( A, B \in S \) are both \(\geq k\), then \( A \rightarrow B \) implies \( B \rightarrow A \). We claim that the smallest integer \( k \) is \( 8 \). **Proof:** 1. **\( k \leq 7 \) does not satisfy the condition:** Consider the counterexample \( A = (1956, 7, 7, 7, 7, 7, 7, 7, 7, 7) \) and \( B = (1938, 9, 9, 9, 9, 9, 9, 9, 9, 9) \). It is clear that \( A \rightarrow B \). However, \( B \not\rightarrow A \) because each of the entries from the 2nd to the 10th in \( B \) must be subtracted by 9 at least once. After the last subtraction, the remaining entries should be greater than or equal to 8, which contradicts the condition. 2. **\( k = 8 \) does satisfy the condition:** Denote the numbers in the \( i \)-th entry as \( x_i \). For any \( i \) and \( j \), \( x_i - x_j \mod 10 \) is conserved. Assume \( 8 \leq a_1 \leq a_2 \leq \cdots \leq a_{10} \) for \( A \). We need to show that for each \( i \), we can operate on \( B \) so that \( x_1 - a_1 = x_2 - a_2 = \cdots = x_i - a_i \). We prove this by induction on \( i \). The base case is trivial. Suppose \( x_1 - a_1 = x_2 - a_2 = \cdots = x_i - a_i \). Since \( x_i - x_j \mod 10 \) is conserved and \( A \rightarrow B \), \( x_i - a_i \mod 10 \) should be equal for all \( i \). Repeat operations on \( x_1, x_2, \cdots, x_{i+1} \) equal times so that \( x_1 \) or \( x_{i+1} \leq 8 \). If \( x_{i+1} - a_{i+1} < x_i - a_i \), let \( t = \frac{(x_i - a_i) - (x_{i+1} - a_{i+1})}{10} \). After some calculations, we can subtract 9 from all \( x_j (j \neq i+1) \) \( t \) times, including necessary operations subtracting 9 from \( x_j (j \geq i+2) \), while not subtracting 9 from \( x_{i+1} \). If \( x_{i+1} - a_{i+1} > x_i - a_i \), proceed similarly. Thus, the smallest \( k \) is \( 8 \). ### Part 2: For the \( k \) obtained in Part 1, we need to find how many tuples can be picked from \( S \) such that any two distinct tuples \( A, B \) satisfy \( A \not\rightarrow B \). We have practically shown that \( A \rightarrow B \) is equivalent to \( a_1 - b_1 \equiv a_2 - b_2 \equiv \cdots \equiv a_{10} - b_{10} \pmod{10} \). We need to count the number of tuples that cannot be derived from each other, ensuring \( x_1 + x_2 + \cdots + x_{10} = 2019 \equiv 9 \pmod{10} \). The number of such tuples is \( 10^8 \). The answer is: \boxed{10^8}.
10^8
china_team_selection_test
[ "Mathematics -> Discrete Mathematics -> Combinatorics", "Mathematics -> Applied Mathematics -> Statistics -> Probability -> Other" ]
8
A blackboard contains 68 pairs of nonzero integers. Suppose that for each positive integer $k$ at most one of the pairs $(k, k)$ and $(-k, -k)$ is written on the blackboard. A student erases some of the 136 integers, subject to the condition that no two erased integers may add to 0. The student then scores one point for each of the 68 pairs in which at least one integer is erased. Determine, with proof, the largest number $N$ of points that the student can guarantee to score regardless of which 68 pairs have been written on the board.
Answer: 43 Attainability: Consider 8 distinct positive numbers. Let there be 5 pairs for each of the numbers including 2 clones of that number. Let there also be 28 pairs that include the negatives of those numbers such that each negative associates with another negative once and exactly once (in graph theoretic terms, a K8). Let x be the number of positives chosen out of these 8 (assume the other chosen numbers are negatives) for erasing; then, the number of points the student scores is $28 + 5x - {x \choose 2}$ , which is maximized at x=5 and x=6, and the maximum value is $43$ . Choosing the first 5 numbers as positive and the other three as negative attains this. Hence, 43 is a possible maximum possible score. Bounding: We use expected values. WLOG all the pairs of numbers with both numbers identical have only positive values. Consider flipping a weighted coin whether to choose the positive number or its negation for each positive number; it chooses positive with probability p. The pairs with both numbers same are chosen with probability p, the pairs (k, -k) are chosen with probability 1, and the pairs (x, y) for distinct x and y that sum to a nonzero number are chosen with probability $1-p^2$ . We are trying to minimize the expected value, so we can assume that no pairs (k, -k) exist. Let A be the number of (k, k) pairs, and 68-A be the number of (x, y) pairs. The expected number of points scored is $Ap + (68-A)(1-p^2)$ . We want to prove this is larger than 42 at all times for some choice of p. If $A < 36$ , $1/2$ works for p to give this bound. If $A > 36$ , $5/8$ works for p for p to give the desired bound. If $A = 36$ , we can use $3/5$ for p to get the desired bound. Hence, in any case, the expected value for the number of points scored across all of these weighted processes is larger than 42, so there exists some case that gives a score of 43. Hence, bounding is complete. We are done with both parts. Q.E.D. -Solution by thanosaops
\[ 43 \]
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Other" ]
5
Let $S$ be the set of all points in the plane whose coordinates are positive integers less than or equal to 100 (so $S$ has $100^{2}$ elements), and let $\mathcal{L}$ be the set of all lines $\ell$ such that $\ell$ passes through at least two points in $S$. Find, with proof, the largest integer $N \geq 2$ for which it is possible to choose $N$ distinct lines in $\mathcal{L}$ such that every two of the chosen lines are parallel.
Let the lines all have slope $\frac{p}{q}$ where $p$ and $q$ are relatively prime. Without loss of generality, let this slope be positive. Consider the set of points that consists of the point of $S$ with the smallest coordinates on each individual line in the set $L$. Consider a point $(x, y)$ in this, because there is no other point in $S$ on this line with smaller coordinates, either $x \leq q$ or $y \leq p$. Additionally, since each line passes through at least two points in $S$, we need $x+q \leq 100$ and $y+p \leq 100$. The shape of this set of points will then be either a rectangle from $(1,1)$ to $(100-q, 100-p)$ with the rectangle from $(q+1, p+1)$ to $(100-q, 100-p)$ removed, or if $100-q<q+1$ or $100-p<p+1$, just the initial rectangle. This leads us to two formulas for the number of lines, $$N= \begin{cases}(100-p)(100-q)-(100-2 p)(100-2 q) & p, q<50 \\ (100-p)(100-q) & \text { otherwise }\end{cases}$$ In the first case, we need to minimize the quantity $$(100-p)(100-q)-(100-2 p)(100-2 q) =100(p+q)-3 p q =\frac{10000}{3}-3\left(q-\frac{100}{3}\right)\left(p-\frac{100}{3}\right)$$ if one of $p, q$ is above $100 / 3$ and the other is below it, we would want to maximize how far these two are from $100 / 3$. The case $(p, q)=(49,1)$ will be the optimal case since all other combinations will have $p, q$ 's closer to $100 / 3$, this gives us 4853 cases. In the second case, we need to minimize $p$ and $q$ while keeping at least one above 50 and them relatively prime. From here we need only check $(p, q)=(50,1)$ since for all other cases, we can reduce either $p$ or $q$ to increase the count. This case gives a maximum of 4950.
4950
HMMT_2
[ "Mathematics -> Algebra -> Abstract Algebra -> Field Theory" ]
6
Find all functions $ f: \mathbb{Q}^{\plus{}} \mapsto \mathbb{Q}^{\plus{}}$ such that: \[ f(x) \plus{} f(y) \plus{} 2xy f(xy) \equal{} \frac {f(xy)}{f(x\plus{}y)}.\]
Let \( f: \mathbb{Q}^{+} \to \mathbb{Q}^{+} \) be a function such that: \[ f(x) + f(y) + 2xy f(xy) = \frac{f(xy)}{f(x+y)} \] for all \( x, y \in \mathbb{Q}^{+} \). First, we denote the assertion of the given functional equation as \( P(x, y) \). 1. From \( P(1, 1) \), we have: \[ f(1) + f(1) + 2 \cdot 1 \cdot 1 \cdot f(1) = \frac{f(1)}{f(2)} \] \[ 4f(1) = \frac{f(1)}{f(2)} \] \[ f(2) = \frac{1}{4} \] 2. From \( P(n, 1) \), we get: \[ f(n) + f(1) + 2n f(n) = \frac{f(n)}{f(n+1)} \] \[ f(n+1) = \frac{f(n)}{(2n+1)f(n) + f(1)} \] 3. Using the above recurrence relation, we find: \[ f(3) = \frac{f(2)}{(2 \cdot 2 + 1)f(2) + f(1)} = \frac{\frac{1}{4}}{5 \cdot \frac{1}{4} + f(1)} = \frac{1}{5 + 4f(1)} \] \[ f(4) = \frac{f(3)}{(2 \cdot 3 + 1)f(3) + f(1)} = \frac{\frac{1}{5 + 4f(1)}}{7 \cdot \frac{1}{5 + 4f(1)} + f(1)} = \frac{1}{7 + 5f(1) + 4f(1)^2} \] 4. From \( P(2, 2) \), we have: \[ f(2) + f(2) + 2 \cdot 2 \cdot 2 \cdot f(4) = \frac{f(4)}{f(4)} \] \[ 2f(2) + 8f(4) = 1 \] \[ 2 \cdot \frac{1}{4} + 8f(4) = 1 \] \[ 8f(4) = \frac{1}{2} \] \[ f(4) = \frac{1}{16} \] 5. Equating the two expressions for \( f(4) \), we get: \[ \frac{1}{7 + 5f(1) + 4f(1)^2} = \frac{1}{16} \] \[ 7 + 5f(1) + 4f(1)^2 = 16 \] \[ 4f(1)^2 + 5f(1) - 9 = 0 \] \[ f(1) = 1 \] 6. Substituting \( f(1) = 1 \) into the recurrence relation, we get: \[ f(n+1) = \frac{f(n)}{(2n+1)f(n) + 1} \] By induction, we can show that \( f(n) = \frac{1}{n^2} \) for all \( n \in \mathbb{N} \). 7. Now, we claim that \( f(x) = \frac{1}{x^2} \) for all \( x \in \mathbb{Q}^{+} \). Using the functional equation and induction, we can extend this result to all positive rational numbers. Thus, the function that satisfies the given functional equation is: \[ f(x) = \frac{1}{x^2} \] The answer is: \boxed{\frac{1}{x^2}}.
\frac{1}{x^2}
china_team_selection_test
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
8
Consider an $n$ -by- $n$ board of unit squares for some odd positive integer $n$ . We say that a collection $C$ of identical dominoes is a maximal grid-aligned configuration on the board if $C$ consists of $(n^2-1)/2$ dominoes where each domino covers exactly two neighboring squares and the dominoes don't overlap: $C$ then covers all but one square on the board. We are allowed to slide (but not rotate) a domino on the board to cover the uncovered square, resulting in a new maximal grid-aligned configuration with another square uncovered. Let $k(C)$ be the number of distinct maximal grid-aligned configurations obtainable from $C$ by repeatedly sliding dominoes. Find the maximum value of $k(C)$ as a function of $n$ .
We claim the answer is $(\frac{n+1}{2})^2$ . First, consider a checkerboard tiling of the board with 4 colors: R, G, B, Y. Number each column from $1$ to $n$ from left to right and each row from $1$ to $n$ from top to bottom. We color a tile R if its row and column are odd, a tile G is its row is even but its column is odd, a tile B if its row and column is even, and a tile Y if its row is odd but its column is even. Lemma 1: Throughout our moves, the color of the uncolored tile stays an invariant. Consider that a domino can either only change rows or can only change columns. Therefore, sliding a domino into the hole and creating a new one has two possible colors. Of these, note that the new hole will always trivially be two tiles away from the old hole, meaning that the parity of both the row and column number stays the same. Thus, the lemma holds. Lemma 2: There are more red tiles than any other color. Because each color is uniquely defined by the parity of a pair of column and row number, it satisfies to show that given an odd integer $n$ , there are more odd positive integers less than or equal to $n$ than even ones. Obviously, this is true, and so red will have more tiles than any other color. Lemma 3: For any starting configuration $C$ and any blank tile $B$ such that the blank tile's color matches the blank tile's color of $C$ , there is no more than one unique configuration $C'$ that can be produced from $C$ using valid moves. We will use proof by contradiction. Assume there exists two different $C'$ . We can get from one of these $C'$ to another using moves. However, we have to finish off with the same hole as before. Before the last move, the hole must be two tiles away from the starting hole. However, because the domino we used to move into the blank tile's spot is in the way, that hole must be congruent to the hole produced after the first move. We can induct this logic, and because there is a limited amount of tiles with the same color, eventually we will run out of tiles to apply this to. Therefore, having two distinct $C'$ with the same starting hole $B$ is impossible with some $C$ . We will now prove that $(\frac{n+1}{2})^2$ is the answer. There are $\frac{n+1}{2}$ rows and $\frac{n+1}{2}$ columns that are odd, and thus there are $(\frac{n+1}{2})^2$ red tiles. Given lemma 3, this is our upper bound for a maximum. To establish that $(\frac{n+1}{2})^2$ is indeed possible, we construct such a $C$ : In the first column, leave the first tile up blank. Then, continuously fill in vertically oriented dominos in that column until it reaches the bottom. In the next $n-1$ columns, place $\frac{n-1}{2}$ vertically oriented dominos in a row starting from the top. At the bottom row, starting with the first unfilled tile on the left, place horizontally aligned dominos in a row until you reach the right. Obviously, the top left tile is red. It suffices to show that any red tile may be uncovered. For the first column, one may slide some dominos on the first column until the desired tile is uncovered. For the bottom row, all the first dominos may be slid up, and then the bottom dominos may be slid to the left until the desired red tile is uncovered. Finally, for the rest of the red tiles, the bottom red tile in the same color may be revealed, and then vertically aligned dominos in the same column may be slid down until the desired tile is revealed. Therefore, this configuration may produce $(\frac{n+1}{2})^2$ different configurations with moves. Hence, we have proved that $(\frac{n+1}{2})^2$ is the maximum, and we are done. $\blacksquare{}$ ~SigmaPiE
\[ \left(\frac{n+1}{2}\right)^2 \]
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Triangulations" ]
3.5
In triangle $A B C, A C=3 A B$. Let $A D$ bisect angle $A$ with $D$ lying on $B C$, and let $E$ be the foot of the perpendicular from $C$ to $A D$. Find $[A B D] /[C D E]$.
By the Angle Bisector Theorem, $D C / D B=A C / A B=3$. We will show that $A D=$ $D E$. Let $C E$ intersect $A B$ at $F$. Then since $A E$ bisects angle $A, A F=A C=3 A B$, and $E F=E C$. Let $G$ be the midpoint of $B F$. Then $B G=G F$, so $G E \| B C$. But then since $B$ is the midpoint of $A G, D$ must be the midpoint of $A E$, as desired. Then $[A B D] /[C D E]=(A D \cdot B D) /(E D \cdot C D)=1 / 3$.
\[\frac{[ABD]}{[CDE]} = \frac{1}{3}\]
HMMT_2
[ "Mathematics -> Algebra -> Intermediate Algebra -> Other", "Mathematics -> Algebra -> Algebra -> Polynomial Operations" ]
8
Find all functions $f:\mathbb{R}\rightarrow \mathbb{R}$ such that for all real numbers $x$ and $y$ , \[(f(x)+xy)\cdot f(x-3y)+(f(y)+xy)\cdot f(3x-y)=(f(x+y))^2.\]
Step 1: Set $x = y = 0$ to obtain $f(0) = 0.$ Step 2: Set $x = 0$ to obtain $f(y)f(-y) = f(y)^2.$ $\indent$ In particular, if $f(y) \ne 0$ then $f(y) = f(-y).$ $\indent$ In addition, replacing $y \to -t$ , it follows that $f(t) = 0 \implies f(-t) = 0$ for all $t \in \mathbb{R}.$ Step 3: Set $x = 3y$ to obtain $\left[f(y) + 3y^2\right]f(8y) = f(4y)^2.$ $\indent$ In particular, replacing $y \to t/8$ , it follows that $f(t) = 0 \implies f(t/2) = 0$ for all $t \in \mathbb{R}.$ Step 4: Set $y = -x$ to obtain $f(4x)\left[f(x) + f(-x) - 2x^2\right] = 0.$ $\indent$ In particular, if $f(x) \ne 0$ , then $f(4x) \ne 0$ by the observation from Step 3, because $f(4x) = 0 \implies f(2x) = 0 \implies f(x) = 0.$ Hence, the above equation implies that $2x^2 = f(x) + f(-x) = 2f(x)$ , where the last step follows from the first observation from Step 2. $\indent$ Therefore, either $f(x) = 0$ or $f(x) = x^2$ for each $x.$ $\indent$ Looking back on the equation from Step 3, it follows that $f(y) + 3y^2 \ne 0$ for any nonzero $y.$ Therefore, replacing $y \to t/4$ in this equation, it follows that $f(t) = 0 \implies f(2t) = 0.$ Step 5: If $f(a) = f(b) = 0$ , then $f(b - a) = 0.$ $\indent$ This follows by choosing $x, y$ such that $x - 3y = a$ and $3x - y = b.$ Then $x + y = \tfrac{b - a}{2}$ , so plugging $x, y$ into the given equation, we deduce that $f\left(\tfrac{b - a}{2}\right) = 0.$ Therefore, by the third observation from Step 4, we obtain $f(b - a) = 0$ , as desired. Step 6: If $f \not\equiv 0$ , then $f(t) = 0 \implies t = 0.$ $\indent$ Suppose by way of contradiction that there exists an nonzero $t$ with $f(t) = 0.$ Choose $x, y$ such that $f(x) \ne 0$ and $x + y = t.$ The following three facts are crucial: $\indent$ 1. $f(y) \ne 0.$ This is because $(x + y) - y = x$ , so by Step 5, $f(y) = 0 \implies f(x) = 0$ , impossible. $\indent$ 2. $f(x - 3y) \ne 0.$ This is because $(x + y) - (x - 3y) = 4y$ , so by Step 5 and the observation from Step 3, $f(x - 3y) = 0 \implies f(4y) = 0 \implies f(2y) = 0 \implies f(y) = 0$ , impossible. $\indent$ 3. $f(3x - y) \ne 0.$ This is because by the second observation from Step 2, $f(3x - y) = 0 \implies f(y - 3x) = 0.$ Then because $(x + y) - (y - 3x) = 4x$ , Step 5 together with the observation from Step 3 yield $f(3x - y) = 0 \implies f(4x) = 0 \implies f(2x) = 0 \implies f(x) = 0$ , impossible. $\indent$ By the second observation from Step 4, these three facts imply that $f(y) = y^2$ and $f(x - 3y) = \left(x - 3y\right)^2$ and $f(3x - y) = \left(3x - y\right)^2.$ By plugging into the given equation, it follows that \begin{align*} \left(x^2 + xy\right)\left(x - 3y\right)^2 + \left(y^2 + xy\right)\left(3x - y\right)^2 = 0. \end{align*} But the above expression miraculously factors into $\left(x + y\right)^4$ ! This is clearly a contradiction, since $t = x + y \ne 0$ by assumption. This completes Step 6. Step 7: By Step 6 and the second observation from Step 4, the only possible solutions are $f \equiv 0$ and $f(x) = x^2$ for all $x \in \mathbb{R}.$ It's easy to check that both of these work, so we're done.
The functions \( f \) that satisfy the given equation are: \[ f(x) = 0 \quad \text{and} \quad f(x) = x^2 \]
usamo
[ "Mathematics -> Algebra -> Abstract Algebra -> Group Theory" ]
6.5
A finite set $S$ of positive integers has the property that, for each $s \in S,$ and each positive integer divisor $d$ of $s$ , there exists a unique element $t \in S$ satisfying $\text{gcd}(s, t) = d$ . (The elements $s$ and $t$ could be equal.) Given this information, find all possible values for the number of elements of $S$ .
The answer is $0$ (left out by problem author) or $2^n$ for any non-negative integer $n$ . To construct $|S|=2^n$ , take $2n$ distinct primes $p_1, q_1, p_2, q_2, \dots p_n, q_n$ and construct $2^n$ elements for $S$ by taking the product over all $n$ indices $i$ of either $p_i$ or $q_i$ for every $i$ . Note that (from the problem statement) each element of $S$ has $|S|$ divisors. Now it suffices to show that there is no element $x \in S$ such that $x$ is divisible by more than one power of any prime. Suppose there does exist an element $x$ and a prime $p$ such that $p^2 \mid x$ . This implies that there exists an element $c \in S$ that is divisible by $p$ but not $p^2$ by $\gcd (x,c)=p$ . Exactly half of $c$ 's divisors are not divisible by $p$ , so it follows that exactly half of the elements in $S$ are not divisible by $p$ . However, if $p^0 , p^1 , \dots p^k$ are the powers of $p$ dividing $x$ for some $k \ge 2$ , only $\frac{1}{k+1}$ of the divisors of $x$ are not divisible by $p$ . But this means that $\frac{1}{k+1}$ of the elements of $S$ are not divisible by $p$ , contradiction. Therefore, $|S|$ must be some power of $2$ .
The number of elements of \( S \) is \( 2^n \) for any non-negative integer \( n \).
usajmo
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
6
Each cell of an $m\times n$ board is filled with some nonnegative integer. Two numbers in the filling are said to be adjacent if their cells share a common side. (Note that two numbers in cells that share only a corner are not adjacent). The filling is called a garden if it satisfies the following two conditions: (i) The difference between any two adjacent numbers is either $0$ or $1$ . (ii) If a number is less than or equal to all of its adjacent numbers, then it is equal to $0$ . Determine the number of distinct gardens in terms of $m$ and $n$ .
We claim that any configuration of $0$ 's produces a distinct garden. To verify this claim, we show that, for any cell that is nonzero, the value of that cell is its distance away from the nearest zero, where distance means the shortest chain of adjacent cells connecting two cells. Now, since we know that any cell with a nonzero value must have a cell adjacent to it that is less than its value, there is a path that goes from this cell to the $0$ that is decreasing, which means that the value of the cell must be its distance from the $0 \rightarrow$ as the path must end. From this, we realize that, for any configuration of $0$ 's, the value of each of the cells is simply its distance from the nearest $0$ , and therefore one garden is produced for every configuration of $0$ 's. However, we also note that there must be at least one $0$ in the garden, as otherwise the smallest number in the garden, which is less than or equal to all of its neighbors, is $>0$ , which violates condition $(ii)$ . There are $2^{mn}$ possible configurations of $0$ and not $0$ in the garden, one of which has no $0$ 's, so our total amount of configurations is $\boxed{2^{mn} -1}$
\boxed{2^{mn} - 1}
usajmo
[ "Mathematics -> Geometry -> Solid Geometry -> 3D Shapes" ]
8
Determine the greatest positive integer $ n$ such that in three-dimensional space, there exist n points $ P_{1},P_{2},\cdots,P_{n},$ among $ n$ points no three points are collinear, and for arbitary $ 1\leq i < j < k\leq n$, $ P_{i}P_{j}P_{k}$ isn't obtuse triangle.
To determine the greatest positive integer \( n \) such that in three-dimensional space, there exist \( n \) points \( P_{1}, P_{2}, \cdots, P_{n} \) where no three points are collinear and for any \( 1 \leq i < j < k \leq n \), the triangle \( P_{i}P_{j}P_{k} \) is not obtuse, we need to consider the geometric constraints. In three-dimensional space, the maximum number of points that can be arranged such that no three are collinear and no triangle formed by any three points is obtuse is 8. This arrangement can be visualized as the vertices of a cube. If we attempt to add a ninth point, it is inevitable that at least one of the triangles formed will be obtuse. This is because in any arrangement of more than 8 points, there will be at least one set of three points where the angle between two of the points exceeds \( \frac{\pi}{2} \). Therefore, the greatest positive integer \( n \) such that no three points are collinear and no triangle is obtuse is 8. The answer is: \(\boxed{8}\).
8
china_team_selection_test
[ "Mathematics -> Algebra -> Intermediate Algebra -> Functional Analysis" ]
7.5
A function $f: \mathbb{R}\to \mathbb{R}$ is $\textit{essentially increasing}$ if $f(s)\leq f(t)$ holds whenever $s\leq t$ are real numbers such that $f(s)\neq 0$ and $f(t)\neq 0$ . Find the smallest integer $k$ such that for any 2022 real numbers $x_1,x_2,\ldots , x_{2022},$ there exist $k$ essentially increasing functions $f_1,\ldots, f_k$ such that \[f_1(n) + f_2(n) + \cdots + f_k(n) = x_n\qquad \text{for every } n= 1,2,\ldots 2022.\]
Coming soon.
The smallest integer \( k \) is 2022.
usamo
[ "Mathematics -> Algebra -> Algebra -> Equations and Inequalities", "Mathematics -> Number Theory -> Other" ]
8
Find all positive real numbers $t$ with the following property: there exists an infinite set $X$ of real numbers such that the inequality \[ \max\{|x-(a-d)|,|y-a|,|z-(a+d)|\}>td\] holds for all (not necessarily distinct) $x,y,z\in X$, all real numbers $a$ and all positive real numbers $d$.
To find all positive real numbers \( t \) with the property that there exists an infinite set \( X \) of real numbers such that the inequality \[ \max\{|x-(a-d)|,|y-a|,|z-(a+d)|\} > td \] holds for all \( x, y, z \in X \), all real numbers \( a \), and all positive real numbers \( d \), we proceed as follows: Firstly, we show that for all \( x < y < z \in \mathbb{R} \), there exist \( a \in \mathbb{R} \) and \( d > 0 \) such that \[ \max\{|x-(a-d)|,|y-a|,|z-(a+d)|\} \leq \frac{1}{2}d. \] Assume without loss of generality that \( x-y < y-z \). Let \( c = \frac{1}{4}(x + y - 2z) \). Then there exist \( a \) and \( d \) such that: \[ \begin{align*} z - c &= a + d, \\ y + c &= a, \\ x - c &= a - d. \end{align*} \] Let \( k = \frac{y - x}{x + y - 2z} \). Then \[ \frac{c}{d} = \frac{1}{2 + 4k} < \frac{1}{2}. \] Conversely, we may choose \( X = \{1, m, m^2, \ldots, m^n, \ldots\} \), where \( m \) is very large. Suppose \( m^{n_1}, m^{n_2}, m^{n_3} \) are elements of \( X \). Suppose \( a_1, a_2, a_3 \) is an arithmetic sequence. Define \( c_i = m^{n_i} - a_i \). Then: \[ m^{n_1} + m^{n_3} - 2m^{n_2} = c_1 + c_3 - 2c_2 \leq |c_1| + |c_3| + 2|c_2|. \] Let \( d = a_2 - a_1 = m^{n_2} - c_2 - (m^{n_1} - c_1) \leq m^{n_2} - m^{n_1} + |c_2| + |c_1| \). Hence: \[ \frac{|c_1| + |c_3|}{d} \geq \frac{|c_1| + |c_3|}{m^{n_2} - m^{n_1} + |c_2| + |c_1|} \geq \frac{|c_1| + |c_2|}{(2k + 1)|c_2| + (k + 1)|c_1| + k|c_3|} \geq \frac{1}{2k + 1}. \] By the pigeonhole principle, the maximum of \[ \left\{\frac{|c_1|}{d}, \frac{|c_3|}{d}\right\} \] can be made arbitrarily close to \( \frac{1}{2} \). Therefore, the answer is: \boxed{t < \frac{1}{2}}.
t < \frac{1}{2}
china_national_olympiad
[ "Mathematics -> Algebra -> Algebra -> Equations and Inequalities", "Mathematics -> Number Theory -> Congruences" ]
4.5
At a tennis tournament there were $2n$ boys and $n$ girls participating. Every player played every other player. The boys won $\frac 75$ times as many matches as the girls. It is knowns that there were no draws. Find $n$ .
The total number of games played in the tournament is $\tfrac{3n(3n-1)}{2}.$ Since the boys won $\tfrac75$ as many matches as the girls, the boys won $\tfrac{7}{12}$ of all the games played, so the total number of games that a boy won is $\tfrac{7}{12} \cdot \tfrac{3n(3n-1)}{2} = \tfrac{7n(3n-1)}{8}.$ Since the number of games that a boy won is a whole number, $n(3n-1)$ must be a multiple of 8. Testing each residue, we find that $n \equiv 0,3 \pmod{8}.$ For $n$ to be valid, the number of games the boys won must be less than or equal to the number of games where a boy has played. The number of games with only girls is $\tfrac{n(n-1)}{2},$ so the number of games where there is at least one boy playing is $\tfrac{3n(3n-1)}{2} - \tfrac{n(n-1)}{2}.$ This means we can write and solve the following inequality. \begin{align*} \frac{3n(3n-1)}{2} - \frac{n(n-1)}{2} &\ge \frac{7n(3n-1)}{8} \\ \frac{5n(3n-1)}{8} &\ge \frac{n(n-1)}{2} \end{align*} Since $n > 0,$ we do not have to change the inequality sign when we divided by $n.$ \begin{align*} \frac{5(3n-1)}{4} &\ge n-1 \\ \frac{15n}{4} - \frac54 &\ge n-1 \\ \frac{11n}{4} &\ge \frac14 \\ n &\ge \frac{1}{11} \end{align*} Thus, we can confirm that all positive integers congruent to 0 or 3 modulo 8 satisfy the conditions. In summary, $\boxed{n \in \{\mathbf{N} \equiv 0,3 \pmod{8}\}}.$
\[ n \in \{ \mathbf{N} \equiv 0, 3 \pmod{8} \} \]
jbmo
[ "Mathematics -> Geometry -> Plane Geometry -> Triangulations", "Mathematics -> Algebra -> Intermediate Algebra -> Other" ]
6
Let $ABC$ be a triangle with $\angle A = 90^{\circ}$ . Points $D$ and $E$ lie on sides $AC$ and $AB$ , respectively, such that $\angle ABD = \angle DBC$ and $\angle ACE = \angle ECB$ . Segments $BD$ and $CE$ meet at $I$ . Determine whether or not it is possible for segments $AB, AC, BI, ID, CI, IE$ to all have integer lengths.
We know that angle $BIC = 135^{\circ}$ , as the other two angles in triangle $BIC$ add to $45^{\circ}$ . Assume that only $AB, AC, BI$ , and $CI$ are integers. Using the Law of Cosines on triangle BIC, $BC^2 = BI^2 + CI^2 - 2BI\cdot CI \cdot \cos 135^{\circ}$ . Observing that $BC^2 = AB^2 + AC^2$ is an integer and that $\cos 135^{\circ} = -\frac{\sqrt{2}}{2},$ we have and therefore, The LHS ( $\sqrt{2}$ ) is irrational, while the RHS is the quotient of the division of two integers and thus is rational. Clearly, there is a contradiction. Therefore, it is impossible for $AB, AC, BI$ , and $CI$ to all be integers, which invalidates the original claim that all six lengths are integers, and we are done.
It is impossible for \( AB, AC, BI, ID, CI, IE \) to all have integer lengths.
usamo
[ "Mathematics -> Number Theory -> Congruences", "Mathematics -> Algebra -> Algebra -> Equations and Inequalities" ]
7
Determine all integral solutions of $a^2+b^2+c^2=a^2b^2$ .
Either $a^2=0$ or $a^2>0$ . If $a^2=0$ , then $b^2=c^2=0$ . Symmetry applies for $b$ as well. If $a^2,b^2\neq 0$ , then $c^2\neq 0$ . Now we look at $a^2\bmod{4}$ : $a^2\equiv 0\bmod{4}$ : Since a square is either 1 or 0 mod 4, then all the other squares are 0 mod 4. Let $a=2a_1$ , $b=2b_1$ , and $c=2c_1$ . Thus $a_1^2+b_1^2+c_1^2=4a_1^2b_1^2$ . Since the LHS is divisible by four, all the variables are divisible by 4, and we must do this over and over again, and from infinite descent, there are no non-zero solutions when $a^2\equiv 0\bmod{4}$ . $a^2\equiv 1\bmod{4}$ : Since $b^2\neq 0\bmod{4}$ , $b^2\equiv 1\bmod{4}$ , and $2+c^2\equiv 1\bmod{4}$ . But for this to be true, $c^2\equiv 3\bmod{4}$ , which is an impossibility. Thus there are no non-zero solutions when $a^2\equiv 1\bmod{4}$ . Thus the only solution is the solution above: $(a,b,c)=0$ . Alternate solutions are always welcome. If you have a different, elegant solution to this problem, please add it to this page.
The only integral solution is \((a, b, c) = (0, 0, 0)\).
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Polygons", "Mathematics -> Discrete Mathematics -> Combinatorics" ]
7
Find the smallest positive integer $n$ such that if $n$ squares of a $1000 \times 1000$ chessboard are colored, then there will exist three colored squares whose centers form a right triangle with sides parallel to the edges of the board.
We claim that $n = 1999$ is the smallest such number. For $n \le 1998$ , we can simply color any of the $1998$ squares forming the top row and the left column, but excluding the top left corner square. [asy] for(int i = 0; i < 10; ++i){ for(int j = 0; j < 10; ++j){ if((i == 0 || j == 9) && !(j-i == 9)) fill(shift(i,j)*unitsquare,rgb(0.3,0.3,0.3)); else draw(shift(i,j)*unitsquare); } } [/asy] We now show that no configuration with no colored right triangles exists for $n = 1999$ . We call a row or column filled if all $1000$ of its squares are colored. Then any of the remaining $999$ colored squares must share a column or row, respectively, with one of the colored squares in a filled row or column. These two squares, and any other square in the filled row or column, form a colored right triangle, giving us a contradiction. Hence, no filled row or column may exist. Let $m$ be the number of columns with $1$ colored square. Then there are $1999-m$ colored squares in the remaining columns, and in each of these $< 1999-m$ columns that have a colored square must have at least two colored squares in them. These two colored squares will form a triangle with any other colored square in either of the rows containing the colored squares. Hence, each of the $1999-m$ colored squares must be placed in different rows, but as there are only $1000$ rows, the inequality $1999 - m \le 1000 \Longrightarrow m \ge 999$ holds. If $m = 1000$ , then each column only has $1$ colored square, leaving no place for the remaining $999$ , contradiction. If $m = 999$ , then each of the $1000$ rows has $1$ black square, leaving no place for the other $999$ , contradiction. Hence $n = \boxed{1999}$ is the minimal value.
\boxed{1999}
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Angles" ]
7
Let $A B C$ be an equilateral triangle. Let $P$ be a point on the side $A C$ and $Q$ be a point on the side $A B$ so that both triangles $A B P$ and $A C Q$ are acute. Let $R$ be the orthocentre of triangle $A B P$ and $S$ be the orthocentre of triangle $A C Q$. Let $T$ be the point common to the segments $B P$ and $C Q$. Find all possible values of $\angle C B P$ and $\angle B C Q$ such that triangle $T R S$ is equilateral.
We are going to show that this can only happen when $\angle C B P=\angle B C Q=15^{\circ}$. Lemma. If $\angle C B P>\angle B C Q$, then $R T>S T$. Proof. Let $A D, B E$ and $C F$ be the altitudes of triangle $A B C$ concurrent at its centre $G$. Then $P$ lies on $C E, Q$ lies on $B F$, and thus $T$ lies in triangle $B D G$. Note that $\angle F A S=\angle F C Q=30^{\circ}-\angle B C Q>30^{\circ}-\angle C B P=\angle E B P=\angle E A R$. Since $A F=A E$, we have $F S>E R$ so that $G S=G F-F S<G E-E R=G R$. Let $T_{x}$ be the projection of $T$ onto $B C$ and $T_{y}$ be the projection of $T$ onto $A D$, and similarly for $R$ and $S$. We have $R_{x} T_{x}=D R_{x}+D T_{x}>\left|D S_{x}-D T_{x}\right|=S_{x} T_{x}$ and $R_{y} T_{y}=G R_{y}+G T_{y}>G S_{y}+G T_{y}=S_{y} T_{y}$. It follows that $R T>S T$. Thus, if $\triangle T R S$ is equilateral, we must have $\angle C B P=\angle B C Q$. It is clear from the symmetry of the figure that $T R=T S$, so $\triangle T R S$ is equilateral if and only if $\angle R T A=30^{\circ}$. Now, as $B R$ is an altitude of the triangle $A B C, \angle R B A=30^{\circ}$. So $\triangle T R S$ is equilateral if and only if $R T B A$ is a cyclic quadrilateral. Therefore, $\triangle T R S$ is equilateral if and only if $\angle T B R=\angle T A R$. But $90^{\circ}=\angle T B A+\angle B A R=(\angle T B R+30^{\circ})+(30^{\circ}+\angle T A R)$ and so $30^{\circ}=\angle T A R+\angle T B R$. But these angles must be equal, so $\angle T A R=\angle T B R=15^{\circ}$. Therefore $\angle C B P=\angle B C Q=15^{\circ}$.
\[ \angle C B P = \angle B C Q = 15^\circ \]
apmoapmo_sol
[ "Mathematics -> Number Theory -> Factorization", "Mathematics -> Discrete Mathematics -> Graph Theory" ]
6.5
( Zuming Feng ) Determine all composite positive integers $n$ for which it is possible to arrange all divisors of $n$ that are greater than 1 in a circle so that no two adjacent divisors are relatively prime.
Solution 1 (official solution) No such circular arrangement exists for $n=pq$ , where $p$ and $q$ are distinct primes. In that case, the numbers to be arranged are $p$ ; $q$ and $pq$ , and in any circular arrangement, $p$ and $q$ will be adjacent. We claim that the desired circular arrangement exists in all other cases. If $n=p^e$ where $e\ge2$ , an arbitrary circular arrangement works. Henceforth we assume that $n$ has prime factorization $p^{e_1}_{1}p^{e_2}_{2}\cdots p^{e_k}_k$ , where $p_1<p_2<\cdots<p_k$ and either $k>2$ or else $\max(e1,e2)>1$ . To construct the desired circular arrangement of $D_n:=\lbrace d:d|n\ \text{and}\ d>1\rbrace$ , start with the circular arrangement of $n,p_{1}p_{2},p_{2}p_{3}\ldots,p_{k-1}p_{k}$ as shown. Then between $n$ and $p_{1}p_{2}$ , place (in arbitrary order) all other members of $D_n$ that have $p_1$ as their smallest prime factor. Between $p_{1}p_{2}$ and $p_{2}p_{3}$ , place all members of $D_n$ other than $p_{2}p_{3}$ that have $p_2$ as their smallest prime factor. Continue in this way, ending by placing $p_k,p^{2}_{k},\ldots,p^{e_k}_{k}$ between $p_{k-1}p_k$ and $n$ . It is easy to see that each element of $D_n$ is placed exactly one time, and any two adjacent elements have a common prime factor. Hence this arrangement has the desired property. Note. In graph theory terms, this construction yields a Hamiltonian cycle in the graph with vertex set $D_n$ in which two vertices form an edge if the two corresponding numbers have a common prime factor. The graphs below illustrate the construction for the special cases $n=p^{2}q$ and $n=pqr$ . Solution 2 The proof that no arrangement exists for $n=pq$ , where $p,q$ are distinct primes follows from above. Apply induction to prove all other cases are possible Base case: , where is a prime and is a positive integer. Any arrangement suffices , where are distinct primes. The following configuration works \[p,pq,pr,r,qr,q,pqr\] Inductive step: Suppose the desired arrangement exists for a composite $n$ , show the arrangement exists for $np^r$ , where $p$ is a prime relatively prime to $n$ and $r$ is a positive integer Let $a_1,a_2,\cdots,a_m$ be the arrangement of divisors of $n$ , then $(a_i,a_{i+1})>1$ for $i=1,2,\cdots,m$ , where $a_{m+1}=a_1$ . The divisors of $np^r$ greater than 1 are of the form \[a_ip^j,p^j\qquad 1\leq i\leq m,1\leq j\leq r\] The following sequence works \[a_1,\cdots,a_{m-1}, a_{m-1}p,\text{other divisors in arbitrary order},a_mp,a_m\] since all other divisors are divisible by $p$ . Alternate solutions are always welcome. If you have a different, elegant solution to this problem, please add it to this page.
The composite positive integers \( n \) for which it is possible to arrange all divisors of \( n \) that are greater than 1 in a circle so that no two adjacent divisors are relatively prime are those \( n \) that are not of the form \( pq \) where \( p \) and \( q \) are distinct primes.
usamo
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
7
Each of eight boxes contains six balls. Each ball has been colored with one of $n$ colors, such that no two balls in the same box are the same color, and no two colors occur together in more than one box. Determine, with justification, the smallest integer $n$ for which this is possible.
We claim that $n=23$ is the minimum. Consider the following construction (replacing colors with numbers) which fulfills this: \[\left[ \begin{array}{cccccccc} 1 & 1 & 1 & 2 & 3 & 4 & 5 & 6 \\ 2 & 7 & 12 & 7 & 8 & 9 & 10 & 11 \\ 3 & 8 & 13 & 12 & 13 & 14 & 15 & 16 \\ 4 & 9 & 14 & 17 & 17 & 17 & 18 & 19 \\ 5 & 10 & 15 & 18 & 20 & 22 & 20 & 21 \\ 6 & 11 & 16 & 19 & 21 & 23 & 22 & 23 \end{array} \right]\] Suppose a configuration exists with $n \le 22$ . Suppose a ball appears $5$ or more times. Then the remaining balls of the $5$ boxes must be distinct, so that there are at least $n \ge 5 \cdot 5 + 1 = 26$ balls, contradiction. If a ball appears $4$ or more times, the remaining balls of the $4$ boxes must be distinct, leading to $5 \cdot 4 + 1 = 21$ balls. The fifth box can contain at most four balls from the previous boxes, and then the remaining two balls must be distinct, leading to $n \ge 2 + 21 = 23$ , contradiction. However, by the Pigeonhole Principle , at least one ball must appear $3$ times. Without loss of generality suppose that $1$ appears three times, and let the boxes that contain these have balls with colors $\{1,2,3,4,5,6\},\{1,7,8,9,10,11\},\{1,12,13,14,15,16\}$ . Each of the remaining five boxes can have at most $3$ balls from each of these boxes. Thus, each of the remaining five boxes must have $3$ additional balls $> 16$ . Thus, it is necessary that we use $\le 22 - 16 = 6$ balls to fill a $3 \times 5$ grid by the same rules. Again, no balls may appear $\ge 4$ times, but by Pigeonhole, one ball must appear $3$ times. Without loss of generality , let this ball have color $17$ ; then the three boxes containing $17$ must have at least $2 \cdot 3 + 1 = 7$ balls, contradiction. Therefore, $n = 23$ is the minimum.
The smallest integer \( n \) for which this is possible is \( 23 \).
usamo
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
8
Let $S = \{(x,y) | x = 1, 2, \ldots, 1993, y = 1, 2, 3, 4\}$. If $T \subset S$ and there aren't any squares in $T.$ Find the maximum possible value of $|T|.$ The squares in T use points in S as vertices.
Let \( S = \{(x,y) \mid x = 1, 2, \ldots, 1993, y = 1, 2, 3, 4\} \). We aim to find the maximum possible value of \( |T| \) for a subset \( T \subset S \) such that there are no squares in \( T \). To solve this, we need to ensure that no four points in \( T \) form the vertices of a square. The key observation is that for any square in \( S \), we can have at most 3 of its vertices in \( T \). This gives a weak upper bound: \[ |T| \leq \frac{3}{4} |S|. \] We will use a more refined approach to maximize \( |T| \). Consider the columns of \( S \). If a column \( C \) contains all its elements in \( T \), then the adjacent columns can have at most 2 elements in \( T \) to avoid forming squares. Thus, it is more efficient to avoid having all elements of any column in \( T \). We can choose 3 elements from each column in \( T \) without forming squares. To achieve this, we can use a pattern where each set of 4 adjacent columns has distinct permutations of 3 elements in \( T \) and 1 element not in \( T \). This pattern avoids forming squares and maximizes the number of elements in \( T \). For example, consider the following arrangement for 4 columns: \[ \begin{array}{cccc} \bullet & \circ & \circ & \circ \\ \circ & \circ & \bullet & \circ \\ \circ & \bullet & \circ & \circ \\ \circ & \circ & \circ & \bullet \\ \end{array} \] Here, \( \bullet \) represents an element in \( T \) and \( \circ \) represents an element not in \( T \). This pattern can be repeated, with a separating column containing only 1 element in \( T \) to avoid forming squares. Given that there are 1993 columns, we can divide them into groups of 5 columns (4 columns with 3 elements each and 1 separating column with 1 element). Thus, we have: \[ 1993 = 5 \cdot 398 + 3. \] The maximum number of elements in \( T \) is: \[ 398 \cdot 13 + 3 \cdot 3 = 5183. \] Therefore, the maximum possible value of \( |T| \) is: \[ \boxed{5183}. \]
5183
china_team_selection_test
[ "Mathematics -> Algebra -> Sequences and Series -> Other" ]
5.5
Let $n > 1$ be an integer. Find, with proof, all sequences $x_1, x_2, \ldots, x_{n-1}$ of positive integers with the following three properties: (a). ; (b). for all ; (c). given any two indices and (not necessarily distinct) for which , there is an index such that .
The sequence is $2, 4, 6, \ldots, 2n-2$ . Proof 1 We will prove that any sequence $x_1, \ldots, x_{n-1}$ , that satisfies the given conditions, is an arithmetic progression with $x_1$ as both the first term and the increment. Once this is proved, condition (b) implies that $x_1 + x_{n-1} = x_1 + (n-1)x_1 = nx_1 = 2n$ . Therefore $x_1 = 2$ , and the sequence is just the even numbers from $2$ to $2n-2$ . The sequence of successive even numbers clearly satisfies all three conditions, and we are done. First a degenerate case. If $n = 2$ , there is only one element $x_1$ , and condition (b) gives $x_1 + x_1 = 4$ or $x_1 = 2$ . Conditions (a) and (c) are vacuously true. Otherwise, for $n > 2$ , we will prove by induction on $m$ that the difference $x_{n-m} - x_{n-1-m} = x_1$ for all $m \in [1, n-2]$ , which makes all the differences $x_{n-1} - x_{n-2} = \ldots = x_2 - x_1 = x_1$ , i.e. the sequence is an arithmetic progression with $x_1$ as the first term and increment as promised. So first the $m=1$ case. With $n > 2$ , $x_{n-2}$ exists and is less than $x_{n-1}$ by condition (a). Now since by condition (b) $x_1 + x_{n-1} = 2n$ , we conclude that $x_1 + x_{n-2} < 2n$ , and therefore by condition (c) $x_1 + x_{n-2} = x_k$ for some $k$ . Now, since $x_1 > 0$ , $x_k > x_{n-2}$ and can only be $x_{n-1}$ . So $x_1 + x_{n-2} = x_{n-1}$ . Now for the induction step on all values of $m$ . Suppose we have shown that for all $i \le m$ , $x_1 + x_{n-1-i} = x_{n-i}$ . If $m = n-2$ we are done, otherwise $m < n-2$ , and by condition (c) $x_1 + x_{n-2-m} = x_k$ for some $k$ . This $x_k$ is larger than $x_{n-2-m}$ , but smaller than $x_1 + x_{n-1-m} = x_{n-m}$ by the inductive hypothesis. It then follows that $x_1 + x_{n-2-m} = x_{n-1-m}$ , the only element of the sequence between $x_{n-2-m}$ and $x_{n-m}$ . This establishes the result for $i=m+1$ . So, by induction $x_1 + x_{n-1-m} = x_{n-m}$ for all $m \in [1, n-2]$ , which completes the proof. Proof 2 Let $S=\{x_1,x_2,...,x_{n-1}\}$ . Notice that \[x_1<x_1+x_1<x_1+x_2<\dots <x_1+x_{n-2}<2n.\] Then by condition (c), we must have $x_1,x_1+x_1,...,x_1+x_{n-2}\in S$ . This implies that $x_1=x_1,x_1+x_1=x_2,...,x_1+x_{n-2}=x_{n-1}$ , or that $x_k=kx_1$ . Then we have $x_1+x_{n-1}=n(x_1)=2n\rightarrow x_1=2$ , and the rest is trivial.
The sequence is \(2, 4, 6, \ldots, 2n-2\).
usajmo
[ "Mathematics -> Algebra -> Algebra -> Equations and Inequalities", "Mathematics -> Number Theory -> Congruences" ]
8
Suppose that $(a_1, b_1), (a_2, b_2), \ldots , (a_{100}, b_{100})$ are distinct ordered pairs of nonnegative integers. Let $N$ denote the number of pairs of integers $(i, j)$ satisfying $1 \le i < j \le 100$ and $|a_ib_j - a_j b_i|=1$ . Determine the largest possible value of $N$ over all possible choices of the $100$ ordered pairs.
Let's start off with just $(a_1, b_1), (a_2, b_2)$ and suppose that it satisfies the given condition. We could use $(1, 1), (1, 2)$ for example. We should maximize the number of conditions that the third pair satisfies. We find out that the third pair should equal $(a_1+a_2, b_1+b_2)$ : We know this must be true: \[|a_1b_2-a_2b_1| = 1\] So \[a_1b_2-a_2b_1 = 1\] We require the maximum conditions for $(a_3, b_3)$ \[|a_3b_2-a_2b_3| = 1\] \[|a_3b_1-a_1b_3| = 1\] Then one case can be: \[a_3b_2-a_2b_3 = 1\] \[a_3b_1-a_1b_3 = -1\] We try to do some stuff such as solving for $a_3$ with manipulations: \[a_3b_2a_1-a_2b_3a_1 = a_1\] \[a_3b_1a_2-a_1b_3a_2 = -a_2\] \[a_3(a_1b_2-a_2b_1) = a_1+a_2\] \[a_3 = a_1+a_2\] \[a_3b_2b_1-a_2b_3b_1 = b_1\] \[a_3b_1b_2-a_1b_3b_2 = -b_2\] \[b_3(a_1b_2-a_2b_1) = b_1+b_2\] \[b_3 = b_1+b_2\] We showed that 3 pairs are a complete graph; however, 4 pairs are not a complete graph. We will now show that: \[a_4 = a_1+2a_2\] \[b_4 = b_1+2b_2\] \[|a_1b_1+2a_2b_1-a_1b_1-2a_1b_2| = 1\] \[2|a_2b_1-a_1b_2| = 1\] This is clearly impossible because $1$ is not even and also $|a_2b_1-a_1b_2| = 1$ . The answer is as follows: \[0+1+2+\ldots+2\] $a_1$ has $0$ subtractions that follow condition while $a_2$ has $1$ and then the rest has $2$ . There are $n$ terms, so our answer be $2n-3$ and in case of $n=100$ that means \[\boxed{N=197}.\] ~Lopkiloinm
\[\boxed{N=197}\]
usamo
[ "Mathematics -> Precalculus -> Functions" ]
7.5
Let $\mathbb{Z}$ be the set of integers. Find all functions $f : \mathbb{Z} \rightarrow \mathbb{Z}$ such that \[xf(2f(y)-x)+y^2f(2x-f(y))=\frac{f(x)^2}{x}+f(yf(y))\] for all $x, y \in \mathbb{Z}$ with $x \neq 0$ .
Note: This solution is kind of rough. I didn't want to put my 7-page solution all over again. It would be nice if someone could edit in the details of the expansions. Lemma 1: $f(0) = 0$ . Proof: Assume the opposite for a contradiction. Plug in $x = 2f(0)$ (because we assumed that $f(0) \neq 0$ ), $y = 0$ . What you get eventually reduces to: \[4f(0)-2 = \left( \frac{f(2f(0))}{f(0)} \right)^2\] which is a contradiction since the LHS is divisible by 2 but not 4. Then plug in $y = 0$ into the original equation and simplify by Lemma 1. We get: \[x^2f(-x) = f(x)^2\] Then: \begin{align*} x^6f(x) &= x^4\bigl(x^2f(x)\bigr)\\ &= x^4\bigl((-x)^2f(-(-x))\bigr)\\ &= x^4(-x)^2f(-(-x))\\ &= x^4f(-x)^2\\ &= f(x)^4 \end{align*} Therefore, $f(x)$ must be 0 or $x^2$ . Now either $f(x)$ is $x^2$ for all $x$ or there exists $a \neq 0$ such that $f(a)=0$ . The first case gives a valid solution. In the second case, we let $y = a$ in the original equation and simplify to get: \[xf(-x) + a^2f(2x) = \frac{f(x)^2}{x}\] But we know that $xf(-x) = \frac{f(x)^2}{x}$ , so: \[a^2f(2x) = 0\] Since $a$ is not 0, $f(2x)$ is 0 for all $x$ (including 0). Now either $f(x)$ is 0 for all $x$ , or there exists some $m \neq 0$ such that $f(m) = m^2$ . Then $m$ must be odd. We can let $x = 2k$ in the original equation, and since $f(2x)$ is 0 for all $x$ , stuff cancels and we get: \[y^2f(4k - f(y)) = f(yf(y))\] for . Now, let $y = m$ and we get: \[m^2f(4k - m^2) = f(m^3)\] Now, either both sides are 0 or both are equal to $m^6$ . If both are $m^6$ then: \[m^2(4k - m^2)^2 = m^6\] which simplifies to: \[4k - m^2 = \pm m^2\] Since $k \neq 0$ and $m$ is odd, both cases are impossible, so we must have: \[m^2f(4k - m^2) = f(m^3) = 0\] Then we can let $k$ be anything except 0, and get $f(x)$ is 0 for all $x \equiv 3 \pmod{4}$ except $-m^2$ . Also since $x^2f(-x) = f(x)^2$ , we have $f(x) = 0 \Rightarrow f(-x) = 0$ , so $f(x)$ is 0 for all $x \equiv 1 \pmod{4}$ except $m^2$ . So $f(x)$ is 0 for all $x$ except $\pm m^2$ . Since $f(m) \neq 0$ , $m = \pm m^2$ . Squaring, $m^2 = m^4$ and dividing by $m$ , $m = m^3$ . Since $f(m^3) = 0$ , $f(m) = 0$ , which is a contradiction for $m \neq 1$ . However, if we plug in $x = 1$ with $f(1) = 1$ and $y$ as an arbitrary large number with $f(y) = 0$ into the original equation, we get $0 = 1$ which is a clear contradiction, so our only solutions are $f(x) = 0$ and $f(x) = x^2$ .
The functions that satisfy the given equation are: \[ f(x) = 0 \] and \[ f(x) = x^2 \]
usamo
[ "Mathematics -> Algebra -> Intermediate Algebra -> Other", "Mathematics -> Discrete Mathematics -> Other" ]
7
Prove by induction on \(n\) that \(\frac{(m+n)!}{(m-n)!}=\prod_{i=1}^{n}\left(m^{2}+m-i^{2}+i\right)\).
1. \(\frac{(m+1)!}{(m-1)!}=m(m+1)=m^{2}+m\). 2. \(\frac{(m+n+1)!}{(m-n-1)!}=\left(\prod_{i=1}^{n}\left(m^{2}+m-i^{2}+i\right)\right)(m+n+1)(m-n)\) (by induction) \(=\left(\prod_{i=1}^{n}\left(m^{2}+m-i^{2}+i\right)\right)\left(m^{2}+m-n^{2}-n\right)=\prod_{i=1}^{n+1}\left(m^{2}+m-i^{2}+i\right)\). But \(m^{2}+m \geq m^{2}+m-i^{2}+i \geq i^{2}+i-i^{2}+i=2 i\), for \(i \geq m\). Therefore \(2^{n} n!\leq \frac{(m+n)!}{(m-n)!} \leq\left(m^{2}+m\right)^{n}\).
\[\frac{(m+n)!}{(m-n)!} = \prod_{i=1}^{n}\left(m^{2}+m-i^{2}+i\right)\]
apmoapmo_sol
[ "Mathematics -> Algebra -> Algebra -> Equations and Inequalities", "Mathematics -> Number Theory -> Other" ]
5
Determine all the sets of six consecutive positive integers such that the product of some two of them, added to the product of some other two of them is equal to the product of the remaining two numbers.
$x_1 x_2 + x_3 x_4 = x_5 x_6$ Every set which is a solution must be of the form $Y_k = \{k, k+1, k+2, k+3, k+4, k+5\}$ Since they are consecutive, it follows that $x_2, x_4, x_6$ are even and $x_1, x_3, x_5$ are odd. In addition, exactly two of the six integers are multiples of $3$ and need to be multiplied together. Exactly one of these two integers is even (and also the only one which is multiple of $6$ ) and the other one is odd. Also, each pair of positive integers destined to be multiplied together can have a difference of either $1$ or $3$ or $5$ . So, we only have to consider integers from $1$ up to $11$ since $k \leq 6$ (see inequalities below). Therefore we calculate the following products: A = { 2⋅1, 4⋅5, 8⋅7, 13⋅14, 10⋅11} B = { 1⋅4, 2⋅5, 3⋅6, 4⋅7, 5⋅8, 6⋅9, 7⋅10, 8⋅11} C = { 2⋅7, 5⋅10} In any case, either 3⋅6 or 6⋅9 needs to be included in every solution set. {7⋅10, 8⋅11} cannot be part of any potential solution set, but they were included here just for completeness. Method 1: Using a graph One could also construct a graph G=(V,E) with the set V of vertices (also called nodes or points) and the set E of edges (also called arcs or line). The elements of all sets A,B,C will be the vertices. The edges will be the possibles combinations, so the candidate solutions will form a cycle of exactly three vertices. So, there should be eight such cycles. Only three of them will be the valid solution sets: $2 \cdot 5 + 3 \cdot 6 = 4 \cdot 7$ $1 \cdot 2 + 3 \cdot 6 = 4 \cdot 5$ $7 \cdot 8 + 6 \cdot 9 = 10 \cdot 11$ Rejected: { 1⋅4, 2⋅5, 3⋅6 }, {3⋅6, 4⋅7 5⋅8}, { 3⋅6, 4⋅5, 8⋅7 }, {6⋅9, 4⋅5, 8⋅7}, {3⋅6, 2⋅7, 4⋅5} Method 2: Taking cases Alternatively, we can have five cases at most (actually only three) : Case 1: $|x_1-x_2| = |x_3 -x_4| = |x_5 -x_6| = 3$ $k(k+3)+(k+1)(k+4) \leq (k+2)(k+5) \to k \leq 3$ We just have to look at set B in this case. $2 \cdot 5 + 3 \cdot 6 = 4 \cdot 7$ Rejected: { 1⋅4, 2⋅5, 3⋅6 }, {3⋅6, 4⋅7 5⋅8} $Y_2$ is the only solution set for this case. Case 2: $|x_1-x_2| =1, |x_3 -x_4| = 3, |x_5 -x_6| = 1$ $k(k+1)+(k+2)(k+5) \leq (k+3)(k+4) \to k \leq 3$ $k(k+3)+(k+1)(k+2) \leq (k+4)(k+5) \to k \leq 6$ $1 \cdot 2 + 3 \cdot 6 = 4 \cdot 5$ $7 \cdot 8 + 6 \cdot 9 = 10 \cdot 11$ Rejected: { 3⋅6, 4⋅5, 8⋅7 }, {6⋅9, 4⋅5, 8⋅7} $Y_1$ and $Y_6$ are the only solution sets for this case. Case 3: $|x_1-x_2| = 3 |x_3 -x_4| = 5, |x_5 -x_6| = 1$ $k(k+5)+(k+1)(k+4) \leq (k+3)(k+5) \to k \leq 2$ Rejected: {3⋅6, 2⋅7, 4⋅5} No solution set for this case since they were all rejected. Case 4: $|x_1-x_2| = |x_3 -x_4| = |x_5 -x_6| = 1$ No solution set for this case, as the multiples of three need to be multiplied together. (This case is actually not realistic and it was included here just for completeness.) Case 5: $|x_1-x_2| = 1, |x_3 -x_4| = 5, |x_5 -x_6| = 1$ No solution set for this case, as the multiples of three need to be multiplied together. (This case is actually not realistic and it was included here just for completeness.) Solution 2 Let the six numbers be $a, a+1, a+2, a+3, a+4, a+5$ . We can bound the graph to restrict the values of $a$ by setting the inequality $(a+5)(a+4)\geq a(a+3)+(a+2)(a+1)$ . If we sum the pairwise products, we show that the minimum sum obtainable is $a(a+3)+(a+2)(a+1)$ by first establishing that the $a^2$ and $a$ terms will be the same no matter how you pair them. However, we can manipulate the products such that we obtain the least possible constant, which is achievable by treating $a=a+0$ so we remove the greatest constant factor which is 3. Solving the inequality and making the left side 0, we get $0\geq a^2-6a-18=(a+3)(a-6)$ . If the quadratic is greater than 0, it means that the RHS will be too large for any $a$ to suffice because we have already chosen the minimum value for the RHS and the maximum for the LHS. Clearly, $a=6$ works as a set. Now, we find the second largest product for the LHS and the smallest pairwise product-sum for the RHS. This is achievable at $(a+5)(a+3)\geq a(a+4)+(a+2)(a+1) \Longrightarrow 0\geq a^2-a-13$ . Solving for the quadratic, we see that there are no integer roots for $a$ but we can bound $a\leq 4$ . Now, we take the third largest product for the LHS which is $(a+4)(a+3)\geq a(a+5)+(a+2)(a+1) \Longrightarrow 0\geq a^2+a-10$ bounding $a\leq 2$ . Now, we can further bound(simply use the method above) to see no solutions here.
The sets of six consecutive positive integers that satisfy the given condition are: \[ \{2, 3, 4, 5, 6, 7
jbmo
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
8
Consider a \( 2018 \times 2019 \) board with integers in each unit square. Two unit squares are said to be neighbours if they share a common edge. In each turn, you choose some unit squares. Then for each chosen unit square the average of all its neighbours is calculated. Finally, after these calculations are done, the number in each chosen unit square is replaced by the corresponding average. Is it always possible to make the numbers in all squares become the same after finitely many turns?
No. Let \( n \) be a positive integer relatively prime to 2 and 3. We may study the whole process modulo \( n \) by replacing divisions by \( 2,3,4 \) with multiplications by the corresponding inverses modulo \( n \). If at some point the original process makes all the numbers equal, then the process modulo \( n \) will also have all the numbers equal. Our aim is to choose \( n \) and an initial configuration modulo \( n \) for which no process modulo \( n \) reaches a board with all numbers equal modulo \( n \). We split this goal into two lemmas. Lemma 1. There is a \( 2 \times 3 \) board that stays constant modulo 5 and whose entries are not all equal. Proof. Here is one such a board: \( \begin{bmatrix} 1 & 1 & 1 \\ 1 & 2 & 2 \end{bmatrix} \). The fact that the board remains constant regardless of the choice of squares can be checked square by square. Lemma 2. If there is an \( r \times s \) board with \( r \geq 2, s \geq 2 \), that stays constant modulo 5, then there is also a \( kr \times ls \) board with the same property. Proof. We prove by a case by case analysis that repeatedly reflecting the \( r \times s \) with respect to an edge preserves the property: - If a cell had 4 neighbors, after reflections it still has the same neighbors. - If a cell with \( a \) had 3 neighbors \( b, c, d \), we have by hypothesis that \( a \equiv 3^{-1}(b+c+d) \equiv 2(b+c+d)(\bmod 5) \). A reflection may add \( a \) as a neighbor of the cell and now \( 4^{-1}(a+b+c+d) \equiv 4(a+b+c+d) \equiv 4a+2a \equiv a \quad(\bmod 5) \). - If a cell with \( a \) had 2 neighbors \( b, c \), we have by hypothesis that \( a \equiv 2^{-1}(b+c) \equiv 3(b+c) \) \((\bmod 5) \). If the reflections add one \( a \) as neighbor, now \( 3^{-1}(a+b+c) \equiv 2(3(b+c)+b+c) \equiv 8(b+c) \equiv 3(b+c) \equiv a \quad(\bmod 5) \). - If a cell with \( a \) had 2 neighbors \( b, c \), we have by hypothesis that \( a \equiv 2^{-1}(b+c)(\bmod 5) \). If the reflections add two \( a \)'s as neighbors, now \( 4^{-1}(2a+b+c) \equiv(2^{-1}a+2^{-1}a) \equiv a \quad(\bmod 5) \). In the three cases, any cell is still preserved modulo 5 after an operation. Hence we can fill in the \( kr \times ls \) board by \( k \times l \) copies by reflection. Since \( 2 \mid 2018 \) and \( 3 \mid 2019 \), we can get through reflections the following board: \( \begin{bmatrix} 1 & 1 & 1 & \ldots & 1 & 1 \\ 1 & 2 & 2 & \ldots & 2 & 2 \\ \vdots & \vdots & \vdots & \ddots & \vdots & \vdots \\ 1 & 1 & 1 & \ldots & 1 & 1 \\ 1 & 2 & 2 & \ldots & 2 & 2 \end{bmatrix} \). By the lemmas above, the board is invariant modulo 5, so the answer is no.
No.
apmoapmo_sol
[ "Mathematics -> Geometry -> Solid Geometry -> 3D Shapes" ]
8
We know that $2021=43 \times 47$. Is there a polyhedron whose surface can be formed by gluing together 43 equal non-planar 47-gons? Please justify your answer with a rigorous argument.
The answer is YES. All we need to do is to construct an example. Let's consider a standard torus $\mathbb{T}$, whose points can be represented by two parameters: $\mathbb{T}=\{\theta, \varphi: 0 \leq \theta, \varphi<2 \pi\}$. One can view the $z$-axis as the axis of symmetry of the torus: $((R+r \cos \varphi) \cos \theta,(R+r \cos \varphi) \sin \theta, r \sin \varphi)$. For $1 \leq k \leq 43$, we consider the following region on the torus $D_{k}=\left\{\theta, \varphi: \frac{2(k-1)}{43} \pi+3 \frac{\varphi}{86} \leq \theta \leq \frac{2 k}{43} \pi+3 \frac{\varphi}{86}\right\}$. Intuitively, what we do here is to divide the torus into 43 equal parts, then cut every part along the circle $\{\varphi=0\}$, keep one side of the cut while sliding the other side along the circle for certain angle. Now, we deform the circle $\{\varphi=0\}$ into a regular 43-gon whose vertices correspond to $\theta=\frac{2 k}{43} \pi$. Then $D_{k}$ has four "sides" of (two of which lie on $\{\varphi=0\}$ ), four "corners" (two of which are adjacent vertices of the 43-gon, while the other two are midpoints of two sides, we need then mark the vertex of the 43-gon between these two midpoints). We denote $C_{k, 0}=\left(\frac{2(k-1)}{43} \pi, 0\right), C_{k, 1}=\left(\frac{2 k}{43} \pi, 0\right)$, $D_{k, 0}=\left(\frac{2 k+1}{43} \pi, 2 \pi\right), D_{k, 1}=\left(\frac{2 k+3}{43} \pi, 2 \pi\right)$, $E_{k}=\left(\frac{2 k+2}{43} \pi, 2 \pi\right)$. Take another "side" of $\partial D_{k}$, mark 21 points, e.g. $A_{k, i}=\left(\frac{2(k-1)}{43} \pi+3 \frac{\varphi}{86} \pi, \frac{i}{11} \pi\right), i=1, \ldots, 21$. Then rotate around $z$-axis by $\frac{2}{43} \pi$ to get another 21 points, denote them by $B_{k, i}, i=1, \ldots, 21$. Now we join $C_{k, 0} C_{k, 1}, C_{k, 0} A_{k, 1}, C_{k, 1} B_{k, 1}, A_{k, i} A_{k, i+1}, B_{k, i} B_{k, i+1}, A_{k, i} B_{k, i}, A_{k, i} B_{k, i+1}(i=1, \ldots, 21)$ and $A_{k, 21} D_{k, 0}, B_{k, 21} D_{k, 1}, A_{k, 21} E_{k}, B_{k, 21} E_{k}, D_{k, 0} E_{k}, E_{k} D_{k, 1}$. We get a non-planar 47-gon. Thus we get 43 congruent (the construction above is independent of $k$ )non-planar 47gons, they can be glue together to form a polyhedron.
YES
alibaba_global_contest
[ "Mathematics -> Number Theory -> Congruences" ]
7
Let $m,n$ be positive integers. Find the minimum positive integer $N$ which satisfies the following condition. If there exists a set $S$ of integers that contains a complete residue system module $m$ such that $| S | = N$, then there exists a nonempty set $A \subseteq S$ so that $n\mid {\sum\limits_{x \in A} x }$.
Let \( m \) and \( n \) be positive integers. We aim to find the minimum positive integer \( N \) which satisfies the following condition: If there exists a set \( S \) of integers that contains a complete residue system modulo \( m \) such that \( |S| = N \), then there exists a nonempty set \( A \subseteq S \) so that \( n \mid \sum_{x \in A} x \). First, let \( d = \gcd(m, n) \), and write \( m = ad \) and \( n = bd \). The answer depends on the relationship between \( bd \) and \( \frac{ad(d+1)}{2} \). The minimum positive integer \( N \) is given by: \[ N = \begin{cases} 1 & \text{if } bd \leq \frac{ad(d+1)}{2}, \\ bd - \frac{ad(d-1)}{2} & \text{otherwise}. \end{cases} \] The answer is: \boxed{\begin{cases} 1 & \text{if } bd \leq \frac{ad(d+1)}{2}, \\ bd - \frac{ad(d-1)}{2} & \text{otherwise}. \end{cases}}.
\begin{cases} 1 & \text{if } bd \leq \frac{ad(d+1)}{2}, \\ bd - \frac{ad(d-1)}{2} & \text{otherwise}. \end{cases}
china_national_olympiad
[ "Mathematics -> Number Theory -> Factorization", "Mathematics -> Algebra -> Algebra -> Equations and Inequalities" ]
8
Find all pairs of positive integers $(x, y)$ such that $(xy+1)(xy+x+2)$ be a perfect square .
To find all pairs of positive integers \((x, y)\) such that \((xy+1)(xy+x+2)\) is a perfect square, we start by analyzing the given expression. First, consider the greatest common divisor (gcd) of the two terms: \[ \gcd(xy+1, xy+x+2). \] Using the properties of gcd, we have: \[ \gcd(xy+1, xy+x+2) = \gcd(xy+1, x+1). \] This simplifies further to: \[ \gcd(xy+1, x+1) = \gcd(y-1, x+1). \] Let \(x+1 = da\) and \(y-1 = db\) where \(\gcd(a, b) = 1\). Then we can express \(xy+1\) and \(xy+x+2\) as: \[ xy+1 = d \cdot u^2 \quad \text{and} \quad xy+x+2 = d \cdot v^2, \] for some relatively prime integers \(u\) and \(v\). Using the relationship \(a = v^2 - u^2\), we can rewrite the equation as: \[ u^2 = (d \cdot b + 1)(v^2 - u^2) - b, \] or equivalently: \[ (d \cdot b + 1)v^2 - (d \cdot b + 2)u^2 = b. \] This is a form of a Pell equation. To solve this, note that \(v > u\). Let \(v = \frac{X+Y}{2}\) and \(u = \frac{X-Y}{2}\) for positive integers \(X\) and \(Y\). Substituting these into the equation, we get: \[ X^2 - (4bd + 6)XY + Y^2 + 4b = 0. \] Using Vieta jumping, assume there is a solution \((X, Y)\) in positive integers with \(X \ge Y\). By symmetry, the pair \(\left( \frac{Y^2+4b}{X}, Y \right)\) is also a solution. Repeating this process, we eventually reach pairs \((X_1, Y)\) and \((X_2, Y)\) with \(X_1 > X_2 \ge Y\). This implies: \[ \begin{align*} X_1 + X_2 &= (4bd + 6)Y, \\ X_1 \cdot X_2 &= Y^2 + 4b. \end{align*} \] If \(\min(X_1, X_2) > Y\) and \(X_1 + X_2 = (4bd + 6)Y\), then: \[ X_1 \cdot X_2 \ge Y \cdot (4bd + 5)Y > Y^2 + 4b, \] which leads to a contradiction. Thus, there are no pairs of positive integers \((x, y)\) such that \((xy+1)(xy+x+2)\) is a perfect square. The answer is: \boxed{\text{No solutions}}.
\text{No solutions}
china_team_selection_test
[ "Mathematics -> Algebra -> Algebra -> Equations and Inequalities", "Mathematics -> Number Theory -> Congruences" ]
7.5
Problem Solve in integers the equation \[x^2+xy+y^2 = \left(\frac{x+y}{3}+1\right)^3.\] Solution We first notice that both sides must be integers, so $\frac{x+y}{3}$ must be an integer. We can therefore perform the substitution $x+y = 3t$ where $t$ is an integer. Then: $(3t)^2 - xy = (t+1)^3$ $9t^2 + x (x - 3t) = t^3 + 3t^2 + 3t + 1$ $4x^2 - 12xt + 9t^2 = 4t^3 - 15t^2 + 12t + 4$ $(2x - 3t)^2 = (t - 2)^2(4t + 1)$ $4t+1$ is therefore the square of an odd integer and can be replaced with $(2n+1)^2 = 4n^2 + 4n +1$ By substituting using $t = n^2 + n$ we get: $(2x - 3n^2 - 3n)^2 = [(n^2 + n - 2)(2n+1)]^2$ $2x - 3n^2 - 3n = \pm (2n^3 + 3n^2 -3n -2)$ $x = n^3 + 3n^2 - 1$ or $x = -n^3 + 3n + 1$ Using substitution we get the solutions: $(n^3 + 3n^2 - 1, -n^3 + 3n + 1) \cup (-n^3 + 3n + 1, n^3 + 3n^2 - 1)$
Let $n = \frac{x+y}{3}$ . Thus, $x+y = 3n$ . We have \[x^2+xy+y^2 = \left(\frac{x+y}{3}+1\right)^3 \implies (x+y)^2 - xy = \left(\frac{x+y}{3}+1\right)^3\] Substituting $n$ for $\frac{x+y}{3}$ , we have \[9n^2 - x(3n-x) = (n+1)^3\] Treating $x$ as a variable and $n$ as a constant, we have \[9n^2 - 3nx + x^2 = (n+1)^3,\] which turns into \[x^2 - 3nx + (9n^2 - (n+1)^3) = 0,\] a quadratic equation. By the quadratic formula, \[x = \frac{1}{2} \left(3n \pm \sqrt{9n^2 - 4(9n^2 - (n+1)^3)} \right)\] which simplifies to \[x = \frac{1}{2} \left(3n \pm \sqrt{4(n+1)^3 - 27n^2} \right)\] Since we want $x$ and $y$ to be integers, we need $4(n+1)^3 - 27n^2$ to be a perfect square. We can factor the aforementioned equation to be \[(n-2)^2 (4n+1) = k^2\] for an integer $k$ . Since $(n-2)^2$ is always a perfect square, for $(n-2)^2 (4n+1)$ to be a perfect square, $4n + 1$ has to be a perfect square as well. Since $4n + 1$ is odd, the square root of the aforementioned equation must be odd as well. Thus, we have $4n + 1 = a^2$ for some odd $a$ . Thus, \[n = \frac{a^2 - 1}{4},\] in which by difference of squares it is easy to see that all the possible values for $n$ are just $n = p(p-1)$ , where $p$ is a positive integer. Thus, \[x+y = 3n = 3p(p-1).\] Thus, the general form for \[x = \frac{1}{2} \left(3p(p-1) \pm \sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \right)\] for a positive integer $p$ . (This is an integer since $4(p(p-1)+1)^3 - 27(p(p-1))^2$ is an even perfect square (since $4(p(p-1)+1)$ is always even, as well as $27(p(p-1))^2$ being always even) as established, and $3p(p-1)$ is always even as well. Thus, the whole numerator is even, which makes the quantity of that divided by $2$ always an integer.) Since $y = 3n - x$ , the general form for $y$ is just \[y = 3p(p-1) - \frac{1}{2} \left(3p(p-1) \pm \sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \right)\] (This is an integer since $4(p(p-1)+1)^3 - 27(p(p-1))^2$ is an even perfect square (since $4(p(p-1)+1)$ is always even, as well as $27(p(p-1))^2$ being always even) as established, and $3p(p-1)$ is always even as well. Thus, the whole numerator is even, which makes the quantity of that divided by $2$ always an integer, which thus trivially makes \[3p(p-1) - \frac{1}{2} \left(3p(p-1) \pm \sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \right)\] an integer.) for a positive integer $p$ . Thus, our general in integers $(x, y)$ is \[(\frac{1}{2} \left(3p(p-1) \pm \sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \right), 3p(p-1) - \frac{1}{2} \left(3p(p-1) \pm \sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \right).\] $\boxed{}$ -fidgetboss_4000
\[ \left( \frac{1}{2} \left(3p(p-1) \pm \sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \right), 3p(p-1) - \frac{1}{2} \left(3p(p-1) \pm \sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \right) \right) \]
usamo
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations" ]
9
Find out all the integer pairs $(m,n)$ such that there exist two monic polynomials $P(x)$ and $Q(x)$ ,with $\deg{P}=m$ and $\deg{Q}=n$,satisfy that $$P(Q(t))\not=Q(P(t))$$ holds for any real number $t$.
To find all integer pairs \((m,n)\) such that there exist two monic polynomials \(P(x)\) and \(Q(x)\) with \(\deg{P}=m\) and \(\deg{Q}=n\) satisfying \(P(Q(t)) \neq Q(P(t))\) for any real number \(t\), we analyze the given conditions and cases. ### Analysis: 1. **Case \((m,n) = (1,1)\):** - If \(P(x) = x + a\) and \(Q(x) = x + b\), then \(P(Q(x)) = Q(P(x)) = x + a + b\). Thus, \(P(Q(t)) = Q(P(t))\) for any \(t\). 2. **Case \((m,n) = (2k,1)\) or \((1,2k)\):** - Let \(P(x) = x^m + a\) and \(Q(x) = x + b\). Then \(P(Q(x)) = (x + b)^m + a\) and \(Q(P(x)) = x^m + a + b\). Since \(m\) is even, \(P(Q(x)) - Q(P(x))\) will have an odd degree term, ensuring there exists a real \(t\) such that \(P(Q(t)) = Q(P(t))\). 3. **Case where \((m,n) \neq (1,1), (1,2k), (2k,1)\):** - **Subcase 1: At least one of \(m\) or \(n\) is even.** - Without loss of generality, assume \(n\) is even. Choose \(P(x) = x^m\) and \(Q(x) = x^n + 3\). Then, \[ P(Q(x)) - Q(P(x)) = (x^n + 3)^m - x^{mn} - 3 \] is a polynomial with nonnegative coefficients and only even degrees, with a positive constant term. Hence, \(P(Q(x)) - Q(P(x)) > 0\) for all \(x \in \mathbb{R}\). - **Subcase 2: Both \(m\) and \(n\) are odd.** - Assume \(m > 1\). Choose \(P(x) = x^m\) and \(Q(x) = x^n + 3\). Then, \[ P(Q(x)) - Q(P(x)) = (x^n + 3)^m - x^{mn} - 3 \] can be shown to be positive for all \(x \in \mathbb{R}\) using properties of sums of powers and polynomial inequalities. ### Conclusion: The integer pairs \((m,n)\) such that \(P(Q(t)) \neq Q(P(t))\) for any real number \(t\) are all pairs except \((1,1)\), \((1,2k)\), and \((2k,1)\). The answer is: \(\boxed{\text{All pairs except } (1,1), (1,2k), (2k,1)}\).
\text{All pairs except } (1,1), (1,2k), (2k,1)
china_team_selection_test
[ "Mathematics -> Number Theory -> Congruences" ]
6
For distinct positive integers $a$ , $b < 2012$ , define $f(a,b)$ to be the number of integers $k$ with $1 \le k < 2012$ such that the remainder when $ak$ divided by 2012 is greater than that of $bk$ divided by 2012. Let $S$ be the minimum value of $f(a,b)$ , where $a$ and $b$ range over all pairs of distinct positive integers less than 2012. Determine $S$ .
Solution 1 First we'll show that $S \geq 502$ , then we'll find an example $(a, b)$ that have $f(a, b)=502$ . Let $x_k$ be the remainder when $ak$ is divided by 2012, and let $y_k$ be defined similarly for $bk$ . First, we know that, if $x_k > y_k >0$ , then $x_{2012-k} \equiv a(2012-k) \equiv 2012-ak \equiv 2012-x_k \pmod {2012}$ and $y_{2012-k} \equiv 2012-y_k \pmod {2012}$ . This implies that, since $2012 - x_k \neq 0$ and $2012 -y_k \neq 0$ , $x_{2012-k} < y_{2012-k}$ . Similarly, if $0< x_k < y_k$ then $x_{2012-k} > y_{2012-k}$ , establishing a one-to-one correspondence between the number of $k$ such that $x_k < y_k$ . Thus, if $n$ is the number of $k$ such that $x_k \neq y_k$ and $y_k \neq 0 \neq x_k$ , then $S \geq \frac{1}{2}n$ . Now I'll show that $n \geq 1004$ . If $gcd(k, 2012)=1$ , then I'll show you that $x_k \neq y_k$ . This is actually pretty clear; assume that's not true and set up a congruence relation: \[ak \equiv bk \pmod {2012}\] Since $k$ is relatively prime to 2012, it is invertible mod 2012, so we must have $a \equiv b \pmod {2012}$ . Since $0 < a, b <2012$ , this means $a=b$ , which the problem doesn't allow, thus contradiction, and $x_k \neq y_k$ . Additionally, if $gcd(k, 2012)=1$ , then $x_k \neq 0 \neq y_k$ , then based on what we know about $n$ from the previous paragraph, $n$ is at least as large as the number of k relatively prime to 2012. Thus, $n \geq \phi(2012) = \phi(503*4) = 1004$ . Thus, $S \geq 502$ . To show 502 works, consider $(a, b)=(1006, 2)$ . For all even $k$ we have $x_k=0$ , so it doesn't count towards $f(1006, 2)$ . Additionally, if $k = 503, 503*3$ then $x_k = y_k = 1006$ , so the only number that count towards $f(1006, 2)$ are the odd numbers not divisible by 503. There are 1004 such numbers. However, for all such odd k not divisible by 503 (so numbers relatively prime to 2012), we have $x_k \neq 0 \neq y_k$ and $2012-k$ is also relatively prime to 2012. Since under those conditions exactly one of $x_k > y_k$ and $x_{2012-k} > y_{2012-k}$ is true, we have at most 1/2 of the 1004 possible k actually count to $f(1006, 2)$ , so $\frac{1004}{2} = 502 \geq f(1006, 2) \geq S \geq 502$ , so $S=502$ . Solution 2 Let $ak \equiv r_{a} \pmod{2012}$ and $bk \equiv r_{b} \pmod{2012}$ . Notice that this means $a(2012 - k) \equiv 2012 - r_{a} \pmod{2012}$ and $b(2012 - k) \equiv 2012 - r_{b} \pmod{2012}$ . Thus, for every value of $k$ where $r_{a} > r_{b}$ , there is a value of $k$ where $r_{b} > r_{a}$ . Therefore, we merely have to calculate $\frac{1}{2}$ times the number of values of $k$ for which $r_{a} \neq r_{b}$ and $r_{a} \neq 0$ . However, the answer is NOT $\frac{1}{2}(2012) = 1006$ ! This is because we must count the cases where the value of $k$ makes $r_{a} = r_{b}$ or where $r_{a} = 0$ . So, let's start counting. If $k$ is even, we have either $a \equiv 0 \pmod{1006}$ or $a - b \equiv 0 \pmod{1006}$ . So, $a = 1006$ or $a = b + 1006$ . We have $1005$ even values of $k$ (which is all the possible even values of $k$ , since the two above requirements don't put any bounds on $k$ at all). If $k$ is odd, if $k = 503$ or $k = 503 \cdot 3$ , then $a \equiv 0 \pmod{4}$ or $a \equiv b \pmod{4}$ . Otherwise, $ak \equiv 0 \pmod{2012}$ or $ak \equiv bk \pmod{2012}$ , which is impossible to satisfy, given the domain $a, b < 2012$ . So, we have $2$ values of $k$ . In total, we have $2 + 1005 = 1007$ values of $k$ which makes $r_{a} = r_{b}$ or $r_{a} = 0$ , so there are $2011 - 1007 = 1004$ values of $k$ for which $r_{a} \neq r_{b}$ and $r_{a} \neq 0$ . Thus, by our reasoning above, our solution is $\frac{1}{2} \cdot 1004 = \boxed{502}$ . Solution by $\textbf{\underline{Invoker}}$ Solution 3 The key insight in this problem is noticing that when $ak$ is higher than $bk$ , $a(2012-k)$ is lower than $b(2012-k)$ , except at $2 \pmod{4}$ residues*. Also, they must be equal many times. $2012=2^2*503$ . We should have multiples of $503$ . After trying all three pairs and getting $503$ as our answer, we win. But look at the $2\pmod{4}$ idea. What if we just took $2$ and plugged it in with $1006$ ? We get $502$ . -- Va2010 11:12, 28 April 2012 (EDT)va2010 Solution 4 Say that the problem is a race track with $2012$ spots. To intersect the most, we should get next to each other a lot so the negation is high. As $2012=2^2*503$ , we intersect at a lot of multiples of $503$ .
\[ S = 502 \]
usajmo
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations" ]
7
Find all real-coefficient polynomials $f(x)$ which satisfy the following conditions: [b]i.[/b] $f(x) = a_0 x^{2n} + a_2 x^{2n - 2} + \cdots + a_{2n - 2} x^2 + a_{2n}, a_0 > 0$; [b]ii.[/b] $\sum_{j=0}^n a_{2j} a_{2n - 2j} \leq \left( \begin{array}{c} 2n\\ n\end{array} \right) a_0 a_{2n}$; [b]iii.[/b] All the roots of $f(x)$ are imaginary numbers with no real part.
We are tasked with finding all real-coefficient polynomials \( f(x) \) that satisfy the following conditions: 1. \( f(x) = a_0 x^{2n} + a_2 x^{2n - 2} + \cdots + a_{2n - 2} x^2 + a_{2n} \), where \( a_0 > 0 \). 2. \( \sum_{j=0}^n a_{2j} a_{2n - 2j} \leq \binom{2n}{n} a_0 a_{2n} \). 3. All the roots of \( f(x) \) are imaginary numbers with no real part. To solve this, we note that by condition (iii), the roots of \( f(x) \) are purely imaginary. Let the roots be \( \pm \alpha_1 i, \pm \alpha_2 i, \ldots, \pm \alpha_n i \), where \( \alpha_j > 0 \) for all \( j \). This implies that \( f(x) \) can be factored as: \[ f(x) = a_0 (x^2 + \alpha_1^2)(x^2 + \alpha_2^2) \cdots (x^2 + \alpha_n^2). \] Using Vieta's formulas, we express the coefficients \( a_{2k} \) in terms of the roots: \[ \frac{a_{2k}}{a_0} = \sum_{|S|=k, S \in Z'} \left( \prod_{a \in S} a \right)^2, \] where \( Z' = \{ \alpha_1 i, \alpha_2 i, \ldots, \alpha_n i \} \). Condition (ii) can be rewritten using these coefficients: \[ \sum_{j=0}^n a_{2j} a_{2n - 2j} \leq \binom{2n}{n} a_0 a_{2n}. \] By applying the Cauchy-Schwarz inequality and the Vandermonde identity, we find that equality holds if and only if all \( \alpha_j \) are equal. Therefore, all \( \alpha_j \) must be the same, say \( \alpha \). Thus, the polynomial \( f(x) \) simplifies to: \[ f(x) = a_0 (x^2 + \alpha^2)^n, \] where \( a_0 > 0 \) and \( \alpha \in \mathbb{R} \setminus \{0\} \). Hence, the polynomials that satisfy the given conditions are: \[ f(x) = a_0 (x^2 + \alpha^2)^n, \] where \( a_0 > 0 \) and \( \alpha \in \mathbb{R} \setminus \{0\} \). The answer is: \boxed{f(x) = a_0 (x^2 + \alpha^2)^n \text{ where } a_0 > 0 \text{ and } \alpha \in \mathbb{R} \setminus \{0\}}.
f(x) = a_0 (x^2 + \alpha^2)^n \text{ where } a_0 > 0 \text{ and } \alpha \in \mathbb{R} \setminus \{0\}
china_team_selection_test
[ "Mathematics -> Algebra -> Abstract Algebra -> Group Theory" ]
6.5
Let $\mathbb Z$ be the set of all integers. Find all pairs of integers $(a,b)$ for which there exist functions $f:\mathbb Z\rightarrow\mathbb Z$ and $g:\mathbb Z\rightarrow\mathbb Z$ satisfying \[f(g(x))=x+a\quad\text{and}\quad g(f(x))=x+b\] for all integers $x$ .
We claim that the answer is $|a|=|b|$ . Proof: $f$ and $g$ are surjective because $x+a$ and $x+b$ can take on any integral value, and by evaluating the parentheses in different order, we find $f(g(f(x)))=f(x+b)=f(x)+a$ and $g(f(g(x)))=g(x+a)=g(x)+b$ . We see that if $a=0$ then $g(x)=g(x)+b$ to $b=0$ as well, so similarly if $b=0$ then $a=0$ , so now assume $a, b\ne 0$ . We see that if $x=|b|n$ then $f(x)\equiv f(0) \pmod{|a|}$ , if $x=|b|n+1$ then $f(x)\equiv f(1)\pmod{|a|}$ , if $x=|b|n+2$ then $f(x)\equiv f(2)\pmod{|a|}$ ... if $x=|b|(n+1)-1$ then $f(x)\equiv f(|b|-1)\pmod{|a|}$ . This means that the $b$ -element collection $\left\{f(0), f(1), f(2), ... ,f(|b|-1)\right\}$ contains all $|a|$ residues mod $|a|$ since $f$ is surjective, so $|b|\ge |a|$ . Doing the same to $g$ yields that $|a|\ge |b|$ , so this means that only $|a|=|b|$ can work. For $a=b$ let $f(x)=x$ and $g(x)=x+a$ , and for $a=-b$ let $f(x)=-x$ and $g(x)=-x-a$ , so $|a|=|b|$ does work and are the only solutions, as desired. -Stormersyle
\[ |a| = |b| \]
usajmo
[ "Mathematics -> Number Theory -> Prime Numbers", "Mathematics -> Algebra -> Abstract Algebra -> Group Theory" ]
7
Determine which integers $n > 1$ have the property that there exists an infinite sequence $a_1$ , $a_2$ , $a_3$ , $\dots$ of nonzero integers such that the equality \[a_k + 2a_{2k} + \dots + na_{nk} = 0\] holds for every positive integer $k$ .
(Since Bertrand's is well known and provable using elementary techniques, I see nothing wrong with this-tigershark22) For $n=2$ , $|a_1| = 2 |a_2| = \cdots = 2^m |a_{2^m}|$ implies that for any positive integer $m$ , $|a_1| \ge 2^m$ , which is impossible. We proceed to prove that the infinite sequence exists for all $n\ge 3$ . First, one notices that if we have $a_{xy} = a_x a_y$ for any integers $x$ and $y$ , then it is suffice to define all $a_x$ for $x$ prime, and one only needs to verify the equation (*) \[a_1+2a_2+\cdots+na_n=0\] for the other equations will be automatically true. To proceed with the construction, I need the following fact: for any positive integer $m>2$ , there exists a prime $p$ such that $\frac{m}{2} <p \le m$ . To prove this, I am going to use Bertrand's Postulate ( [1] ) without proof. The Theorem states that, for any integer $n>1$ , there exists a prime $p$ such that $n<p\le 2n-1$ . In other words, for any positive integer $m>2$ , if $m=2n$ with $n>1$ , then there exists a prime $p$ such that $\frac{m}{2} < p < m$ , and if $m=2n-1$ with $n>1$ , then there exists a prime $p$ such that $\frac{m+1}{2} <p\le m$ , both of which guarantees that for any integer $m>2$ , there exists a prime $p$ such that $\frac{m}{2} <p \le m$ . Go back to the problem. Suppose $n\ge 3$ . Let the largest two primes not larger than $n$ are $P$ and $Q$ , and that $n\ge P > Q$ . By the fact stated above, one can conclude that $2P > n$ , and that $4Q = 2(2Q) \ge 2P > n$ . Let's construct $a_n$ : Let $a_1=1$ . There will be three cases: (i) $Q>\frac{n}{2}$ , (ii) $\frac{n}{2} \ge Q > \frac{n}{3}$ , and (iii) $\frac{n}{3} \ge Q > \frac{n}{4}$ . Case (i): $2Q>n$ . Let $a_x = 1$ for all prime numbers $x<Q$ , and $a_{xy}=a_xa_y$ , then (*) becomes: \[Pa_P + Qa_Q = C_1\] Case (ii): $2Q\le n$ but $3Q > n$ . In this case, let $a_2=-1$ , and $a_x = 1$ for all prime numbers $2<x<Q$ , and $a_{xy}=a_xa_y$ , then (*) becomes: \[Pa_P + Qa_Q - Qa_{2Q} = C_2\] or \[Pa_P - Qa_Q = C_2\] Case (iii): $3Q\le n$ . In this case, let $a_2=3$ , $a_3=-2$ , and $a_x = 1$ for all prime numbers $3<x<Q$ , and $a_{xy}=a_xa_y$ , then (*) becomes: \[Pa_P + Qa_Q + 3Qa_{2Q} - 2Qa_{3Q} = C_3\] or \[Pa_P + Qa_Q = C_3\] In each case, by Bezout's Theorem, there exists non zero integers $a_P$ and $a_Q$ which satisfy the equation. For all other primes $p > P$ , just let $a_p=1$ (or any other non-zero integer). This construction is correct because, for any $k> 1$ , \[a_k + 2 a_{2k} + \cdots n a_{nk} = a_k (1 + 2 a_2 + \cdots n a_n ) = 0\] Since Bertrand's Theorem is not elementary, we still need to wait for a better proof. -- Lightest 21:24, 2 May 2012 (EDT)
The integers \( n > 1 \) that have the property are \( n \geq 3 \).
usamo
[ "Mathematics -> Number Theory -> Other", "Mathematics -> Algebra -> Algebra -> Equations and Inequalities" ]
7
( Dick Gibbs ) For a given positive integer $k$ find, in terms of $k$ , the minimum value of $N$ for which there is a set of $2k+1$ distinct positive integers that has sum greater than $N$ but every subset of size $k$ has sum at most $N/2$ .
Solution 1 Let one optimal set of integers be $\{a_1,\dots,a_{2k+1}\}$ with $a_1 > a_2 > \cdots > a_{2k+1} > 0$ . The two conditions can now be rewritten as $a_1+\cdots + a_k \leq N/2$ and $a_1+\cdots +a_{2k+1} > N$ . Subtracting, we get that $a_{k+1}+\cdots + a_{2k+1} > N/2$ , and hence $a_{k+1}+\cdots + a_{2k+1} > a_1+\cdots + a_k$ . In words, the sum of the $k+1$ smallest numbers must exceed the sum of the $k$ largest ones. Let $a_{k+1}=C$ . As all the numbers are distinct integers, we must have $\forall i \in\{1,\dots,k\}:~ a_{k+1-i} \geq C+i$ , and also $\forall i \in\{1,\dots,k\}:~ a_{k+1+i} \leq C-i$ . Thus we get that $a_1+\cdots + a_k \geq kC + \dfrac{k(k+1)}2$ , and $a_{k+1}+\cdots + a_{2k+1} \leq (k+1)C - \dfrac{k(k+1)}2$ . As we want the second sum to be larger, clearly we must have $(k+1)C - \dfrac{k(k+1)}2 > kC + \dfrac{k(k+1)}2$ . This simplifies to $C > k(k+1)$ . Hence we get that: \begin{align*} N & \geq 2(a_1+\cdots + a_k) \\ & \geq 2\left( kC + \dfrac{k(k+1)}2 \right) \\ & = 2kC + k(k+1) \\ & \geq 2k(k^2+k+1) + k(k+1) \\ & = 2k^3 + 3k^2 + 3k \end{align*} On the other hand, for the set $\{ k^2+k+1+i ~|~ i\in\{-k,\dots,k\} \, \}$ the sum of the largest $k$ elements is exactly $k^3 + k^2 + k + \dfrac{k(k+1)}2$ , and the sum of the entire set is $(k^2+k+1)(2k+1) = 2k^3 + 3k^2 + 3k + 1$ , which is more than twice the sum of the largest set. Hence the smallest possible $N$ is $\boxed{ N = 2k^3 + 3k^2 + 3k }$ . Alternate solutions are always welcome. If you have a different, elegant solution to this problem, please add it to this page.
\[ N = 2k^3 + 3k^2 + 3k \]
usamo
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations", "Mathematics -> Algebra -> Algebra -> Equations and Inequalities" ]
8
Find all polynomials $P$ with real coefficients such that \[\frac{P(x)}{yz}+\frac{P(y)}{zx}+\frac{P(z)}{xy}=P(x-y)+P(y-z)+P(z-x)\] holds for all nonzero real numbers $x,y,z$ satisfying $2xyz=x+y+z$ .
If $P(x)=c$ for a constant $c,$ then $\dfrac{c(x+y+z)}{xyz}=3c$ . We have $2c=3c.$ Therefore $c=0.$ Now consider the case of non-constant polynomials. First we have \[xP(x)+yP(y)+zP(z)=xyz(P(x-y)+P(y-z)+P(z-x))\] for all nonzero real numbers $x,y,z$ satisfying $2xyz=x+y+z$ . Both sides of the equality are polynomials (of $x,y,z$ ). They have the same values on the 2-dimensional surface $2xyz=x+y+z$ , except for some 1-dimensional curves in it. By continuity, the equality holds for all points on the surface, including those with $z=0.$ Let $z=0,$ we have $y=-x$ and $x(P(x)-P(-x))=0.$ Therefore $P$ is an even function. (Here is a sketch of an elementary proof. Let $z=\dfrac{x+y}{2xy-1}.$ We have \[xP(x)+yP(y)+\dfrac{x+y}{2xy-1}P(\dfrac{x+y}{2xy-1})=xy\dfrac{x+y}{2xy-1}(P(x-y)+P(y-\dfrac{x+y}{2xy-1})+P(\dfrac{x+y}{2xy-1}-x)).\] This is an equality of rational expressions. By multiplying $(2xy-1)^N$ on both sides for a sufficiently large $N$ , they become polynomials, say $A(x,y)=B(x,y)$ for all real $x, y$ with $x\ne 0, y\ne 0, x+y\ne 0$ and $2xy-1\ne 0.$ For a fixed $x,$ we have two polynomials (of $y$ ) having same values for infinitely many $y$ . They must be identical. Let $y=0,$ we have $x^{N+1}(P(x)-P(-x))=0.$ ) Notice that if $P(x)$ is a solution, then is $cP(x)$ for any constant $c.$ For simplicity, we assume the leading coefficient of $P$ is $1$ : \[P(x)=x^n+a_{n-2}x^{n-2}+\cdots +a_2x^2+a_0,\] where $n$ is a positive even number. Let $y=\dfrac{1}{x}$ , $z=x+\dfrac{1}{x}.$ we have \[xP(x)+\dfrac{1}{x}P\left (\dfrac{1}{x}\right )+\left ( x+\dfrac{1}{x}\right ) P\left ( x+\dfrac{1}{x}\right ) =\left (x+\dfrac{1}{x}\right )\left ( P\left (x-\dfrac{1}{x}\right )+P(-x)+P\left (\dfrac{1}{x}\right )\right ).\] Simplify using $P(x)=P(-x),$ \[\left (x+\dfrac{1}{x}\right ) \left (P\left (x+\dfrac{1}{x}\right )-P\left (x-\dfrac{1}{x}\right )\right )=\dfrac{1}{x}P(x)+xP\left (\dfrac{1}{x}\right ).\] Expand and combine like terms, both sides are of the form \[c_{n-1}x^{n-1}+c_{n-3}x^{n-3}+\cdots+c_1x+c_{-1}x^{-1}+\cdots+c_{-n+1}x^{-n+1}.\] They have the same values for infinitely many $x.$ They must be identical. We just compare their leading terms. On the left hand side it is $2nx^{n-1}$ . There are two cases for the right hand sides: If $n>2$ , it is $x^{n-1}$ ; If $n=2$ , it is $(1+a_0)x.$ It does not work for $n>2.$ When $n=2,$ we have $4=1+a_0.$ therefore $a_0=3.$ The solution: $P(x)=c(x^2+3)$ for any constant $c.$ -JZ
\[ P(x) = c(x^2 + 3) \text{ for any constant } c. \]
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Other" ]
7
In a fictional world, each resident (viewed as geometric point) is assigned a number: $1,2, \cdots$. In order to fight against some epidemic, the residents take some vaccine and they stay at the vaccination site after taking the shot for observation. Now suppose that the shape of the Observation Room is a circle of radius $\frac{1}{4}$, and one requires that the distance $d_{m, n}$ between the Resident No. $m$ and the Resident No. $n$ must satisfy $(m+n) d_{m, n} \geq 1$. Where we consider the distance on the circle, i.e., the length of the minor arc between two points. Proof Question: Give a proof of your answer to Question (i).
Solution I. We can place the Residents No. $1,2, \ldots$ according to the following rule. First, put Resident No. 1 arbitrarily. For $n>2$, if Residents No. $1,2, \ldots, n-1$ have already been placed, we consider the positions where Resident No. n cannot be placed. For $1 \leq m \leq n-1$, by $d_{m, n} \geq \frac{1}{m+n}$, we know that the Resident No. $n$ cannot be placed in the arc that is centered at Resident No. $m$, and of the length $\frac{2}{m+n}$. The total length of these arcs is $\frac{2}{n+1}+\frac{2}{n+2}+\cdots+\frac{2}{2 n-1}<2\left(\ln \frac{n+1}{n}+\ln \frac{n+2}{n+1}+\cdots+\ln \frac{2 n-1}{2 n-2}\right)=2 \ln \frac{2 n-1}{n}<2 \ln 2$. Therefore, the total length of the union of these arcs does not exceed $2 \ln 2$, while the perimeter of the circle is $\frac{1}{4} \cdot 2 \pi=\frac{\pi}{2}$. It is easy to observe that $\frac{\pi}{2}>1.5>2 \ln 2$, so these arcs would not cover the whole circle, hence it is always possible to find a place for Resident No. $n$ such that its distances to Residents No. $1,2, \ldots, n-1$ satisfy the requirement. By induction we conclude that the circle can accommodate any quantity of residents.
The circle can accommodate any quantity of residents.
alibaba_global_contest
[ "Mathematics -> Algebra -> Intermediate Algebra -> Recursive Sequences -> Other" ]
8
Let $ \left(a_{n}\right)$ be the sequence of reals defined by $ a_{1}=\frac{1}{4}$ and the recurrence $ a_{n}= \frac{1}{4}(1+a_{n-1})^{2}, n\geq 2$. Find the minimum real $ \lambda$ such that for any non-negative reals $ x_{1},x_{2},\dots,x_{2002}$, it holds \[ \sum_{k=1}^{2002}A_{k}\leq \lambda a_{2002}, \] where $ A_{k}= \frac{x_{k}-k}{(x_{k}+\cdots+x_{2002}+\frac{k(k-1)}{2}+1)^{2}}, k\geq 1$.
Let \( \left(a_n\right) \) be the sequence of reals defined by \( a_1 = \frac{1}{4} \) and the recurrence \( a_n = \frac{1}{4}(1 + a_{n-1})^2 \) for \( n \geq 2 \). We aim to find the minimum real \( \lambda \) such that for any non-negative reals \( x_1, x_2, \dots, x_{2002} \), it holds that \[ \sum_{k=1}^{2002} A_k \leq \lambda a_{2002}, \] where \( A_k = \frac{x_k - k}{(x_k + \cdots + x_{2002} + \frac{k(k-1)}{2} + 1)^2} \) for \( k \geq 1 \). First, we simplify the problem by setting \( t = 2002 \). For \( k = 1, 2, \dots, t \), define \[ y_k = x_k + x_{k+1} + \dots + x_t + \frac{k(k-1)}{2} + 1, \] and let \( L = y_{t+1} = \frac{(t+1)t}{2} + 1 \). Notice that \( y_k - y_{k+1} = x_k - k \) for \( 1 \leq k \leq t \). Thus, we need to maximize the sum \[ S = \sum_{k=1}^{2002} A_k = \sum_{k=1}^{2002} \frac{y_k - y_{k+1}}{y_k^2}. \] We use the following lemma to proceed: **Lemma 1.** The inequality \( \frac{ax - b}{x^2} \leq \frac{a^2}{4} \cdot \frac{1}{b} \) holds for all \( x \in \mathbb{R} \setminus \{0\} \) with equality when \( x = \frac{2b}{a} \), where \( a, b > 0 \). **Proof.** Multiplying by \( 4bx^2 > 0 \), we need \( 4abx - 4b^2 \leq a^2x^2 \), which simplifies to \( (ax - 2b)^2 \geq 0 \). \( \blacksquare \) **Lemma 2.** Define the sequence \( b_1 = 0 \) and \( b_n = \frac{1}{4}(1 + b_{n-1})^2 \) for \( n \geq 2 \). Then \[ \frac{b_k}{y_k} + \frac{y_k - y_{k+1}}{y_k^2} \leq \frac{b_{k+1}}{y_{k+1}} \] for all \( 1 \leq k \leq n \). **Proof.** Using Lemma 1, we find \[ \frac{b_k}{y_k} + \frac{y_k - y_{k+1}}{y_k^2} = \frac{(b_k + 1)y_k - y_{k+1}}{y_k^2} \leq \frac{(b_k + 1)^2}{4} \cdot \frac{1}{y_{k+1}} = \frac{b_{k+1}}{y_{k+1}}. \quad \blacksquare \] Summing these inequalities for \( k = 1, 2, \dots, t \) gives \[ 0 \geq \sum_{k=1}^t \left( \frac{b_k}{y_k} + \frac{y_k - y_{k+1}}{y_k^2} - \frac{b_{k+1}}{y_{k+1}} \right) = S - \frac{b_{t+1}}{y_{t+1}}, \] so \( S \leq \frac{b_{t+1}}{L} \). To achieve the maximum with non-negative \( x_k \), equality holds if and only if \( y_k = \frac{2y_{k+1}}{b_k + 1} \) for \( k = 1, 2, \dots, t \). This ensures all \( y_k \) are positive. Induction shows \( 0 \leq b_n \leq 1 \) for all \( n \geq 1 \), implying \( y_k = \frac{2}{b_k + 1} y_{k+1} \geq y_{k+1} \), ensuring \( x_k \geq 0 \). Since \( b_2 = \frac{1}{4} \) and \( b_{n+1} = a_n \), the maximum \( S = \frac{b_{t+1}}{L} = \frac{1}{L} a_t \). Thus, the constant \( \lambda \) is \[ \lambda = \frac{1}{\frac{2003 \cdot 2002}{2} + 1} = \frac{1}{2005004}. \] The answer is: \boxed{\frac{1}{2005004}}.
\frac{1}{2005004}
china_team_selection_test
[ "Mathematics -> Number Theory -> Prime Numbers", "Mathematics -> Algebra -> Abstract Algebra -> Field Theory" ]
6.5
Let $\mathbb{N}$ denote the set of positive integers. Find all functions $f : \mathbb{N} \rightarrow \mathbb{N}$ such that for positive integers $a$ and $b,$ \[f(a^2 + b^2) = f(a)f(b) \text{ and } f(a^2) = f(a)^2.\]
I claim that the only function $f$ that satisfies the constraints outlined within the problem is the function $f(n) = 1$ for all positive integers $n$ . We will proceed with strong induction. The base case is simple, as plugging $a=1$ into the second equation given within the problem gives $f(1)=f(1)^2$ . Since $f(n)$ can only return a positive integer value, we have that $f(1)=1$ . Now we proceed with the inductive step. If the next number $n$ is either a perfect square or can be represented as a sum of two perfect squares, then obviously $f(n)=1$ , as it is either the product of two $f$ -values that are both equal to $1$ from the inductive assumption or is the square of an $f$ -value that is equal to $1$ , again due to the inductive assumption. Otherwise, we can use the Sum of Two Squares Theorem, which tells us that $n$ has at least one prime in its prime factorization that is $3(\mod 4)$ and is raised to an odd power. Lemma 1: Given that $a^2+b^2=c^2$ and $f(b)=f(c)=1$ , we then have $f(a)=1$ . Proof: Note that the first condition in the problem tells us that $f(a^2+b^2)=f(a)f(b)$ , or $f(c^2)=f(a)f(b)$ . Using the second condition gives us $f(c)^2=f(a)f(b)$ . Plugging in the values of $f(b)$ and $f(c)$ gives us that $f(a)=1$ . Now we will attempt to repeatedly remove prime factors that are $3(mod 4)$ and taken to an odd power, and we will move from the largest prime down to the smallest prime that satisfies above conditions. The prime factors will be removed by constructing a Pythagorean Triple with the prime being the smallest leg in the form of $p,\frac{p^2-1}{2},\frac{p^2+1}{2}$ (this will always work as $p\neq 2$ (not 3 mod 4), and this method works via Lemma 1). We will then prove the ending numbers that we achieve via removing all the $(\text{mod } 4)$ primes(which I will refer to as "tips") are equal to 1 b/c they can be expressed as the sum of two squares or is a perfect square(Sum of Two Squares Theorem). For example, if we took the number $21$ , we would first aim to remove the $7$ by splitting it into $24$ and $25$ , so $21$ would become $72$ and $75$ . $75$ is divisible by $3$ to an odd power, so we transform it into $100$ and $125$ . These two don't have any divisors that are $3 (\text{mod } 4)$ raised to an odd power so we leave it alone, as sum of two squares will work on them or they are perfect squares. $72$ does have a prime factor that is $3(\text{mod } 4)$ , but it is an even power so we leave it alone. In this case, the tips are $72$ , $100$ , and $125$ . However, now we need to prove two key facts: using this Pythagorean Triple Method will never generate another $3(\text{mod } 4)$ prime that is bigger than the current one we are working on or create more primes or keep the same number of primes in the tips' prime factorizations., as otherwise it could cause an infinite cycle, and also we must prove when we use Sum of Two Squares/Perfect Square given in the second condition on the tips the square root of the square(s) used will never be greater than or equal to $n$ . The first claim can be proved rather simply. Note that $\frac{p^2-1}{2}$ can have no prime factors greater than or equal to $p$ , as it can be factored as $(p-1)(\frac{p+1}{2})$ , which are both less than $p$ and are integers( $p+1$ must be an integer due to $p$ must being odd(not equal to $2$ )). For $\frac{p^2+1}{2}$ , we can prove something a bit more general. Lemma 2: For any positive integer $n$ , all prime factors of $n^2+1$ must be $1(\text{mod } 4)$ . Proof: Note that if $n^2+1\equiv 0(\text{mod } p)$ , then we also must have $(n^2+1)(n^2-1)=(n^4-1)\equiv 0(\text{mod } p)$ , or $n^4 \equiv 1(\text{mod } p)$ . Now we can apply Fermat's Little Theorem to obtain $n^{p-1}\equiv 1(\text{mod } p)$ . Note that since $n$ and $n^2$ are obviously not $1(\text{mod } p)$ , as $n^2\equiv -1 (\text{mod } p)$ , we have that $p-1$ is a multiple of $4$ , or $p \equiv 1(\text{mod } 4)$ . We can simply apply Lemma 2 to $p^2+1$ as $2$ is obviously not $3(\text{mod } 4)$ , and this means that a prime factor $3(\text{mod } 4)$ will never be generated from this term. This completes the first of our two claims. Now we proceed to the second of our two claims. Note that every time we use the method on $n$ based on prime $p$ , we will multiply by around $\frac{p}{2}$ , clearly less than $p$ (if we multiply by $p$ we would get $p^2$ which is clearly greater than $\frac{p^2+1}{2}$ ). We will never have to use the same prime twice in our method, so at max in the end we will multiply $n$ by the product of a little less than all $3(mod 4)$ primes that divide it, which is less than $n$ itself for $n$ greater than or equal to $3$ (the smallest 3 mod 4 prime), meaning that the largest number that we must use two squares on that is generated by out method is less than $n^2$ . We need to prove that the two squares that sum to this are both less than $n$ , which is quite trivial, as they are less than $\sqrt{n^2}$ , which obviously means it must be less than $n$ . This proves the second of our claims. This completes the second case of the inductive step, and therefore completes both the induction and the problem. ~Solution by hyxue
The only function \( f \) that satisfies the given conditions is \( f(n) = 1 \) for all positive integers \( n \).
usajmo
[ "Mathematics -> Number Theory -> Factorization", "Mathematics -> Algebra -> Algebra -> Equations and Inequalities" ]
6.5
( Ricky Liu ) Find all positive integers $n$ such that there are $k\ge 2$ positive rational numbers $a_1, a_2, \ldots, a_k$ satisfying $a_1 + a_2 + \cdots + a_k = a_1\cdot a_2\cdots a_k = n$ .
Solution 1 First, consider composite numbers. We can then factor $n$ into $p_1p_2.$ It is easy to see that $p_1+p_2\le n$ , and thus, we can add $(n-p_1-p_2)$ 1s in order to achieve a sum and product of $n$ . For $p_1+p_2=n$ , which is only possible in one case, $n=4$ , we consider $p_1=p_2=2$ . Secondly, let $n$ be a prime. Then we can find the following procedure: Let $a_1=\frac{n}{2}, a_2=4, a_3=\frac{1}{2}$ and let the rest of the $a_k$ be 1. The only numbers we now need to check are those such that $\frac{n}{2}+4+\frac{1}{2}>n\Longrightarrow n<9$ . Thus, we need to check for $n=1,2,3,5,7$ . One is included because it is neither prime nor composite. For $n=1$ , consider $a_1a_2\hdots a_k=1$ . Then by AM-GM, $a_1+a_2+\hdots+a_k\ge k\sqrt[k]{1}>1$ for $k\ge 2$ . Thus, $n=1$ is impossible. If $n=2$ , once again consider $a_1a_2\hdots a_k=2$ . Similar to the above, $a_1+a_2+\hdots\ge k\sqrt[k]{2}>2$ for $k\ge 2$ since $\sqrt[k]{2}>1$ and $k>2$ . Obviously, $n=2$ is then impossible. If $n=3$ , let $a_1a_2\hdots a_k=3$ . Again, $a_1+a_2+\hdots\ge k\sqrt[k]{3}>3$ . This is obvious for $k\ge 3$ . Now consider $k=2$ . Then $2\sqrt{3}\approx 3.4$ is obviously greater than $3$ . Thus, $n=3$ is impossible. If $n=5$ , proceed as above and consider $k=2$ . Then $a_1+a_2=5$ and $a_1a_2=5$ . However, we then come to the quadratic $a_1^2-5a_1+5=0 \Longrightarrow a_1=\frac{5\pm\sqrt{5}}{2}$ , which is not rational. For $k=3$ and $k=4$ we note that $\sqrt[3]{5}>\frac{5}{3}$ and $\sqrt[4]{5}>\frac{5}{4}$ . This is trivial to prove. If $k\ge 5$ , it is obviously impossible, and thus $n=5$ does not work. The last case, where $n=7$ , is possible using the following three numbers. $a_1=\frac{9}{2}, a_2=\frac{4}{3}, a_3=\frac{7}{6}$ shows that $n=7$ is possible. Hence, $n$ can be any positive integer greater than $3$ with the exclusion of $5$ . Alternate solutions are always welcome. If you have a different, elegant solution to this problem, please add it to this page.
The positive integers \( n \) such that there are \( k \geq 2 \) positive rational numbers \( a_1, a_2, \ldots, a_k \) satisfying \( a_1 + a_2 + \cdots + a_k = a_1 \cdot a_2 \cdots a_k = n \) are: \[ n = 4 \text{ or } n \geq 6 \]
usamo
[ "Mathematics -> Applied Mathematics -> Statistics -> Probability -> Counting Methods -> Combinations" ]
4
A random number selector can only select one of the nine integers 1, 2, ..., 9, and it makes these selections with equal probability. Determine the probability that after $n$ selections ( $n>1$ ), the product of the $n$ numbers selected will be divisible by 10.
For the product to be divisible by 10, there must be a factor of 2 and a factor of 5 in there. The probability that there is no 5 is $\left( \frac{8}{9}\right)^n$ . The probability that there is no 2 is $\left( \frac{5}{9}\right)^n$ . The probability that there is neither a 2 nor 5 is $\left( \frac{4}{9}\right)^n$ , which is included in both previous cases. The only possibility left is getting a 2 and a 5, making the product divisible by 10. By complementarity and principle of inclusion-exclusion, the probability of that is $1- \left( \left( \frac{8}{9}\right)^n + \left( \frac{5}{9}\right)^n - \left( \frac{4}{9}\right)^n\right)=\boxed{1-(8/9)^n-(5/9)^n+(4/9)^n}$ . Alternate solutions are always welcome. If you have a different, elegant solution to this problem, please add it to this page.
\[ 1 - \left( \frac{8}{9} \right)^n - \left( \frac{5}{9} \right)^n + \left( \frac{4}{9} \right)^n \]
usamo
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
8
Let $X_1, X_2, \ldots, X_{100}$ be a sequence of mutually distinct nonempty subsets of a set $S$ . Any two sets $X_i$ and $X_{i+1}$ are disjoint and their union is not the whole set $S$ , that is, $X_i\cap X_{i+1}=\emptyset$ and $X_i\cup X_{i+1}\neq S$ , for all $i\in\{1, \ldots, 99\}$ . Find the smallest possible number of elements in $S$ .
The answer is that $|S| \ge 8$ . First, we provide a inductive construction for $S = \left\{ 1, \dots, 8 \right\}$ . Actually, for $n \ge 4$ we will provide a construction for $S = \left\{ 1, \dots, n \right\}$ which has $2^{n-1} + 1$ elements in a line. (This is sufficient, since we then get $129$ for $n = 8$ .) The idea is to start with the following construction for $|S| = 4$ : \[\begin{array}{ccccccccc} 34 & 1 & 23 & 4 & 12 & 3 & 14 & 2 & 13 \end{array}.\] Then inductively, we do the following procedure to move from $n$ to $n+1$ : take the chain for $n$ elements, delete an element, and make two copies of the chain (which now has even length). Glue the two copies together, joined by $\varnothing$ in between. Then place the element $n+1$ in alternating positions starting with the first (in particular, this hits $n+1$ ). For example, the first iteration of this construction gives: \[\begin{array}{ccccccccc} 345 & 1 & 235 & 4 & 125 & 3 & 145 & 2 & 5 \\ 34 & 15 & 23 & 45 & 12 & 35 & 14 & 25 & \end{array}\] Now let's check $|S| \ge 8$ is sufficient. Consider a chain on a set of size $|S| = 7$ . (We need $|S| \ge 7$ else $2^{|S|} < 100$ .) Observe that there are sets of size $\ge 4$ can only be neighbored by sets of size $\le 2$ , of which there are $\binom 71 + \binom 72 = 28$ . So there are $\le 30$ sets of size $\ge 4$ . Also, there are $\binom 73 = 35$ sets of size $3$ . So the total number of sets in a chain can be at most $30 + 28 + 35 = 93 < 100$ .
\[ |S| \ge 8 \]
usamo
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations" ]
5
Let $f(x)$ be a degree 2006 polynomial with complex roots $c_{1}, c_{2}, \ldots, c_{2006}$, such that the set $$\left\{\left|c_{1}\right|,\left|c_{2}\right|, \ldots,\left|c_{2006}\right|\right\}$$ consists of exactly 1006 distinct values. What is the minimum number of real roots of $f(x)$ ?
The complex roots of the polynomial must come in pairs, $c_{i}$ and $\overline{c_{i}}$, both of which have the same absolute value. If $n$ is the number of distinct absolute values $\left|c_{i}\right|$ corresponding to those of non-real roots, then there are at least $2 n$ non-real roots of $f(x)$. Thus $f(x)$ can have at most $2006-2 n$ real roots. However, it must have at least $1006-n$ real roots, as $\left|c_{i}\right|$ takes on $1006-n$ more values. By definition of $n$, these all correspond to real roots. Therefore $1006-n \leq \#$ real roots $\leq 2006-2 n$, so $n \leq 1000$, and \# real roots $\geq 1006-n \geq 6$. It is easy to see that equality is attainable.
The minimum number of real roots of \( f(x) \) is \( 6 \).
HMMT_2
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
4.5
Each lattice point with nonnegative coordinates is labeled with a nonnegative integer in such a way that the point $(0,0)$ is labeled by 0 , and for every $x, y \geq 0$, the set of numbers labeled on the points $(x, y),(x, y+1)$, and $(x+1, y)$ is \{n, n+1, n+2\} for some nonnegative integer $n$. Determine, with proof, all possible labels for the point $(2000,2024)$.
We claim the answer is all multiples of 3 from 0 to $2000+2 \cdot 2024=6048$. First, we prove no other values are possible. Let $\ell(x, y)$ denote the label of cell $(x, y)$. \section*{The label is divisible by 3.} Observe that for any $x$ and $y, \ell(x, y), \ell(x, y+1)$, and \ell(x+1, y)$ are all distinct mod 3 . Thus, for any $a$ and $b, \ell(a+1, b+1)$ cannot match \ell(a+1, b)$ or \ell(a, b+1) \bmod 3$, so it must be equivalent to \ell(a, b)$ modulo 3 . Since \ell(a, b+1), \ell(a, b+2), \ell(a+1, b+1)$ are all distinct \bmod 3$, and \ell(a+1, b+1)$ and \ell(a, b)$ are equivalent \bmod 3$, then \ell(a, b), \ell(a, b+1), \ell(a, b+2)$ are all distinct \bmod 3$, and thus similarly \ell(a, b+$ $1), \ell(a, b+2), \ell(a, b+3)$ are all distinct \bmod 3$, which means that \ell(a, b+3)$ must be neither \ell(a, b+1)$ or \ell(a, b+2) \bmod 3$, and thus must be equal to \ell(a, b) \bmod 3$. These together imply that $$\ell(w, x) \equiv \ell(y, z) \bmod 3 \Longleftrightarrow w-x \equiv y-z \bmod 3$$ It follows that \ell(2000,2024)$ must be equivalent to \ell(0,0) \bmod 3$, which is a multiple of 3 . \section*{The label is at most 6048 .} Note that since \ell(x+1, y), \ell(x, y+1)$, and \ell(x, y)$ are 3 consecutive numbers, \ell(x+1, y)-\ell(x, y)$ and \ell(x, y+1)-\ell(x, y)$ are both \leq 2$. Moreover, since \ell(x+1, y+1) \leq \ell(x, y)+4$, since it is also the same mod 3 , it must be at most \ell(x, y)+3$. Thus, \ell(2000,2000) \leq \ell(0,0)+3 \cdot 2000$, and \ell(2000,2024) \leq \ell(2000,2000)+2 \cdot 24$, so \ell(2000,2024) \leq 6048$. \section*{Construction.} Consider lines \ell_{n}$ of the form $x+2 y=n$ (so $(2000,2024)$ lies on \ell_{6048}$ ). Then any three points of the form $(x, y),(x, y+1)$, and $(x+1, y)$ lie on three consecutive lines \ell_{n}, \ell_{n+1}, \ell_{n+2}$ in some order. Thus, for any $k$ which is a multiple of 3 , if we label every point on line \ell_{i}$ with \max (i \bmod 3, i-k)$, any three consecutive lines \ell_{n}, \ell_{n+1}, \ell_{n+2}$ will either be labelled 0,1 , and 2 in some order, or $n-k, n-k+1$, $n-k+2$, both of which consist of three consecutive numbers. Below is an example with $k=6$. \begin{tabular}{|l|l|l|l|l|l|l|l|l|} \hline 8 & 9 & 10 & 11 & 12 & 13 & 14 & 15 \\ \hline 6 & 7 & 8 & 9 & 10 & 11 & 12 & 13 \\ \hline 4 & 5 & 6 & 7 & 8 & 9 & 10 & 11 \\ \hline 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 \\ \hline 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 \\ \hline 1 & 2 & 0 & 1 & 2 & 3 & 4 & 5 \\ \hline 2 & 0 & 1 & 2 & 0 & 1 & 2 & 3 \\ \hline 0 & 1 & 2 & 0 & 1 & 2 & 0 & 1 \\ \hline \end{tabular} Any such labelling is valid, and letting $k$ range from 0 to 6048 , we see $(2000,2024)$ can take any label of the form $6048-k$, which spans all such multiples of 3 . Hence the possible labels are precisely the multiples of 3 from 0 to 6048.
The possible labels for the point $(2000, 2024)$ are precisely the multiples of 3 from 0 to 6048.
HMMT_2
[ "Mathematics -> Number Theory -> Congruences" ]
7
Determine all non-negative integral solutions $(n_1,n_2,\dots , n_{14})$ if any, apart from permutations, of the Diophantine Equation $n_1^4+n_2^4+\cdots +n_{14}^4=1599$ .
Recall that $n_i^4\equiv 0,1\bmod{16}$ for all integers $n_i$ . Thus the sum we have is anything from 0 to 14 modulo 16. But $1599\equiv 15\bmod{16}$ , and thus there are no integral solutions to the given Diophantine equation.
There are no integral solutions to the given Diophantine equation.
usamo
[ "Mathematics -> Number Theory -> Congruences", "Mathematics -> Algebra -> Abstract Algebra -> Group Theory" ]
8
Let $p$ be an odd prime. An integer $x$ is called a quadratic non-residue if $p$ does not divide $x - t^2$ for any integer $t$ . Denote by $A$ the set of all integers $a$ such that $1 \le a < p$ , and both $a$ and $4 - a$ are quadratic non-residues. Calculate the remainder when the product of the elements of $A$ is divided by $p$ .
This problem needs a solution. If you have a solution for it, please help us out by adding it . 2020 USAMO ( Problems • Resources ) Preceded by Problem 2 Followed by Problem 4 1 • 2 • 3 • 4 • 5 • 6 All USAMO Problems and Solutions The problems on this page are copyrighted by the Mathematical Association of America 's American Mathematics Competitions .
There is no solution provided for this problem.
usamo
[ "Mathematics -> Discrete Mathematics -> Algorithms", "Mathematics -> Algebra -> Intermediate Algebra -> Other" ]
7
A game of solitaire is played with $R$ red cards, $W$ white cards, and $B$ blue cards. A player plays all the cards one at a time. With each play he accumulates a penalty. If he plays a blue card, then he is charged a penalty which is the number of white cards still in his hand. If he plays a white card, then he is charged a penalty which is twice the number of red cards still in his hand. If he plays a red card, then he is charged a penalty which is three times the number of blue cards still in his hand. Find, as a function of $R, W,$ and $B,$ the minimal total penalty a player can amass and all the ways in which this minimum can be achieved.
We claim (inductively) that the minimum is just going to be $\min(BW,2WR,3RB)$ . We'll start our induction with the case where one of the three quantities is zero, in which case we verify that we can indeed get away without any penalty by, for example, discarding blue if we are out of white. Now, for the inductive step, let $f(B,W,R)$ be the minimum we seek. Note that \[f(B,W,R) = \min(W+f(B-1,W,R),2R+f(B,W-1,R),3B+f(B,W,R-1))\] By our inductive hypothesis, $f(B-1,W,R) = \min((B-1)W,2WR,3R(B-1))$ . In order for this to cause our inductive step not to hold, we would require that $W+\min((B-1)W,2WR,3R(B-1)) < \min(BW,2WR,3RB)$ . It is evident that the first two entries in the $min$ expression cannot cause this to happen, so that we need only consider $W+3R(B-1) < \min(BW,2WR,3RB)$ . So $W+3R(B-1) < BW$ , whence $3R < W$ . But $W+3R(B-1) < 3RB$ , so that $W < 3R$ , a contradiction. For the other two cases, we can get similar contradictions, so that our inductive step must hold, and so $f(B,W,R)$ is indeed $\min(BW,2WR,3RB)$ . We now need only establish how many ways to do this. If one of these quantities is smaller, our induction and the fact that it is eventually zero guarantees that it will continue to be the smallest quantity as cards are discarded. (For example, if it is currently optimal to discard a blue card, it will continue to be so until we run out of blue cards.) Therefore, assuming that there is currently only one best choice of card to discard, this will continue to be so in the future, whence if $BW \neq 2WR \neq 3RB$ , there is only $1$ optimal strategy. Suppose, now, that $BW = 2WR$ . It is thus optimal to discard either a $B$ or $W$ card. If we ever discard a blue card, then we will cause $BW < 2WR$ , whence there is only one possible strategy from then on. However, if we discard a white card, then we will still have $BW = 2WR$ , meaning that we continue to have the choice of discarding a white or blue card. Since we can discard a white card at most $W$ times, there are $W+1$ choices for how many $W$ cards to discard ( $0$ to $W$ ), meaning that there are $W+1$ optimal strategies. By similar logic, we get $R+1$ optimal strategies if $2WR = 3RB$ , and $B+1$ optimal strategies if $3RB = BW$ . The final case, then, is if $BW = 2WR = 3RB$ . In this case, if we first discard a white card, we are left with the $BW = 2WR$ case, and similarly for a blue and red card. The total number of optimal strategies in this case is then the sum of the optimal strategies in those cases, or, in other words, $B+W+R$ . To summarize: The minimum penalty is $\min(BW,2WR,3RB)$ . If $BW \neq 2WR \neq 3RB$ , there is $1$ optimal strategy. If $BW = 2WR < 3RB$ , there are $W+1$ strategies. If $2WR = 3RB < BW$ , there are $R+1$ strategies. If $3RB = BW < 2WR$ , there are $B+1$ strategies. If $BW = 2WR = 3RB$ , there are $R+B+W$ strategies. By J Steinhardt, from AoPS Community
The minimal total penalty a player can amass is: \[ \min(BW, 2WR, 3RB) \] The number of optimal strategies is: - If \( BW \neq 2WR \neq 3RB \), there is \( 1 \) optimal strategy. - If \( BW = 2WR < 3RB \), there are \( W+1 \) strategies. - If \( 2WR = 3RB < BW \), there are \( R+1 \) strategies. - If \( 3RB = BW < 2WR \), there are \( B+1 \) strategies. - If \( BW = 2WR = 3RB \), there are \( R+B+W \) strategies.
usamo
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations" ]
5
Given are real numbers $x, y$. For any pair of real numbers $a_{0}, a_{1}$, define a sequence by $a_{n+2}=x a_{n+1}+y a_{n}$ for $n \geq 0$. Suppose that there exists a fixed nonnegative integer $m$ such that, for every choice of $a_{0}$ and $a_{1}$, the numbers $a_{m}, a_{m+1}, a_{m+3}$, in this order, form an arithmetic progression. Find all possible values of $y$.
Note that $x=1$ (or $x=0$ ), $y=0$ gives a constant sequence, so it will always have the desired property. Thus, $y=0$ is one possibility. For the rest of the proof, assume $y \neq 0$. We will prove that $a_{m}$ and $a_{m+1}$ may take on any pair of values, for an appropriate choice of $a_{0}$ and $a_{1}$. Use induction on $m$. The case $m=0$ is trivial. Suppose that $a_{m}$ and $a_{m+1}$ can take on any value. Let $p$ and $q$ be any real numbers. By setting $a_{m}=\frac{q-x p}{y}($ remembering that $y \neq 0)$ and $a_{m+1}=p$, we get $a_{m+1}=p$ and $a_{m+2}=q$. Therefore, $a_{m+1}$ and $a_{m+2}$ can have any values if $a_{m}$ and $a_{m+1}$ can. That completes the induction. Now we determine the nonzero $y$ such that $a_{m}, a_{m+1}, a_{m+3}$ form an arithmetic sequence; that is, such that $a_{m+3}-a_{m+1}=a_{m+1}-a_{m}$. But because $a_{m+3}=\left(x^{2}+y\right) a_{m+1}+x y a_{m}$ by the recursion formula, we can eliminate $a_{m+3}$ from the equation, obtaining the equivalent condition $\left(x^{2}+y-2\right) a_{m+1}+(x y+1) a_{m}=0$. Because the pair $a_{m}, a_{m+1}$ can take on any values, this condition means exactly that $x^{2}+y-2=x y+1=0$. Then $x=-1 / y$, and $1 / y^{2}+y-2=0$, or $y^{3}-2 y^{2}+1=0$. One root of this cubic is $y=1$, and the remaining quadratic factor $y^{2}-y-1$ has the roots $(1 \pm \sqrt{5}) / 2$. Since each such $y$ gives an $x$ for which the condition holds, we conclude that the answer to the problem is $y=0,1$, or $(1 \pm \sqrt{5}) / 2$.
\[ y = 0, 1, \frac{1 + \sqrt{5}}{2}, \frac{1 - \sqrt{5}}{2} \]
HMMT_2
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations", "Mathematics -> Algebra -> Abstract Algebra -> Other", "Mathematics -> Number Theory -> Prime Numbers" ]
8
( Titu Andreescu, Gabriel Dospinescu ) For integral $m$ , let $p(m)$ be the greatest prime divisor of $m$ . By convention, we set $p(\pm 1)=1$ and $p(0)=\infty$ . Find all polynomials $f$ with integer coefficients such that the sequence $\{ p(f(n^2))-2n) \}_{n \in \mathbb{Z} \ge 0}$ is bounded above. (In particular, this requires $f(n^2)\neq 0$ for $n\ge 0$ .)
Solution 1 Let $f(x)$ be a non-constant polynomial in $x$ of degree $d$ with integer coefficients, suppose further that no prime divides all the coefficients of $f$ (otherwise consider the polynomial obtained by dividing $f$ by the gcd of its coefficients). We further normalize $f$ by multiplying by $-1$ , if necessary, to ensure that the leading coefficient (of $x^d$ ) is positive. Let $g(n) = f(n^2)$ , then $g(n)$ is a polynomial of degree $2$ or more and $g(n) = g(-n)$ . Let $g_1, \ldots, g_k$ be the factorization of $g$ into irreducible factors with positive leading coefficients. Such a factorization is unique. Let $d(g_i)$ denote the degree of $g_i$ . Since $g(-n) = g(n)$ the factors $g_i$ are either even functions of $n$ or come in pairs $(g_i, h_i)$ with $g_i(-n) = (-1)^{d(g_i)} h_i(n)$ . Let $P(0) = \infty$ , $P(\pm 1) = 1$ . For any other integer $m$ let $P(m)$ be the largest prime factor of $m$ . Suppose that for some finite constant $C$ and all $n \ge 0$ we have $P(g(n)) - 2n < C$ . Since the polynomials $g_i$ divide $g$ , the same must be true for each of the irreducible polynomials $g_i$ . A theorem of T. Nagell implies that if $d(g_i) \ge 2$ the ratio $P(g_i(n))/n$ is unbounded for large values of $n$ . Since in our case the $P(g_i(n))/n$ is asymptotically bounded above by $2$ for large $n$ , we conclude that all the irreducible factors $g_i$ are linear. Since linear polynomials are not even functions of $n$ , they must occur in pairs $g_i(n) = a_in + b_i$ , $h_i(n) = a_in - b_i$ . Without loss of generality, $b_i \ge 0$ . Since the coefficients of $f$ are relatively prime, so are $a_i$ and $b_i$ , and since $P(0) = \infty$ , neither polynomial can have any non-negative integer roots, so $a_i > 1$ and thus $b_i > 0$ . On the other hand, by Dirichlet's theorem, $a_i \le 2$ , since otherwise the sequence $a_in + b_i$ would yield infinitely many prime values with $P(g_i(n)) = a_in + b_i \ge 3n.$ So $a_i = 2$ and therefore $b_i$ is a positive odd integer. Setting $b_i = 2c_i + 1$ , clearly $P(g_i(n)) - 2n < 2c_i + 2$ . Since this holds for each factor $g_i$ , it is true for the product $g$ of all the factors with the bound determined by the factor with the largest value of $c_i$ . Therefore, for suitable non-negative integers $c_i$ , $g(n)$ is a product of polynomials of the form $4n^2 - (2c_i + 1)^2$ . Now, since $g(n) = f(n^2)$ , we conclude that $f(n)$ is a product of linear factors of the form $4n - (2c_i + 1)^2$ . Since we restricted ourselves to non-constant polynomials with relatively prime coefficients, we can now relax this condition and admit a possibly empty list of linear factors as well as an arbitrary non-zero integer multiple $M$ . Thus for a suitable non-zero integer $M$ and $k \ge 0$ non-negative integers $c_i$ , we have: \[f(n) = M \cdot \prod_{i=1}^k (4n - (2c_i + 1)^2)\] Solution 2 The polynomial $f$ has the required properties if and only if \[f(x) = c(4x - a_1^2)(4x - a_2^2)\cdots (4x - a_k^2),\qquad\qquad (*)\] where $a_1, a_2, \ldots, a_k$ are odd positive integers and $c$ is a nonzero integer. It is straightforward to verify that polynomials given by $(*)$ have the required property. If $p$ is a prime divisor of $f(n^2)$ but not of $c$ , then $p|(2n - a_j)$ or $p|(2n + a_j)$ for some $j\leq k$ . Hence $p - 2n\leq \max\{a_1, a_2, \ldots, a_k\}$ . The prime divisors of $c$ form a finite set and do not affect whether or not the given sequence is bounded above. The rest of the proof is devoted to showing that any $f$ for which $\{p(f(n^2)) - 2n\}_{n\geq 0}$ is bounded above is given by $(*)$ . Let $\mathbb{Z}[x]$ denote the set of all polynomials with integer coefficients. Given $f\in\mathbb{Z}[x]$ , let $\mathcal{P}(f)$ denote the set of those primes that divide at least one of the numbers in the sequence $\{f(n)\}_{n\geq 0}$ . The solution is based on the following lemma. Lemma. If $f\in\mathbb{Z}[x]$ is a nonconstant polynomial then $\mathcal{P}(f)$ is infinite. Proof. Repeated use will be made of the following basic fact: if $a$ and $b$ are distinct integers and $f\in\mathbb{Z}[x]$ , then $a - b$ divides $f(a) - f(b)$ . If $f(0) = 0$ , then $p$ divides $f(p)$ for every prime $p$ , so $\mathcal{P}(f)$ is infinite. If $f(0) = 1$ , then every prime divisor $p$ of $f(n!)$ satisfies $p > n$ . Otherwise $p$ divides $n!$ , which in turn divides $f(n!) - f(0) = f(n!) - 1$ . This yields $p|1$ , which is false. Hence $f(0) = 1$ implies that $\mathcal{P}(f)$ is infinite. To complete the proof, set $g(x) = f(f(0)x)/f(0)$ and observe that $g\in\mathcal{Z}[x]$ and $g(0) = 1$ . The preceding argument shows that $\mathcal{P}(g)$ is infinite, and it follows that $\mathcal{P}(f)$ is infinite. $\blacksquare$ Suppose $f\in\mathbb{Z}[x]$ is nonconstant and there exists a number $M$ such that $p(f(n^2)) - 2n\leq M$ for all $n\geq 0$ . Application of the lemma to $f(x^2)$ shows that there is an infinite sequence of distinct primes $\{p_j\}$ and a corresponding infinite sequence of nonnegative integers $\{k_j\}$ such that $p_j|f(k_j)^2$ for all $j\geq 1$ . Consider the sequence $\{r_j\}$ where $r_j = \min\{k_j\pmod{p_j}, p_j - k_j\pmod{p_j}\}$ . Then $0\leq r_j\leq (p_j - 1)/2$ and $p_j|f(r_j)^2$ . Hence $2r_j + 1\leq p_j\leq p(f(r_j^2))\leq M + 2r_j$ , so $1\leq p_j - 2r_\leq M$ for all $j\geq 1$ . It follows that there is an integer $a_1$ such that $1\leq a_1\leq M$ and $a_1 = p_j - 2r_j$ for infinitely many $j$ . Let $m = \deg f$ . Then $p_j|4^mf(((p_j - a_1)/2)^2)$ and $4^mf(((x - a_1)/2)^2)\in\mathbb{Z}[x]$ . Consequently, $p_j|f((a_1/2)^2)$ for infinitely many $j$ , which shows that $(a_1/2)^2$ is a zero of $f$ . Since $f(n^2)\leq 0$ for $n\geq 0$ , $a_1$ must be odd. Then $f(x) = (4x - a_1)^2g(x)$ , where $g\in\mathbb{Z}[x]$ . (See the note below.) Observe that $\{p(g(n^2)) - 2n\}_{n\geq 0}$ must be bounded above. If $g$ is constant, we are done. If $g$ is nonconstant, the argument can be repeated to show that $f$ is given by $(*)$ . Note. The step that gives $f(x) = (4x - a_1^2)g(x)$ where $g\in\mathbb{Z}[x]$ follows immediately using a lemma of Gauss. The use of such an advanced result can be avoided by first writing $f(x) = r(4x - a_1^2)g(x)$ where $r$ is rational and $g\in\mathbb{Z}[x]$ . Then continuation gives $f(x) = c(4x - a_1^2)\cdots (4x - a_k^2)$ where $c$ is rational and the $a_i$ are odd. Consideration of the leading coefficient shows that the denominator of $c$ is $2^s$ for some $s\geq 0$ and consideration of the constant term shows that the denominator is odd. Hence $c$ is an integer. Alternate solutions are always welcome. If you have a different, elegant solution to this problem, please add it to this page.
\[ f(x) = c(4x - a_1^2)(4x - a_2^2)\cdots (4x - a_k^2), \] where \( a_1, a_2, \ldots, a_k \) are odd positive integers and \( c \) is a nonzero integer.
usamo
[ "Mathematics -> Algebra -> Intermediate Algebra -> Sequences -> Other" ]
5.25
A sequence of positive integers $a_{1}, a_{2}, \ldots, a_{2017}$ has the property that for all integers $m$ where $1 \leq m \leq 2017,3\left(\sum_{i=1}^{m} a_{i}\right)^{2}=\sum_{i=1}^{m} a_{i}^{3}$. Compute $a_{1337}$.
I claim that $a_{i}=3 i$ for all $i$. We can conjecture that the sequence should just be the positive multiples of three because the natural numbers satisfy the property that the square of their sum is the sum of their cubes, and prove this by induction. At $i=1$, we have that $3 a_{i}^{2}=a_{i}^{3}$, so $a_{i}=3$. Now assuming this holds for $i=m$, we see that $$\begin{aligned} 3\left(\sum_{i=1}^{m+1} a_{i}\right)^{2} & =3\left(a_{m+1}+\sum_{i=1}^{m} a_{i}\right)^{2} \\ & =3 a_{m+1}^{2}+\sum_{i=1}^{m} a_{i}^{3}+6 a_{m+1} \sum_{i=1}^{m} a_{i} \\ & =3 a_{m+1}^{2}+\sum_{i=1}^{m} a_{i}^{3}+6 a_{m+1} \cdot 3\left(\frac{m(m+1)}{2}\right) \\ & =\sum_{i=1}^{m+1} a_{i}^{3} \end{aligned}$$ Therefore, $$\begin{aligned} a_{m+1}^{3} & =3 a_{m+1}^{2}+a_{m+1}\left(9 m^{2}+9 m\right) \\ 0 & =a_{m+1}^{2}-3 a_{m+1}-\left(9 m^{2}+9 m\right) \\ 0 & =\left(a_{m+1}-(3 m+3)\right)\left(a_{m+1}+3 m\right) \end{aligned}$$ and because the sequence is positive, $a_{m+1}=3 m+3$, which completes the induction. Then $a_{1337}=$ $1337 \cdot 3=4011$
\[ a_{1337} = 4011 \]
HMMT_11
[ "Mathematics -> Discrete Mathematics -> Graph Theory" ]
8
A graph $G(V,E)$ is triangle-free, but adding any edges to the graph will form a triangle. It's given that $|V|=2019$, $|E|>2018$, find the minimum of $|E|$ .
Given a graph \( G(V, E) \) that is triangle-free, but adding any edges to the graph will form a triangle, and with \( |V| = 2019 \) and \( |E| > 2018 \), we need to find the minimum number of edges \( |E| \). We claim that the minimum number of edges is \( 2n - 5 \) where \( n = 2019 \). This bound is attained for a graph constructed as follows: take a 5-cycle \( C_5 \), and replace one of the vertices with an independent set of \( n - 4 \) vertices, each of which is adjacent to the two neighbors of the original vertex. To prove this, consider the following: 1. **Diameter Condition**: The graph \( G \) has diameter 2 because any two vertices with distance greater than 2 could have an edge added between them without forming a triangle. A diameter 1 graph is complete, which is not our case. 2. **Minimum Degree Analysis**: - If \( d \geq 4 \), then \( G \) has at least \( \frac{4n}{2} > 2n - 5 \) edges. - If \( d = 1 \), let \( v \) be a vertex connected only to \( w \). Then every other vertex must be connected to \( w \), making \( G \) a star graph, which contradicts \( |E| > n - 1 \). - If \( d = 2 \), let \( v \) be connected to \( w \) and \( x \). By the diameter 2 condition, every other vertex is connected to \( w \), \( x \), or both. Let \( A \) be the set of vertices adjacent to \( w \) but not \( x \), \( B \) be the set adjacent to both \( w \) and \( x \), and \( C \) be the set adjacent to \( x \) but not \( w \). Then \( |A| + |B| + |C| = n - 2 \). The only edges we can add are between \( A \) and \( C \), ensuring \( |E| \geq 2n - 5 \). - If \( d = 3 \), let \( v \) be adjacent to \( w \), \( x \), and \( y \). Each vertex in \( S = V \setminus \{v, w, x, y\} \) is adjacent to one of \( w \), \( x \), or \( y \). The degree sum gives \( \deg(w) + \deg(x) + \deg(y) \geq n - 1 \), leading to \( |E| \geq 2n - 5 \). Thus, the minimum number of edges \( |E| \) in such a graph is: \[ |E| = 2 \cdot 2019 - 5 = 4033. \] The answer is: \boxed{4033}.
4033
china_team_selection_test
[ "Mathematics -> Precalculus -> Functions", "Mathematics -> Algebra -> Algebra -> Equations and Inequalities" ]
7
Let $\mathbb{R}$ be the set of real numbers . Determine all functions $f : \mathbb{R} \rightarrow \mathbb{R}$ such that for all pairs of real numbers $x$ and $y$ .
Solution 1 We first prove that $f$ is odd . Note that $f(0) = f(x^2 - x^2) = xf(x) - xf(x) = 0$ , and for nonzero $y$ , $xf(x) + yf(-y) = f(x^2 - y^2) = xf(x) - yf(y)$ , or $yf(-y) = -yf(y)$ , which implies $f(-y) = -f(y)$ . Therefore $f$ is odd. Henceforth, we shall assume that all variables are non-negative. If we let $y = 0$ , then we obtain $f(x^2) = xf(x)$ . Therefore the problem's condition becomes . But for any $a,b$ , we may set $x = \sqrt{a}$ , $y = \sqrt{b}$ to obtain . (It is well known that the only continuous solutions to this functional equation are of the form $f(x) = kx$ , but there do exist other solutions to this which are not solutions to the equation of this problem.) We may let $a = 2t$ , $b = t$ to obtain $2f(t) = f(2t)$ . Letting $x = t+1$ and $y = t$ in the original condition yields But we know $f(2t + 1) = f(2t) + f(1) = 2f(t) + f(1)$ , so we have $2f(t) + f(1) = f(t) + tf(1) + f(1)$ , or . Hence all solutions to our equation are of the form $f(x) = kx$ . It is easy to see that real value of $k$ will suffice. Solution 2 As in the first solution, we obtain the result that $f$ satisfies the condition . We note that . Since $f(2t) = 2f(t)$ , this is equal to It follows that $f$ must be of the form $f(x) = kx$ . Alternate solutions are always welcome. If you have a different, elegant solution to this problem, please add it to this page.
The function \( f(x) = kx \) for any real value of \( k \).
usamo
[ "Mathematics -> Algebra -> Abstract Algebra -> Group Theory", "Mathematics -> Number Theory -> Other", "Mathematics -> Discrete Mathematics -> Combinatorics" ]
8
Determine (with proof) whether there is a subset $X$ of the integers with the following property: for any integer $n$ there is exactly one solution of $a + 2b = n$ with $a,b \in X$ .
Start with an incomplete subset $S = (S_1, S_2, S_3, ... S_m)$ , such that for any integer n, there is exactly zero or one solutions to $a + 2b = n$ with $a,b \in S$ . Let $N$ be the smallest integer such that for any $S_i$ , $|S_i| < N$ . Note that $|S_i+2S_j| < 3N$ for any $S_i$ and $S_j$ Suppose $M$ is the smallest non-negative integer without a solution in $S$ yet. Clearly, $0 \le M \le 3N$ . Generate $S_{m+1}$ and $S_{m+2}$ such that $S_{m+1} = -10N - M$ , and $S_{m+2} = 5N + M$ . Thus, we now have the solution $S_{m+1}+2S_{m+2} = M$ . Note: The values 10 and 5 can be replaced by any sufficiently large values such that the first is twice the second. Now, we must prove that the addition of these two terms to $S$ does not result in an integer n that has two solutions. Of course, $S_{m+1} + 2S_{m+2} = M$ which previously had no solutions. Furthermore, $S_{m+1} + 2S_{m+2} = -15N - M$ . For any , , and . Since , we get that and Similarly, , and . Since , we get that and . Since all of these sums (other than $M$ ) are either greater than $3N$ or less than $-3N$ , they are all sums that previously had no solutions. Furthermore, none of these sums are duplicated, as sums of different forms are contained in disjoint ranges of integers. Thus, we have proved that we can generate a subset $S$ such that all non-negative integers n have a unique solution $a + 2b = n$ . For negative integers M that have no solutions in $S$ a similar proof holds, but instead generating the terms $S_{m+1} = 10N - M$ and $S_{m+2} = -5N + M$ . For any integer M that currently has no solution in S, we can always add two terms $S_{m+1}$ and $S_{m+2}$ such that $S_{m+1} + 2S_{m+2} = M$ that do not result in duplicated sums. Thus, there does exist a subset $X$ of the integers such that for any integer $n$ there is exactly one solution to $a + 2b = n$ with $a, b \in X$ .
Yes, there does exist a subset \( X \) of the integers such that for any integer \( n \) there is exactly one solution to \( a + 2b = n \) with \( a, b \in X \).
usamo
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations", "Mathematics -> Algebra -> Algebra -> Equations and Inequalities" ]
6
If $P(x)$ denotes a polynomial of degree $n$ such that \[P(k)=\frac{k}{k+1}\] for $k=0,1,2,\ldots,n$ , determine $P(n+1)$ .
Let $Q(x) = (x+1)P(x) - x$ , and clearly, $Q(x)$ has a degree of $n+1$ . Then, for $k=0,1,2,\ldots,n$ , $Q(k) = (k+1)P(k) - k = (k+1)\cdot \dfrac{k}{k+1} - k = 0$ . Thus, $k=0,1,2,\ldots,n$ are the roots of $Q(x)$ . Since these are all $n+1$ of the roots of the $n+1^{\text{th}}$ degree polynomial, by the Factor Theorem , we can write $Q(x)$ as \[Q(x) = c(x)(x-1)(x-2) \cdots (x-n)\] where $c$ is a constant. Thus, \[(x+1)P(x) - x = c(x)(x-1)(x-2) \cdots (x-n).\] We plug in $x = -1$ to cancel the $(x+1)P(x)$ and find $c$ : \begin{align*} -(-1) &= c(-1)(-1-1)(-1-2) \cdots (-1-n) \\ 1 &= c(-1)^{n+1}(1)(2) \cdots (n+1) \\ c &= (-1)^{n+1}\dfrac{1}{(n+1)!} \\ \end{align*} Finally, plugging in $x = n+1$ to find $P(n+1)$ gives: \begin{align*} Q(n+1)&=(n+2)P(n+1)-(n+1)\\ (-1)^{n+1}\dfrac{1}{(n+1)!}\cdot(n+1)! &=(n+2)P(n+1)-(n+1)\\ (-1)^{n+1}&=(n+2)P(n+1)-(n+1)\\ (-1)^{n+1}+(n+1)&=(n+2)P(n+1)\\ P(n+1) &= \dfrac{(-1)^{n+1} + (n+1)}{n+2}\\ \end{align*} If $n$ is even, this simplifies to $P(n+1) = \dfrac{n}{n+2}$ . If $n$ is odd, this simplifies to $P(n+1) = 1$ . $\Box$ ~Edits by BakedPotato66
\[ P(n+1) = \begin{cases} \dfrac{n}{n+2} & \text{if } n \text{ is even} \\ 1 & \text{if } n \text{ is odd} \end{cases} \]
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Polygons", "Mathematics -> Algebra -> Algebra -> Equations and Inequalities" ]
8
A convex hexagon $A B C D E F$ is inscribed in a circle. Prove the inequality $A C \cdot B D \cdot C E \cdot D F \cdot A E \cdot B F \geq 27 A B \cdot B C \cdot C D \cdot D E \cdot E F \cdot F A$.
Let $d_{1}=A B \cdot B C \cdot C D \cdot D E \cdot E F \cdot F A, d_{2}=A C \cdot B D \cdot C E \cdot D F \cdot A E \cdot B F, d_{3}=A D \cdot B E \cdot C F$. Applying Ptolemy's theorem to quadrilaterals $A B C D, B C D E, C D E F, D E F A, E F A B, F A B C$, we obtain six equations $A C \cdot B D-A B \cdot C D=B C \cdot A D, \ldots, F B \cdot A C-F A \cdot B C=A B \cdot F C$. Putting these equations in the well-known inequality $$\sqrt[6]{\left(a_{1}-b_{1}\right)\left(a_{2}-b_{2}\right) \cdot \ldots \cdot\left(a_{6}-b_{6}\right)} \leq \sqrt[6]{a_{1} a_{2} \ldots a_{6}}-\sqrt[6]{b_{1} b_{2} \ldots b_{6}} \quad\left(a_{i} \geq b_{i}>0, i=1, \ldots, 6\right)$$ we get $$\sqrt[3]{d_{3}} \sqrt[6]{d_{1}} \leq \sqrt[3]{d_{2}}-\sqrt[3]{d_{1}} \tag{1}$$ Applying Ptolemy's theorem to quadrilaterals $A C D F, A B D E$ и $B C E F$, we obtain three equations $A D \cdot C F=A C \cdot D F+A F \cdot C D, A D \cdot B E=B D \cdot A E+A B \cdot D E, B E \cdot C F=B F \cdot C E+B C \cdot E F$. Putting these equations in the well-known inequality $$\sqrt[3]{\left(a_{1}+b_{1}\right)\left(a_{2}+b_{2}\right)\left(a_{3}+b_{3}\right)} \geq \sqrt[3]{a_{1} a_{2} a_{3}}+\sqrt[3]{b_{1} b_{2} b_{3}}\left(a_{i}>0, b_{i}>0, i=1,2,3\right)$$ we get $$\sqrt[3]{d_{3}^{2}} \geq \sqrt[3]{d_{2}}+\sqrt[3]{d_{1}} \tag{2}$$ It follows from (1) and (2) that $(\sqrt[3]{d_{2}}-\sqrt[3]{d_{1}})^{2} \geq \sqrt[3]{d_{1}}(\sqrt[3]{d_{2}}+\sqrt[3]{d_{1}})$, that is, $\sqrt[3]{d_{2}} \geq 3 \sqrt[3]{d_{1}}$ and $d_{2} \geq 27 d_{1}$, q.e.d.
\[ A C \cdot B D \cdot C E \cdot D F \cdot A E \cdot B F \geq 27 A B \cdot B C \cdot C D \cdot D E \cdot E F \cdot F A \]
izho
[ "Mathematics -> Geometry -> Plane Geometry -> Angles", "Mathematics -> Geometry -> Plane Geometry -> Triangulations" ]
6
$P$ lies between the rays $OA$ and $OB$ . Find $Q$ on $OA$ and $R$ on $OB$ collinear with $P$ so that $\frac{1}{PQ} + \frac{1}{PR}$ is as large as possible.
Perform the inversion with center $P$ and radius $\overline{PO}.$ Lines $OA,OB$ go to the circles $(O_1),(O_2)$ passing through $P,O$ and the line $QR$ cuts $(O_1),(O_2)$ again at the inverses $Q',R'$ of $Q,R.$ Hence $\frac{1}{PQ}+\frac{1}{PR}=\frac{PQ'+PR'}{PO^2}=\frac{Q'R'}{PO^2}$ Thus, it suffices to find the line through $P$ that maximizes the length of the segment $\overline{Q'R'}.$ If $M,N$ are the midpoints of $PQ',PR',$ i.e. the projections of $O_1,O_2$ onto $QR,$ then from the right trapezoid $O_1O_2NM,$ we deduce that $O_1O_2 \ge MN = \frac{_1}{^2}Q'R'.$ Consequently, $2 \cdot O_1O_2$ is the greatest possible length of $Q'R',$ which obviously occurs when $O_1O_2NM$ is a rectangle. Hence, $Q,R$ are the intersections of $OA,OB$ with the perpendicular to $PO$ at $P.$
The intersections of \(OA\) and \(OB\) with the perpendicular to \(PO\) at \(P\).
usamo
[ "Mathematics -> Discrete Mathematics -> Combinatorics", "Mathematics -> Algebra -> Abstract Algebra -> Group Theory (related to permutations) -> Other" ]
4.5
Let $\pi$ be a permutation of $\{1,2, \ldots, 2015\}$. With proof, determine the maximum possible number of ordered pairs $(i, j) \in\{1,2, \ldots, 2015\}^{2}$ with $i<j$ such that $\pi(i) \cdot \pi(j)>i \cdot j$.
Let $n=2015$. The only information we will need about $n$ is that $n>5 \sqrt[4]{4}$. For the construction, take $\pi$ to be the $n$-cycle defined by $\pi(k)= \begin{cases}k+1 & \text { if } 1 \leq k \leq n-1 \\ 1 & \text { if } k=n\end{cases}$. Then $\pi(i)>i$ for $1 \leq i \leq n-1$. So $\pi(i) \pi(j)>i j$ for at least $\binom{n-1}{2}$ pairs $i<j$. For convenience let $z_{i}=\frac{\pi(i)}{i}$, so that we are trying to maximize the number of pairs $(i, j), i<j$ with $z_{i} z_{j}>1$. Notice that over any cycle $c=\left(i_{1} i_{2} \cdots i_{k}\right)$ in the cycle decomposition of $\pi$ we have $\prod_{i \in c} z_{i}=1$. In particular, multiplying over all such cycles gives $\prod_{i=1}^{n} z_{i}=1$. Construct a graph $G$ on vertex set $V=[n]$ such that there is an edge between $i$ and $j$ whenever $\pi(i) \pi(j)>i j$. For any cycle $C=\left(v_{1}, v_{2}, \ldots, v_{k}\right)$ in this graph we get $1<\prod_{i=1}^{k} z_{v_{i}} z_{v_{i+1}}=\prod_{v \in C} z_{v}^{2}$. So, we get in particular that $G$ is non-Hamiltonian. By the contrapositive of Ore's theorem there are two distinct indices $u, v \in[n]$ such that $d(u)+d(v) \leq n-1$. The number of edges is then at most $\binom{n-2}{2}+(n-1)=\binom{n-1}{2}+1$. If equality is to hold, we need $G \backslash\{u, v\}$ to be a complete graph. We also need $d(u)+d(v)=n-1$, with $u v$ not an edge. This implies $d(u), d(v) \geq 1$. Since $n>5$, the pigeonhole principle gives that at least one of $u, v$ has degree at least 3. WLOG $d(u) \geq 3$. Let $w$ be a neighbor of $v$ and let $a, b, c$ be neighbors of $u$; WLOG $w \neq a, b$. Since $G \backslash\{u, v, w\}$ is a complete graph, we can pick a Hamiltonian path in $G \backslash\{u, v, w\}$ with endpoints $a, b$. Connecting $u$ to the ends of this path forms an $(n-2)$-cycle $C$. This gives us $\prod_{x \in C} z_{x}^{2}>1$. But we also have $z_{v} z_{w}>1$, so $1=\prod_{i=i}^{n} z_{i}>1$, contradiction. So, $\binom{n-1}{2}+1$ cannot be attained, and $\binom{n-1}{2}$ is indeed the maximum number of pairs possible.
\[ \binom{2014}{2} \]
HMMT_2
[ "Mathematics -> Algebra -> Abstract Algebra -> Other" ]
9
Let $n\geq 2$ be a given integer. Find all functions $f:\mathbb{R}\rightarrow \mathbb{R}$ such that \[f(x-f(y))=f(x+y^n)+f(f(y)+y^n), \qquad \forall x,y \in \mathbb R.\]
Let \( n \geq 2 \) be a given integer. We aim to find all functions \( f: \mathbb{R} \rightarrow \mathbb{R} \) such that \[ f(x - f(y)) = f(x + y^n) + f(f(y) + y^n), \quad \forall x, y \in \mathbb{R}. \] The solutions to this functional equation are: 1. \( f(x) = 0 \) for all \( x \in \mathbb{R} \). 2. \( f(x) = -x^n \) for all \( x \in \mathbb{R} \). To verify, we check both functions: 1. For \( f(x) = 0 \): \[ f(x - f(y)) = f(x - 0) = 0, \] \[ f(x + y^n) + f(f(y) + y^n) = 0 + 0 = 0, \] which satisfies the given equation. 2. For \( f(x) = -x^n \): \[ f(x - f(y)) = f(x + y^n) = -(x + y^n)^n, \] \[ f(x + y^n) + f(f(y) + y^n) = - (x + y^n)^n + (-(-y^n + y^n)^n) = -(x + y^n)^n, \] which also satisfies the given equation. Thus, the only solutions are \( f(x) = 0 \) and \( f(x) = -x^n \). The answer is: \boxed{f(x) = 0 \text{ or } f(x) = -x^n}.
f(x) = 0 \text{ or } f(x) = -x^n
china_team_selection_test
[ "Mathematics -> Discrete Mathematics -> Combinatorics", "Mathematics -> Algebra -> Algebra -> Algebraic Expressions" ]
4.5
For each positive integer $n$ , find the number of $n$ -digit positive integers that satisfy both of the following conditions: $\bullet$ no two consecutive digits are equal, and $\bullet$ the last digit is a prime.
The answer is $\boxed{\frac{2}{5}\left(9^n+(-1)^{n+1}\right)}$ . Suppose $a_n$ denotes the number of $n$ -digit numbers that satisfy the condition. We claim $a_n=4\cdot 9^{n-1}-a_{n-1}$ , with $a_1=4$ . $\textit{Proof.}$ It is trivial to show that $a_1=4$ . Now, we can do casework on whether or not the tens digit of the $n$ -digit integer is prime. If the tens digit is prime, we can choose the digits before the units digit in $a_{n-1}$ ways and choose the units digit in $3$ ways, since it must be prime and not equal to the tens digit. Therefore, there are $3a_{n-1}$ ways in this case. If the tens digit is not prime, we can use complementary counting. First, we consider the number of $(n-1)$ -digit integers that do not have consecutive digits. There are $9$ ways to choose the first digit and $9$ ways to choose the remaining digits. Thus, there are $9^{n-1}$ integers that satisfy this. Therefore, the number of those $(n-1)$ -digit integers whose units digit is not prime is $9^{n-1}-a_{n-1}$ . It is easy to see that there are $4$ ways to choose the units digit, so there are $4\left(9^{n-1}-a_{n-1}\right)$ numbers in this case. It follows that \[a_n=3a_{n-1}+4\left(9^{n-1}-a_{n-1}\right)=4\cdot 9^{n-1}-a_{n-1},\] and our claim has been proven. Then, we can use induction to show that $a_n=\frac{2}{5}\left(9^n+(-1)^{n+1}\right)$ . It is easy to see that our base case is true, as $a_1=4$ . Then, \[a_{n+1}=4\cdot 9^n-a_n=4\cdot 9^n-\frac{2}{5}\left(9^n+(-1)^{n+1}\right)=4\cdot 9^n-\frac{2}{5}\cdot 9^n-\frac{2}{5}(-1)^{n+1},\] which is equal to \[a_{n+1}=\left(4-\frac{2}{5}\right)\cdot 9^n-\frac{2}{5}\cdot\frac{(-1)^{n+2}}{(-1)}=\frac{18}{5}\cdot 9^n+\frac{2}{5}\cdot(-1)^{n+2}=\frac{2}{5}\left(9^{n+1}+(-1)^{n+2}\right),\] as desired. $\square$ Solution by TheUltimate123.
\[ \frac{2}{5}\left(9^n + (-1)^{n+1}\right) \]
usajmo
[ "Mathematics -> Geometry -> Plane Geometry -> Polygons", "Mathematics -> Algebra -> Intermediate Algebra -> Inequalities" ]
7
The area of a convex pentagon $A B C D E$ is $S$, and the circumradii of the triangles $A B C, B C D, C D E, D E A, E A B$ are $R_{1}, R_{2}, R_{3}, R_{4}, R_{5}$. Prove the inequality $R_{1}^{4}+R_{2}^{4}+R_{3}^{4}+R_{4}^{4}+R_{5}^{4} \geqslant \frac{4}{5 \sin ^{2} 108^{\circ}} S^{2}$.
First we prove the following Lemma 1. In a convex $n$-gon $A_{1} A_{2} \ldots A_{n}$ with area $S$ we have $4 S \leqslant A_{n} A_{2} \cdot R_{1}+A_{1} A_{3} \cdot R_{2}+\ldots+A_{n-1} A_{1} \cdot R_{n}$ where $R_{i}$ is the circumradius of the triangle $A_{i-1} A_{i} A_{i+1}, A_{0}=A_{n}, A_{n+1}=A_{n}$. Let $M_{i}$ be the midpoint of $A_{i} A_{i+1}$ for $i=1, \ldots, n$. For each $i$ we consider the quadrilateral formed by the segments $A_{i} M_{i}$ and $A_{i} M_{i-1}$ and the perpendiculars to this segments drawn through $M_{i}$ and $M_{i-1}$, respectively. We claim that these $n$ quadrilateral cover the $n$-gon. Indeed, let $P$ be a point inside the $n$-gon. Let $P A_{k}$ be the minimum among the distances $P A_{1}, P A_{2}, \ldots, P A_{n}$. We have $P A_{k} \leqslant P A_{k+1}$ and $P A_{k} \leqslant P A_{k-11}$, therefore $P$ belongs to the $n$-gon and to each of the two half-planes containing $A_{k}$ and bounded by the perpendicular bisectors to $A_{k} A_{k+1}$ and $A_{k} A_{k-11}$, that is, to the $k$-th quadrilateral. To complete the proof it remains to note that the area of the $i$-th quadrilateral does nor exceed $\frac{1}{2} \cdot \frac{A_{i-1} A_{i+1}}{2} \cdot R_{i}$. For our problem it follows that $4 S \leqslant 2 R_{1}^{2} \sin \angle A_{1}+2 R_{2}^{2} \sin \angle A_{2}+\ldots+2 R_{5}^{2} \sin \angle A_{5}$. Applying Cauchy-Buniakowsky inequality, we obtain $2 S \leqslant R_{1}^{2} \sin \angle A_{1}+R_{2}^{2} \sin \angle A_{2}+\ldots+R_{5}^{2} \sin \angle A_{5} \leqslant \sqrt{\left(R_{1}^{4}+\ldots+R_{5}^{4}\right)\left(\sin ^{2} \angle A_{1}+\ldots+\sin ^{2} \angle A_{5}\right)} \leqslant \sqrt{5\left(R_{1}^{4}+\ldots+R_{5}^{4}\right) \sin ^{2} 108^{\circ}}$ thus $\frac{4 S^{2}}{5 \sin ^{2} 108^{\circ}} \leqslant R_{1}^{4}+R_{2}^{4}+\ldots+R_{5}^{4}$. In the above inequality we made use of the following Lemma 2. If $\alpha_{1}, \alpha_{2}, \ldots, \alpha_{5}$ are angles of a convex pentagon, then $\sin ^{2} \alpha_{1}+\ldots+\sin ^{2} \alpha_{5} \leqslant 5 \sin ^{2} 108^{\circ}$. The sum in question does not depend on the order of the angles, therefore we may assume $\alpha_{1} \leqslant \alpha_{2} \leqslant \ldots \leqslant \alpha_{5}$. If $\alpha_{1}=108^{\circ}$, then $\alpha_{2}=\ldots=\alpha_{5}=108^{\circ}$, and the inequality turns to equality. If $\alpha_{1}<108^{\circ}$, then $\alpha_{5}>108^{\circ}$. Note that $\alpha_{1}+\alpha_{5}<270^{\circ}$ (if $\alpha_{1}+\alpha_{5} \geqslant 270^{\circ}$, then $\alpha_{2}+\alpha_{3}+\alpha_{4} \leqslant 270^{\circ}$, therefore $\alpha_{2} \leqslant 90^{\circ}$, a fortiori $\alpha_{1} \leqslant 90^{\circ}$ and thus $\alpha_{5} \geqslant 180^{\circ}$, a contradiction). Then we have $\sin ^{2} 108^{\circ}+\sin ^{2}\left(\alpha_{1}+\alpha_{5}-108^{\circ}\right)-\sin ^{2} \alpha_{1}-\sin ^{2} \alpha_{5}=2 \cos \left(\alpha_{1}+\alpha_{5}\right) \sin \left(\alpha_{1}-108^{\circ}\right) \sin \left(\alpha_{5}-108^{\circ}\right)>0$. It means that changing the angles $\alpha_{1}$ by $108^{\circ}$ and $\alpha_{5}$ by $\alpha_{1}+\alpha_{5}-108^{\circ}$ increases the sum of squares of the sines. Iterating this operation, we shall make all the angles equal to $108^{\circ}$, thus proving the inequality.
\[ R_{1}^{4} + R_{2}^{4} + R_{3}^{4} + R_{4}^{4} + R_{5}^{4} \geq \frac{4}{5 \sin^{2} 108^{\circ}} S^{2} \]
izho
[ "Mathematics -> Geometry -> Plane Geometry -> Circles", "Mathematics -> Geometry -> Plane Geometry -> Triangulations" ]
6
Let triangle \(ABC\) be an acute triangle with circumcircle \(\Gamma\). Let \(X\) and \(Y\) be the midpoints of minor arcs \(\widehat{AB}\) and \(\widehat{AC}\) of \(\Gamma\), respectively. If line \(XY\) is tangent to the incircle of triangle \(ABC\) and the radius of \(\Gamma\) is \(R\), find, with proof, the value of \(XY\) in terms of \(R\).
Note that \(X\) and \(Y\) are the centers of circles \((AIB)\) and \((AIC)\), respectively, so we have \(XY\) perpendicularly bisects \(AI\), where \(I\) is the incenter. Since \(XY\) is tangent to the incircle, we have \(AI\) has length twice the inradius. Thus, we get \(\angle A=60^{\circ}\). Thus, since \(\widehat{XY}=\frac{\widehat{BAC}}{2}\), we have \(\widehat{XY}\) is a \(120^{\circ}\) arc. Thus, we have \(XY=R \sqrt{3}\).
\[ XY = R \sqrt{3} \]
HMMT_2
[ "Mathematics -> Geometry -> Plane Geometry -> Angles", "Mathematics -> Geometry -> Plane Geometry -> Triangulations" ]
7.5
Let $ABC$ be a triangle. Find all points $P$ on segment $BC$ satisfying the following property: If $X$ and $Y$ are the intersections of line $PA$ with the common external tangent lines of the circumcircles of triangles $PAB$ and $PAC$ , then \[\left(\frac{PA}{XY}\right)^2+\frac{PB\cdot PC}{AB\cdot AC}=1.\]
Let circle $PAB$ (i.e. the circumcircle of $PAB$ ), $PAC$ be $\omega_1, \omega_2$ with radii $r_1$ , $r_2$ and centers $O_1, O_2$ , respectively, and $d$ be the distance between their centers. Lemma. $XY = \frac{r_1 + r_2}{d} \sqrt{d^2 - (r_1 - r_2)^2}.$ Proof. Let the external tangent containing $X$ meet $\omega_1$ at $X_1$ and $\omega_2$ at $X_2$ , and let the external tangent containing $Y$ meet $\omega_1$ at $Y_1$ and $\omega_2$ at $Y_2$ . Then clearly $X_1 Y_1$ and $X_2 Y_2$ are parallel (for they are both perpendicular $O_1 O_2$ ), and so $X_1 Y_1 Y_2 X_2$ is a trapezoid. Now, $X_1 X^2 = XA \cdot XP = X_2 X^2$ by Power of a Point, and so $X$ is the midpoint of $X_1 X_2$ . Similarly, $Y$ is the midpoint of $Y_1 Y_2$ . Hence, $XY = \frac{1}{2} (X_1 Y_1 + X_2 Y_2).$ Let $X_1 Y_1$ , $X_2 Y_2$ meet $O_1 O_2$ s at $Z_1, Z_2$ , respectively. Then by similar triangles and the Pythagorean Theorem we deduce that $X_1 Z_1 = \frac{r_1 \sqrt{d^2 - (r_1 - r_2)^2}}{d}$ and $\frac{r_2 \sqrt{d^2 - (r_1 - r_2)^2}}{d}$ . But it is clear that $Z_1$ , $Z_2$ is the midpoint of $X_1 Y_1$ , $X_2 Y_2$ , respectively, so $XY = \frac{(r_1 + r_2)}{d} \sqrt{d^2 - (r_1 - r_2)^2},$ as desired. Lemma 2. Triangles $O_1 A O_2$ and $BAC$ are similar. Proof. $\angle{AO_1 O_2} = \frac{\angle{PO_1 A}}{2} = \angle{ABC}$ and similarly $\angle{AO_2 O_1} = \angle{ACB}$ , so the triangles are similar by AA Similarity. Also, let $O_1 O_2$ intersect $AP$ at $Z$ . Then obviously $Z$ is the midpoint of $AP$ and $AZ$ is an altitude of triangle $A O_1 O_2$ .Thus, we can simplify our expression of $XY$ : \[XY = \frac{AB + AC}{BC} \cdot \frac{AP}{2 h_a} \sqrt{BC^2 - (AB - AC)^2},\] where $h_a$ is the length of the altitude from $A$ in triangle $ABC$ . Hence, substituting into our condition and using $AB = c, BC = a, CA = b$ gives \[\left( \frac{2a h_a}{(b+c) \sqrt{a^2 - (b-c)^2}} \right)^2 + \frac{PB \cdot PC}{bc} = 1.\] Using $2 a h_a = 4[ABC] = \sqrt{(a + b + c)(a + b - c)(a - b + c)(-a + b + c)}$ by Heron's Formula (where $[ABC]$ is the area of triangle $ABC$ , our condition becomes \[\frac{(a + b + c)(-a + b + c)}{(b + c)^2} + \frac{PB \cdot PC}{bc} = 1,\] which by $(a + b + c)(-a + b + c) = (b + c)^2 - a^2$ becomes \[\frac{PB \cdot PC}{bc} = \frac{a^2 bc}{(b+c)^2}.\] Let $PB = x$ ; then $PC = a - x$ . The quadratic in $x$ is \[x^2 - ax + \frac{a^2 bc}{(b+c)^2} = 0,\] which factors as \[\left(x - \frac{ab}{b+c}\right)\left(x - \frac{ac}{b+c}\right) = 0.\] Hence, $PB = \frac{ab}{b+c}$ or $\frac{ac}{b+c}$ , and so the $P$ corresponding to these lengths are our answer. The problems on this page are copyrighted by the Mathematical Association of America 's American Mathematics Competitions .
The points \( P \) on segment \( BC \) that satisfy the given property are such that: \[ PB = \frac{ab}{b+c} \quad \text{or} \quad PB = \frac{ac}{b+c}. \]
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Triangulations" ]
5.25
Let $A B C$ be a triangle. The following diagram contains points $P_{1}, P_{2}, \ldots, P_{7}$, which are the following triangle centers of triangle $A B C$ in some order: - the incenter $I$; - the circumcenter $O$; - the orthocenter $H$; - the symmedian point $L$, which is the intersections of the reflections of $B$-median and $C$-median across angle bisectors of $\angle A B C$ and $\angle A C B$, respectively; - the Gergonne point $G$, which is the intersection of lines from $B$ and $C$ to the tangency points of the incircle with $\overline{A C}$ and $\overline{A B}$, respectively; - the Nagel point $N$, which is the intersection of line from $B$ to the tangency point between $B$ excircle and $\overline{A C}$, and line from $C$ to the tangency point between $C$-excircle and $\overline{A B}$; and - the Kosnita point $K$, which is the intersection of lines from $B$ and $C$ to the circumcenters of triangles $A O C$ and $A O B$, respectively. Compute which triangle centers $\{I, O, H, L, G, N, K\}$ corresponds to $P_{k}$ for $k \in\{1,2,3,4,5,6,7\}$.
Let $G^{\prime}$ be the centroid of triangle $A B C$. Recall the following. - Points $O, G^{\prime}, H$ lie on Euler's line of $\triangle A B C$ with $O G^{\prime}: G^{\prime} H=1: 2$. - Points $I, G^{\prime}, N$ lie on Nagel's line of $\triangle A B C$ with $I G^{\prime}: G^{\prime} N=1: 2$. Thus, $O I \parallel H N$ with $O I: H N=1: 2$. Therefore, we can detect parallel lines with ratio $2: 1$ in the figure. The only possible pairs are $P_{2} P_{4} \parallel P_{7} P_{5}$. Therefore, there are two possibilities: $\left(P_{2}, P_{7}\right)$ and $\left(P_{4}, P_{5}\right)$ must be $(O, H)$ and $(I, N)$ in some order. Intuitively, $H$ should be further out, so it's not unreasonable to guess that $P_{2}=O, P_{7}=H, P_{4}=I$, and $P_{5}=N$. Alternatively, perform the algorithm below with the other case to see if it fails. To identify the remaining points, we recall that the isogonal conjugate of $G$ and $N$ both lie on $O I$ (they are insimilicenter and exsimilicenter of incircle and circumcircle, respectively). Thus, $H, G, N, I$ lie on isogonal conjugate of $O I$, known as the Feuerbach's Hyperbola. It's also known that $O I$ is tangent to this line, and this hyperbola have perpendicular asymptotes. Using all information in the above paragraph, we can eyeball a rectangular hyperbola passing through $H, G, N, I$ and is tangent to $O I$. It's then not hard to see that $P_{6}=G$. Finally, we need to distinguish between symmedian and Kosnita points. To do that, recall that Kosnita point is isogonal conjugate of the nine-point center (not hard to show). Thus, $H, L, K, O$ lies on isogonal conjugate of $O H$, which is the Jerabek's Hyperbola. One can see that $H, L, K, O$ lies on the same branch. Moreover, they lie on this hyperbola in this order because the isogonal conjugates (in order) are $O$, centroid, nine-point center, and $H$, which lies on $O H$ in this order. Using this fact, we can identity $P_{5}=L$ and $P_{1}=K$, completing the identification.
\[ \begin{aligned} P_1 &= K, \\ P_2 &= O, \\ P_3 &= \text{(not specified)}, \\ P_4 &= I, \\ P_5 &= L, \\ P_6 &= G, \\ P_7 &= H. \end{aligned} \]
HMMT_2
[ "Mathematics -> Discrete Mathematics -> Combinatorics", "Mathematics -> Discrete Mathematics -> Combinatorics" ]
8
Let $n \ge 3$ be an integer. Rowan and Colin play a game on an $n \times n$ grid of squares, where each square is colored either red or blue. Rowan is allowed to permute the rows of the grid, and Colin is allowed to permute the columns of the grid. A grid coloring is $orderly$ if: no matter how Rowan permutes the rows of the coloring, Colin can then permute the columns to restore the original grid coloring; and no matter how Colin permutes the column of the coloring, Rowan can then permute the rows to restore the original grid coloring; In terms of $n$ , how many orderly colorings are there?
We focus on the leftmost column for simplicity. Let $m$ be the number of red squares in this column. We then have five cases: 1. $m=1$ When Rowan permutes the rows of the coloring, we consider only the first column, which by the above contains $m=1$ red colors, so there are ${n \choose 1}=n$ ways to permute the first column’s rows. Thus every other column will have to contain one different permutation of the first column; otherwise, there will be at least one permutation of which there is no corresponding column. Furthermore, each permutation will be different, so each row will contain one and only one red square, which also fulfills the case of if Colin permutes the coloring first. Thus there are $n\cdot (n-1)\cdot(n-2)\cdot\cdot\cdot2\cdot1=n!$ different colorings for this case (the same as choosing squares such as no square is in the same row or column as any other square). 2. $m=n-1$ This is essentially the same as case 1 except for the coloring; now there is one blue square and the rest are red squares. Thus there are also $n!$ different colorings for this case. 3. $m=0$ Since we have an entirely blue column, we are unable to have a column with $1$ red square only as doing so would leave one permutation that is not covered by at least one column (that space is being taken for the blank column). We are also unable to have a completely blue column as doing so would allow for Colin to shift the columns and in doing so fail for Rowan to shift back the columns. We also cannot have a column with any other number of red squares other than $0$ as will be shown below, so there is $1$ case here in which the entire coloring is red. 4. $m=n$ This is the same is an entire blue column, and, similar to above, we have $1$ coloring. 5. $1<m<n-1$ This is the final case and is equivalent to permuting for ${n \choose m}$ different ways. We must prove that this is greater than $n$ to show that the columns are not able to contain every possible permutation of this column for all values of $n$ such that $n>3$ (when $n=3$ , there is no such positive integer $m$ that satisfies the conditions). Note that if we have any column with a different number of red squares, it is an unattainable column and is thus not optimal. Lemma: Given that $m$ and $n$ are positive integers such that $1<m<n-1$ and $n>3$ , it is true for all $m$ and $n$ that ${n \choose m}>n$ . Proof: Assume that $m<\frac{n-1}{2}$ . $\Leftrightarrow$ $m+1<n-m$ $\Leftrightarrow$ $(m+1)!(n-m-1)!<m!(n-m)!$ $\Leftrightarrow$ $\frac{n!}{m!(n-m)!}<\frac{n!}{(m+1)!(n-m-1)!}$ $\Leftrightarrow$ ${n \choose m}<{n \choose m+1}$ Similarly, we can prove that ${n \choose m}>{n \choose m+1}$ for $m>\frac{n-1}{2}$ . Now we split our proof into two cases. Case 1: $n$ is even. The largest integer less than $\frac{n-1}{2}$ is $\frac{n}{2}-1$ , so we know that: ${n \choose \frac{n}{2}}>{n \choose \frac{n}{2}-1}>\cdot\cdot\cdot>{n \choose 2}$ by induction. On the other hand, the smallest integer greater than $\frac{n-1}{2}$ is $\frac{n}{2}$ , so we know that: ${n \choose \frac{n}{2}}>{n \choose \frac{n}{2}+1}>\cdot\cdot\cdot>{n \choose n-2}$ also by induction. Thus out of the given range for $m$ we know that ${n \choose 2}$ and ${n \choose n-2}$ are the minimum values, and all that is left is to prove that they are both greater than $n$ . Furthermore, since ${n \choose 2}={n \choose n-2}$ , we only have to prove that ${n \choose 2}>n$ . We start with the given: $n>3$ $\Leftrightarrow$ $\frac{n-1}{2}>1$ $\Leftrightarrow$ $\frac{n(n-1)}{2}>n$ $\Leftrightarrow$ $\frac{n!}{2!(n-2)!}>n$ $\Leftrightarrow$ ${n \choose 2}>n$ Thus we have proven the inequality for all even $n$ . Case 2: $n$ is odd. The greatest integer less than $\frac{n-1}{2}$ is $\frac{n-3}{2}$ , so we know that: ${n \choose \frac{n-1}{2}}>{n \choose \frac{n-3}{2}}>\cdot\cdot\cdot>{n \choose 2}$ by induction. On the other hand, the smallest integer greater than $\frac{n-1}{2}$ is $\frac{n+1}{2}$ , so we know that: ${n \choose \frac{n+1}{2}}>{n \choose \frac{n+3}{2}}>\cdot\cdot\cdot>{n \choose n-2}$ also by induction. Since ${n \choose \frac{n+1}{2}}={n \choose \frac{n-1}{2}}$ , we know that once again, ${n \choose n-2}={n \choose 2}$ is the minimum of the given range for $m$ , and the same proof applies. Thus, the inequality holds true for odd and in turn all positive integers $n>3$ . As a result, due to our lemma, there are always more permutations of the columns than the number of columns itself, so there will always exist a permutation of the column such that there are no corresponding original columns of which to match with. Thus there are no solutions for this case. In conclusion, there are a total of $2\cdot n!+2$ different colorings for which the above apply. ~eevee9406
\[ 2 \cdot n! + 2 \]
usajmo
[ "Mathematics -> Geometry -> Plane Geometry -> Triangulations" ]
8
Let $ABC$ be a fixed acute triangle inscribed in a circle $\omega$ with center $O$ . A variable point $X$ is chosen on minor arc $AB$ of $\omega$ , and segments $CX$ and $AB$ meet at $D$ . Denote by $O_1$ and $O_2$ the circumcenters of triangles $ADX$ and $BDX$ , respectively. Determine all points $X$ for which the area of triangle $OO_1O_2$ is minimized.
Let $E$ be midpoint $AD.$ Let $F$ be midpoint $BD \implies$ \[EF = ED + FD = \frac {AD}{2} + \frac {BD}{2} = \frac {AB}{2}.\] $E$ and $F$ are the bases of perpendiculars dropped from $O_1$ and $O_2,$ respectively. Therefore $O_1O_2 \ge EF = \frac {AB}{2}.$ \[CX \perp O_1O_2, AX \perp O_1O \implies \angle O O_1O_2 = \angle AXC\] $\angle AXC = \angle ABC (AXBC$ is cyclic) $\implies \angle O O_1O_2 = \angle ABC.$ Similarly $\angle BAC = \angle O O_2 O_1 \implies \triangle ABC \sim \triangle O_2 O_1O.$ The area of $\triangle OO_1O_2$ is minimized if $CX \perp AB$ because \[\frac {[OO_1O_2]} {[ABC]} = \left(\frac {O_1 O_2} {AB}\right)^2 \ge \left(\frac {EF} {AB}\right)^2 = \frac {1}{4}.\] [email protected], vvsss
The area of triangle $OO_1O_2$ is minimized if $CX \perp AB$.
usamo
[ "Mathematics -> Algebra -> Algebra -> Equations and Inequalities", "Mathematics -> Number Theory -> Prime Numbers" ]
5
Find all prime numbers $p,q,r$ , such that $\frac{p}{q}-\frac{4}{r+1}=1$
The given equation can be rearranged into the below form: $4q = (p-q)(r+1)$ $Case 1: 4|(p-q)$ then we have $q = ((p-q)/4)(r+1)$ $=> (p-q)/4 = 1$ and $q = r + 1$ $=> r = 2, q = 3$ and $p = 7$ $Case 2: 4|(r+1)$ then we have $q = (p-q)((r+1)/4)$ $=> (p-q) = 1$ and $q = (r + 1)/4$ $=> p = q + 1 => q = 2, p = 3$ and $r = 7$ note that if $(r+1)/4 = 1$ , then $q = (p-q) => p = 2q$ which is a contradiction. $Case 3: 2|(p-q)$ and $2|(r+1)$ then we have $q = ((p-q)/2)((r+1)/2)$ $=> (p-q)/2 = 1$ and $q = (r+1)/2$ $=> p = q + 2$ and $r = 2q - 1$ We have that exactly one of $q, q + 1, q + 2$ is a multiple of $3$ . $q + 1$ cannot be a multiple of $3$ since $q + 1 = 3k => q = 3k - 1$ . Since $r = 2q - 1$ is prime, then we have $2(3k - 1) - 1 = 3(2k-1)$ is a prime. $=> k = 1 => q = 2 => p = 4$ contradiction. Also, $q + 2$ cannot be a multiple of $3$ since, $q + 2 = 3k => p = 3k => k = 1 => q = 1$ contradiction. So, $q = 3k$ $=> k = 1 => q = 3, p = 5$ and $r = 5$ Thus we have the following solutions: $(7, 3, 2), (3, 2, 7), (5, 3, 5)$ $Kris17$
\[ (7, 3, 2), (3, 2, 7), (5, 3, 5) \]
jbmo
[ "Mathematics -> Discrete Mathematics -> Combinatorics", "Mathematics -> Geometry -> Plane Geometry -> Polygons" ]
7
( Reid Barton ) An animal with $n$ cells is a connected figure consisting of $n$ equal-sized square cells. ${}^1$ The figure below shows an 8-cell animal. A dinosaur is an animal with at least 2007 cells. It is said to be primitive if its cells cannot be partitioned into two or more dinosaurs. Find with proof the maximum number of cells in a primitive dinosaur. Animals are also called polyominoes . They can be defined inductively . Two cells are adjacent if they share a complete edge . A single cell is an animal, and given an animal with cells, one with cells is obtained by adjoining a new cell by making it adjacent to one or more existing cells.
Solution 1 Let a $n$ -dino denote an animal with $n$ or more cells. We show by induction that an $n$ -dino with $4n-2$ or more animal cells is not primitive. (Note: if it had more, we could just take off enough until it had $4n-2$ , which would have a partition, and then add the cells back on.) Base Case: If $n=1$ , we have two cells, which are clearly not primitive. Inductive Step: Assume any $4n-2$ cell animal can be partitioned into two or more $n$ -dinos. For a given $(4n+2)$ -dino, take off any four cells (call them $w,\ x,\ y,\ z$ ) to get an animal with $4n-2$ cells. This can be partitioned into two or more $n$ -dinos, let's call them $A$ and $B$ . This means that $A$ and $B$ are connected. If both $A$ and $B$ are $(n+1)$ -dinos or if $w,\ x,\ y,\ z$ don't all attach to one of them, then we're done. So assume $A$ has $n$ cells and thus $B$ has at least $3n-2$ cells, and that $w,\ x,\ y,\ z$ are added to $B$ . So $B$ has $3n+2$ cells total. Let $C$ denote the cell of $B$ attached to $A$ . There are $3n+1$ cells on $B$ besides $C$ . Thus, of the three (or less) sides of $C$ not attached to $A$ , one of them must have $n+1$ cells by the pigeonhole principle . It then follows that we can add $A$ , $C$ , and the other two sides together to get an $(n+1)$ dino, and the side of $C$ that has $n+1$ cells is also an $n+1$ -dino, so we can partition the animal with $4n+2$ cells into two $(n+1)$ -dinos and we're done. Thus, our answer is $4(2007) - 3 = 8025$ cells. Example of a solution Attempting to partition solution into dinosaurs Solution 2 For simplicity, let $k=2007$ and let $n$ be the number of squares . Let the centers of the squares be vertices , and connect any centers of adjacent squares with edges. Suppose we have some loops . Just remove an edge in the loop. We are still connected since you can go around the other way in the loop. Now we have no loops. Each vertex can have at most 4 edges coming out of it. For each point, assign it the quadruple : $(a,b,c,d)$ where $a$ , $b$ , $c$ , $d$ are the numbers of vertices on each branch, WLOG $a\ge b\ge c\ge d$ . Note $a+b+c+d=n-1$ . Claim: If $n=4k-2$ , then we must be able to divide the animal into two dinosaurs. Chose a vertex, $v$ , for which $a$ is minimal (i.e. out of all maximal elements in a quadruple, choose the one with the least maximal element). We have that $4a \ge a+b+c+d=4k-3$ , so $a\ge k$ . Hence we can just cut off that branch, that forms a dinosaur. But suppose the remaining vertices do not make a dinosaur. Then we have $b+c+d+1\le k-1 \iff n-a\le k-1\iff a\ge 3k-1$ . Now move to the first point on the branch at $a$ . We have a new quadruple $p,\ q,\ r,\ b+c+d+1$ ) where $p+q+r=a-1\ge 3k-2$ . Now consider the maximal element of that quadruple. We already have $b+c+d+1\le k-1$ . WLOG $p\ge q\ge r\ge 0$ , then $3p\ge p+q+r=a-1\ge 3k-2\implies p\ge k$ so $p>k-1=b+c+d+1$ , so $p$ is the maximal element of that quadruple. Also $a-1=p+q+r\ge p+0+0$ , so $p<a$ . But that is a contradiction to the minimality of $a$ . Therefore, we must have that $b+c+d+1\ge k$ , so we have a partition of two dinosaurs. Maximum: $n=4k-3$ . Consider a cross with each branch having $k-1$ verticies. Clearly if we take partition $k$ vertices, we remove the center, and we are not connected. So $k=2007$ : $4\cdot 2007-3=8025$ . Solution 3 (Generalization) Turn the dinosaur into a graph (cells are vertices , adjacent cells connected by an edge) and prove this result about graphs. A connected graph with $V$ vertices, where each vertex has degree less than or equal to $D$ , can be partitioned into connected components of sizes at least $\frac{V-1}{D}$ . So then in this special case, we have $D = 4$ , and so $V = 2006 \times 4+1$ (a possible configuration of this size that works consists of a center and 4 lines of cells each of size 2006 connected to the center). We next throw out all the geometry of this situation, so that we have a completely unconstrained graph. If we prove the above-mentioned result, we can put the geometry back in later by taking the connected components that our partition gives us, then filling back all edges that have to be there due to adjacent cells. This won't change any of the problem constraints, so we can legitimately do this. Going, now, to the case of arbitrary graphs, we WOP on the number of edges. If we can remove any edge and still have a connected graph, then we have found a smaller graph that does not obey our theorem, a contradiction due to the minimality imposed by WOP. Therefore, the only case we have to worry about is when the graph is a tree. If it's a tree, we can root the tree and consider the size of subtrees. Pick the root such that the size of the largest subtree is minimized. This minimum must be at least $\frac{V-1}{D}$ , otherwise the sum of the size of the subtrees is smaller than the size of the graph, which is a contradiction. Also, it must be at most $\frac{V}{2}$ , or else pick the subtree of size greater than $\frac{V}{2}$ and you have decreased the size of the largest subtree if you root from that vertex instead, so you have some subtree with size between $\frac{V-1}{D}$ and $\frac V2$ . Cut the edge connecting the root to that subtree, and use that as your partition. It is easy to see that these partitions satisfy the contention of our theorem, so we are done. Solution 4 Let $s$ denote the minimum number of cells in a dinosaur; the number this year is $s = 2007$ . Claim: The maximum number of cells in a primitive dinosaur is $4(s - 1) + 1$ . First, a primitive dinosaur can contain up to $4(s - 1) + 1$ cells. To see this, consider a dinosaur in the form of a cross consisting of a central cell and four arms with $s - 1$ cells apiece. No connected figure with at least $s$ cells can be removed without disconnecting the dinosaur. The proof that no dinosaur with at least $4(s - 1) + 2$ cells is primitive relies on the following result. Lemma. Let $D$ be a dinosaur having at least $4(s - 1) + 2$ cells, and let $R$ (red) and $B$ (black) be two complementary animals in $D$ , i.e. $R\cap B = \emptyset$ and $R\cup B = D$ . Suppose $|R|\leq s - 1$ . Then $R$ can be augmented to produce animals $\~{R}\subset R$ (Error compiling LaTeX. Unknown error_msg) and $\{B} = D\backslash\{R}$ (Error compiling LaTeX. Unknown error_msg) such that at least one of the following holds: (i) $|\{R}|\geq s$ (Error compiling LaTeX. Unknown error_msg) and $|\~{B}|\geq s$ (Error compiling LaTeX. Unknown error_msg) , (ii) $|\{R}| = |R| + 1$ (Error compiling LaTeX. Unknown error_msg) , (iii) $|R| < |\{R}|\leq s - 1$ (Error compiling LaTeX. Unknown error_msg) . Proof. If there is a black cell adjacent to $R$ that can be made red without disconnecting $B$ , then (ii) holds. Otherwise, there is a black cell $c$ adjacent to $R$ whose removal disconnects $B$ . Of the squares adjacent to $c$ , at least one is red, and at least one is black, otherwise $B$ would be disconnected. Then there are at most three resulting components $\mathcal{C}_1, \mathcal{C}_2, \mathcal{C}_3$ of $B$ after the removal of $c$ . Without loss of generality, $\mathcal{C}_3$ is the largest of the remaining components. (Note that $\mathcal{C}_1$ or $\mathcal{C}_2$ may be empty.) Now $\mathcal{C}_3$ has at least $\lceil (3s - 2)/3\rceil = s$ cells. Let $\{B} = \mathcal{C}_3$ (Error compiling LaTeX. Unknown error_msg) . Then $|\{R}| = |R| + |\mathcal{C}_1| + |\mathcal{C}_2| + 1$ (Error compiling LaTeX. Unknown error_msg) . If $|\{B}|\leq 3s - 2$ (Error compiling LaTeX. Unknown error_msg) , then $|\{R}|\geq s$ (Error compiling LaTeX. Unknown error_msg) and (i) holds. If $|\{B}|\geq 3s - 1$ (Error compiling LaTeX. Unknown error_msg) then either (ii) or (iii) holds, depending on whether $|\{R}|\geq s$ (Error compiling LaTeX. Unknown error_msg) or not. $\blacksquare$ Starting with $|R| = 1$ , repeatedly apply the Lemma. Because in alternatives (ii) and (iii) $|R|$ increases but remains less than $s$ , alternative (i) eventually must occur. This shows that no dinosaur with at least $4(s - 1) + 2$ cells is primitive. Alternate solutions are always welcome. If you have a different, elegant solution to this problem, please add it to this page.
\[ 4 \cdot 2007 - 3 = 8025 \]
usamo
[ "Mathematics -> Applied Mathematics -> Statistics -> Probability -> Counting Methods -> Permutations", "Mathematics -> Number Theory -> Perfect Squares -> Other" ]
6.5
A permutation of the set of positive integers $[n] = \{1, 2, \ldots, n\}$ is a sequence $(a_1, a_2, \ldots, a_n)$ such that each element of $[n]$ appears precisely one time as a term of the sequence. For example, $(3, 5, 1, 2, 4)$ is a permutation of $[5]$ . Let $P(n)$ be the number of permutations of $[n]$ for which $ka_k$ is a perfect square for all $1\leq k\leq n$ . Find with proof the smallest $n$ such that $P(n)$ is a multiple of $2010$ .
We claim that the smallest $n$ is $67^2 = \boxed{4489}$ . Solution 1 Let $S = \{1, 4, 9, \ldots\}$ be the set of positive perfect squares. We claim that the relation $R = \{(j, k)\in [n]\times[n]\mid jk\in S\}$ is an equivalence relation on $[n]$ . It is reflexive because for all . It is symmetric because . It is transitive because if and , then , since is closed under multiplication and a non-square times a square is always a non-square. We are restricted to permutations for which $ka_k \in S$ , in other words to permutations that send each element of $[n]$ into its equivalence class. Suppose there are $N$ equivalence classes: $C_1, \ldots, C_N$ . Let $n_i$ be the number of elements of $C_i$ , then $P(n) = \prod_{i=1}^{N} n_i!$ . Now $2010 = 2\cdot 3\cdot 5\cdot 67$ . In order that $2010\mid P(n)$ , we must have $67\mid n_m!$ for the class $C_m$ with the most elements. This means $n_m\geq 67$ , since no smaller factorial will have $67$ as a factor. This condition is sufficient, since $n_m!$ will be divisible by $30$ for $n_m\geq 5$ , and even more so $n_m\geq 67$ . The smallest element $g_m$ of the equivalence class $C_m$ is square-free, since if it were divisible by the square of a prime, the quotient would be a smaller element of $C_m$ . Also, each prime $p$ that divides $g_m$ divides all the other elements $k$ of $C_m$ , since $p^2\mid kg_m$ and thus $p\mid k$ . Therefore $g_m\mid k$ for all $k\in C_m$ . The primes that are not in $g_m$ occur an even number of times in each $k\in C_m$ . Thus the equivalence class $C_m = \{g_mk^2\leq n\}$ . With $g_m = 1$ , we get the largest possible $n_m$ . This is just the set of squares in $[n]$ , of which we need at least $67$ , so $n\geq 67^2$ . This condition is necessary and sufficient. Solution 2 This proof can also be rephrased as follows, in a longer way, but with fewer highly technical words such as "equivalence relation": It is possible to write all positive integers $n$ in the form $p\cdot m^2$ , where $m^2$ is the largest perfect square dividing $n$ , so $p$ is not divisible by the square of any prime. Obviously, one working permutation of $[n]$ is simply $(1, 2, \ldots, n)$ ; this is acceptable, as $ka_k$ is always $k^2$ in this sequence. Lemma 1. We can permute any numbers that, when each divided by the largest perfect square that divides it, yield equal quantities $p$ . Proof. Let $p_k$ and $m_k$ be the values of $p$ and $m$ , respectively, for a given $k$ as defined above, such that $p$ is not divisible by the square of any prime. We can obviously permute two numbers which have the same $p$ , since if $p_j = p_w$ where $j$ and $w$ are 2 values of $k$ , then $j\cdot w = p_j^2\cdot m_j^2\cdot m_w^2$ , which is a perfect square. This proves that we can permute any numbers with the same value of $p$ . End Lemma Lemma 2. We will prove the converse of Lemma 1: Let one number have a $p$ value of $\phi$ and another, $\gamma$ . $\phi\cdot f$ and $\gamma\cdot g$ are both perfect squares. Proof. $\phi\cdot f$ and $\gamma\cdot g$ are both perfect squares, so for $\phi\cdot \gamma$ to be a perfect square, if $g$ is greater than or equal to $f$ , $g/f$ must be a perfect square, too. Thus $g$ is $f$ times a square, but $g$ cannot divide any squares besides $1$ , so $g = 1f$ ; $g = f$ . Similarly, if $f\geq g$ , then $f = g$ for our rules to keep working. End Lemma We can permute $l$ numbers with the same $p$ in $l!$ ways. We must have at least 67 numbers with a certain $p$ so our product will be divisible by 67. Obviously, then it will also be divisible by 2, 3, and 5, and thus 2010, as well. Toms as $h$ , in general, we need numbers all the way up to $h\cdot 67^2$ , so obviously, $67^2$ is the smallest such number such that we can get a $67!$ term; here 67 $p$ terms are 1. Thus we need the integers $1, 2, \ldots, 67^2$ , so $67^2$ , or $\boxed{4489}$ , is the answer.
\(\boxed{4489}\)
usajmo
[ "Mathematics -> Discrete Mathematics -> Combinatorics" ]
5
For any odd positive integer $n$, let $r(n)$ be the odd positive integer such that the binary representation of $r(n)$ is the binary representation of $n$ written backwards. For example, $r(2023)=r\left(11111100111_{2}\right)=11100111111_{2}=1855$. Determine, with proof, whether there exists a strictly increasing eight-term arithmetic progression $a_{1}, \ldots, a_{8}$ of odd positive integers such that $r\left(a_{1}\right), \ldots, r\left(a_{8}\right)$ is an arithmetic progression in that order.
The main idea is the following claim. Claim: If $a, b, c$ are in arithmetic progression and have the same number of digits in their binary representations, then $r(a), r(b), r(c)$ cannot be in arithmetic progression in that order. Proof. Consider the least significant digit that differs in $a$ and $b$. Then $c$ will have the same value of that digit as $a$, which will be different from $b$. Since this becomes the most significant digit in $r(a), r(b), r(c)$, then of course $b$ cannot be between $a$ and $c$. To finish, we just need to show that if there are 8 numbers in arithmetic progression, which we'll write as $a_{1}, a_{1}+d, a_{1}+2 d, \ldots, a_{1}+7 d$, three of them have the same number of digits. We have a few cases. - If $a_{1}+3 d<2^{k} \leq a_{1}+4 d$, then $a_{1}+4 d, a_{1}+5 d, a_{1}+6 d$ will have the same number of digits. - If $a_{1}+4 d<2^{k} \leq a_{1}+5 d$, then $a_{1}+5 d, a_{1}+6 d, a_{1}+7 d$ will have the same number of digits. - If neither of these assumptions are true, $a_{1}+3 d, a_{1}+4 d, a_{1}+5 d$ will have the same number of digits. Having exhausted all cases, we are done.
There does not exist a strictly increasing eight-term arithmetic progression \(a_{1}, \ldots, a_{8}\) of odd positive integers such that \(r(a_{1}), \ldots, r(a_{8})\) is an arithmetic progression in that order.
HMMT_2
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations", "Mathematics -> Number Theory -> Divisibility -> Other" ]
4.5
Find all ordered pairs $(a,b)$ of positive integers for which the numbers $\dfrac{a^3b-1}{a+1}$ and $\dfrac{b^3a+1}{b-1}$ are both positive integers
Adding $1$ to both the given numbers we get: $\dfrac{a^3b-1}{a+1} + 1$ is also a positive integer so we have: $\dfrac{a^3b+a}{a+1}$ = $\dfrac{a(a^2b+1)}{a+1}$ is a positive integer $\implies (a+1) \mid (a^2b+1)$ $\implies (a+1) \mid (((a+1) - 1)^2b+1)$ $\implies (a+1) \mid (b+1)$ Similarly, $\dfrac{b^3a+1}{b-1} + 1$ is also a positive integer so we have: $\dfrac{b^3a+b}{b-1}$ = $\dfrac{b(b^2a+1)}{b-1}$ is a positive integer $\implies (b-1) | (b^2a+1)$ $\implies (b-1) | (((b-1) + 1)^2a+1)$ $\implies (b-1) | (a+1)$ Combining above $2$ results we get: $(b-1) | (b+1)$ $\implies b=2,3$ $Case 1: b=2$ $\implies a+1|3 \implies a=2$ which is a valid solution. $Case 2: b=3$ $\implies a+1|4 \implies a=1,3$ which are valid solutions. Thus, all solutions are: $(2,2), (1,3), (3,3)$ $Kris17$
\[ \{(2,2), (1,3), (3,3)\} \]
jbmo
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations" ]
8
Determine each real root of $x^4-(2\cdot10^{10}+1)x^2-x+10^{20}+10^{10}-1=0$ correct to four decimal places.
The equation can be re-written as \begin{align}\label{eqn1} (x+10^5)^2(x-10^5)^2 -(x+10^5)(x-10^5) -x-1=0. \end{align} We first prove that the equation has no negative roots. Let $x\le 0.$ The equation above can be further re-arranged as \begin{align*}[(x+10^5)(x-10^5)+1][(x+10^5)(x-10^5)-2]=x-1.\end{align*} The right hand side of the equation is negative. Therefore \[[(x+10^5)(x-10^5)+1][(x+10^5)(x-10^5)-2)]<0,\] and we have $-1<(x+10^5)(x-10^5) <2.$ Then the left hand side of the equation is bounded by \[|[(x+10^5)(x-10^5)+1][(x+10^5)(x-10^5)-2]|\le 3\times 3.\] However, since $|(x+10^5)(x-10^5)|\le 2$ and $x<0,$ it follows that $|x+10^5| <\frac{2}{|x-10^5|}<2\times 10^{-5}$ for negative $x.$ Then $x<2\times 10^{-5}-10^5.$ The right hand side of the equation is then a large negative number. It cannot be equal to the left hand side which is bounded by 9. Now let $x>0.$ When $x=10^5,$ the left hand side of equation (1) is negative. Therefore the equation has real roots on both side of $10^5$ , as its leading coefficient is positive. We will prove that $x=10^5$ is a good approximation of the roots (within $10^{-2}$ ). In fact, we can solve the "quadratic" equation (1) for $(x+10^5)(x-10^5)$ : \[(x+10^5)(x-10^5)=\frac{1\pm\sqrt{1+4(x+1)}}{2}.\] Then \[x-10^5=\frac{1\pm\sqrt{1+4(x+1)}}{2(x+10^5)}.\] Easy to see that $|x-10^5| <1$ for positve $x.$ Therefore, $10^5-1<x<10^5+1.$ Then \begin{align*} |x-10^5|&=\left|\frac{1\pm\sqrt{1+4(x+1)}}{2(x+10^5)}\right |\\ &\le \left |\frac{1}{2(x+10^5)}\right |+\left |\frac{\sqrt{1+4(x+1)}}{2(x+10^5)}\right |\\ &\le \frac{1}{2(10^5-1+10^5)} +\frac{\sqrt{1+4(10^5+1+1)}}{2(10^5-1+10^5)} \\ &<10^{-2}. \end{align*} Let $x_1$ be a root of the equation with $x_1<10^5.$ Then $0<10^5-x_1<10^{-2}$ and \[x_1-10^5=\frac{1-\sqrt{1+4(x_1+1)}}{2(x_1+10^5)}.\] An aproximation of $x_1$ is defined as follows: \[\tilde{x}_1=10^5+\frac{1-\sqrt{1+4(10^5+1)}}{2(10^5+10^5)}.\] We check the error of the estimate: \begin{align*} |\tilde{x}_1-x_1|&=\left | \frac{1-\sqrt{1+4(10^5+1)}}{2(10^5+10^5)}- \frac{1-\sqrt{1+4(x_1+1)}}{2(x_1+10^5)} \right | \\ &\le \left |\frac{1}{2(10^5+10^5)}- \frac{1}{2(x_1+10^5)}\right |+\left |\frac{\sqrt{1+4(10^5+1)}}{2(10^5+10^5)}- \frac{\sqrt{1+4(x_1+1)}}{2(x_1+10^5)}\right |. \end{align*} The first absolute value \[\left |\frac{1}{2(10^5+10^5)}- \frac{1}{2(x_1+10^5)}\right | =\frac{|x_1- 10^5|}{2(10^5+10^5)(x_1+10^5)}<10^{-12}.\] The second absolute value \begin{align*} &\left |\frac{\sqrt{1+4(10^5+1)}}{2(10^5+10^5)} - \frac{\sqrt{1+4(x_1+1)}}{2(x_1+10^5)} \right |\\ &\le \left |\frac{\sqrt{1+4(10^5+1)}}{2(10^5+10^5)}- \frac{\sqrt{1+4(x_1+1)}}{2(10^5+10^5)}\right |+\left |\frac{\sqrt{1+4(x_1+1)}}{2(10^5+10^5)}- \frac{\sqrt{1+4(x_1+1)}}{2(x_1+10^5)}\right |\\ &\le 10^{-7}+10^{-9}, \end{align*} through a rationalized numerator.Therefore $|\tilde{x}_1-x_1|\le 10^{-6}.$ For a real root $x_2$ with $x_2>10^5,$ we choose \[\tilde{x}_2=10^5+\frac{1+\sqrt{1+4(10^5+1)}}{2(10^5+10^5)}.\] We can similarly prove it has the desired approximation.
The real roots of the equation \(x^4-(2\cdot10^{10}+1)x^2-x+10^{20}+10^{10}-1=0\) correct to four decimal places are approximately: \[ x_1 \approx 10^5 - \frac{1 - \sqrt{1 + 4(10^5 + 1)}}{2 \cdot 2 \cdot 10^5} \] \[ x_2 \approx 10^5 + \frac{1 + \sqrt{1 + 4(10^5 + 1)}}{2 \cdot 2 \cdot 10^5} \] Given the approximations: \[ x_1 \approx 99999.5000 \] \[ x_2 \approx 100000.5000 \]
usamo
[ "Mathematics -> Geometry -> Plane Geometry -> Polygons" ]
7
Find the minimum positive integer $n\ge 3$, such that there exist $n$ points $A_1,A_2,\cdots, A_n$ satisfying no three points are collinear and for any $1\le i\le n$, there exist $1\le j \le n (j\neq i)$, segment $A_jA_{j+1}$ pass through the midpoint of segment $A_iA_{i+1}$, where $A_{n+1}=A_1$
To find the minimum positive integer \( n \geq 3 \) such that there exist \( n \) points \( A_1, A_2, \ldots, A_n \) satisfying no three points are collinear and for any \( 1 \leq i \leq n \), there exists \( 1 \leq j \leq n \) (with \( j \neq i \)), such that the segment \( A_jA_{j+1} \) passes through the midpoint of segment \( A_iA_{i+1} \), where \( A_{n+1} = A_1 \), we proceed as follows: First, it is necessary to verify that \( n = 3 \) and \( n = 4 \) do not satisfy the given conditions. Through geometric construction and analysis, it can be shown that no such configurations exist for these values of \( n \). Next, consider \( n = 5 \). We analyze two cases: 1. **Case 1**: There are no parallelograms formed by any four of the points \( A_i \). By detailed geometric analysis and coordinate bashing, it can be shown that no such five points exist. 2. **Case 2**: Assume \( A_1A_4A_2A_3 \) forms a parallelogram. By considering the reflection of points and ensuring no three points are collinear, it leads to a contradiction, proving that \( n = 5 \) is also not possible. Finally, for \( n = 6 \), a construction exists that satisfies all the given conditions. Therefore, the minimum positive integer \( n \) for which the conditions hold is \( n = 6 \). The answer is: \boxed{6}.
6
china_national_olympiad
[ "Mathematics -> Discrete Mathematics -> Combinatorics", "Mathematics -> Geometry -> Plane Geometry -> Area" ]
5
Lily has a $300 \times 300$ grid of squares. She now removes $100 \times 100$ squares from each of the four corners and colors each of the remaining 50000 squares black and white. Given that no $2 \times 2$ square is colored in a checkerboard pattern, find the maximum possible number of (unordered) pairs of squares such that one is black, one is white and the squares share an edge.
First we show an upper bound. Define a grid point as a vertex of one of the squares in the figure. Construct a graph as follows. Place a vertex at each grid point and draw an edge between two adjacent points if that edge forms a black-white boundary. The condition of there being no $2 \times 2$ checkerboard is equivalent to no vertex having degree more than 2. There are $101^{2}+4 \cdot 99^{2}=49405$ vertices that are allowed to have degree 2 and $12 \cdot 99=1188$ vertices (on the boundary) that can have degree 1. This gives us an upper bound of 49999 edges. We will show that exactly this many edges is impossible. Assume for the sake of contradiction that we have a configuration achieving exactly this many edges. Consider pairing up the degree 1 vertices so that those on a horizontal edge pair with the other vertex in the same column and those on a vertical edge pair with the other vertex in the same row. If we combine the pairs into one vertex, the resulting graph must have all vertices with degree exactly 2. This means the graph must be a union of disjoint cycles. However all cycles must have even length and there are an odd number of total vertices so this is impossible. Thus we have an upper bound of 49998. We now describe the construction. The top row alternates black and white. The next 99 rows alternate between all black and all white. Let's say the second row from the top is all white. The $101^{\text {st }}$ row alternates black and white for the first 100 squares, is all black for the next 100 and alternates between white and black for the last 100 squares. The next 98 rows alternate between all black and all white (the $102^{\text {nd }}$ row is all white). Finally, the bottom 101 rows are a mirror of the top 101 rows with the colors reversed. We easily verify that this achieves the desired.
49998
HMMT_2
[ "Mathematics -> Calculus -> Differential Calculus -> Applications of Derivatives", "Mathematics -> Algebra -> Intermediate Algebra -> Inequalities" ]
7
Consider functions $f : [0, 1] \rightarrow \mathbb{R}$ which satisfy (i) for all in , (ii) , (iii) whenever , , and are all in . Find, with proof, the smallest constant $c$ such that $f(x) \le cx$ for every function $f$ satisfying (i)-(iii) and every $x$ in $[0, 1]$ .
My claim: $c\ge2$ Lemma 1 ) $f\left(\left(\frac{1}{2}\right)^n\right)\le\left(\frac{1}{2}\right)^n$ for $n\in \mathbb{Z}, n\ge0$ For $n=0$ , $f(1)=1$ (ii) Assume that it is true for $n-1$ , then $f\left(\left(\frac{1}{2}\right)^{n}\right)+f\left(\left(\frac{1}{2}\right)^{n}\right)\le f\left(\left(\frac{1}{2}\right)^{n-1}\right)\le \left(\frac{1}{2}\right)^{n-1}$ $f\left(\left(\frac{1}{2}\right)^{n}\right)\le \left(\frac{1}{2}\right)^{n}$ By principle of induction, lemma 1 is proven . Lemma 2 ) For any $x$ , $\left(\frac{1}{2}\right)^{n+1}<x\le\left(\frac{1}{2}\right)^n\le1$ and $n\in \mathbb{Z}$ , $f(x)\le\left(\frac{1}{2}\right)^n$ . $f(x)+f\left(\left(\frac{1}{2}\right)^n-x\right)\le f\left(\left(\frac{1}{2}\right)^{n}\right)\le \left(\frac{1}{2}\right)^{n}$ (lemma 1 and (iii) ) $f(x)\le\left(\frac{1}{2}\right)^n$ (because $f\left(\left(\frac{1}{2}\right)^n-x\right)\ge0$ (i) ) $\forall 0\le x\le 1$ , $\left(\frac{1}{2}\right)^{n-1}\ge2x\ge \left(\frac{1}{2}\right)^n\ge f(x)$ . Thus, $c=2$ works. Let's look at a function $g(x)=\left\{\begin{array}{ll}0&0\le x\le \frac{1}{2};\\1&\frac{1}{2}<x\le1;\\\end{array}\right\}$ It clearly have property (i) and (ii). For $0\le x\le\frac{1}{2}$ and WLOG let $x\le y$ , $f(x)+f(y)=0+f(y)\le f(y)$ For $\frac{1}{2}< x\le 1$ , $x+y>1$ . Thus, property (iii) holds too. Thus $g(x)$ is one of the legit function. $\lim_{x\rightarrow\frac{1}{2}^+} cx \ge \lim_{x\rightarrow\frac{1}{2}^+} g(x)=1$ $\frac{1}{2}c>1$ $c>2$ but approach to $2$ when $x$ is extremely close to $\frac{1}{2}$ from the right side. $\mathbb{Q.E.D}$
The smallest constant \( c \) such that \( f(x) \le cx \) for every function \( f \) satisfying the given conditions is \( c = 2 \).
usamo
[ "Mathematics -> Algebra -> Intermediate Algebra -> Exponential Functions" ]
5
Find all pairs of positive integers $(x,y)$ such that \[x^y = y^{x - y}.\]
Note that $x^y$ is at least one. Then $y^{x - y}$ is at least one, so $x \geq y$ . Write $x = a^{b+c}, y = a^c$ , where $\gcd(b, c) = 1$ . (We know that $b$ is nonnegative because $x\geq y$ .) Then our equation becomes $a^{(b+c)*a^c} = a^{c*(a^{b+c} - a^c)}$ . Taking logarithms base $a$ and dividing through by $a^c$ , we obtain $b + c = c*(a^b - 1)$ . Since $c$ divides the RHS of this equation, it must divide the LHS. Since $\gcd(b, c) = 1$ by assumption, we must have $c = 1$ , so that the equation reduces to $b + 1 = a^b - 1$ , or $b + 2 = a^b$ . This equation has only the solutions $b = 1, a = 3$ and $b = 2, a = 2$ . Therefore, our only solutions are $x = 3^{1 + 1} = 9, y = 3^1 = 3$ , and $x = 2^{2+1} = 8, y = 2^1 = 2$ , and we are done.
The pairs of positive integers \((x, y)\) that satisfy the equation \(x^y = y^{x - y}\) are: \[ (x, y) = (9, 3) \quad \text{and} \quad (x, y) = (8, 2) \]
jbmo
[ "Mathematics -> Algebra -> Algebra -> Polynomial Operations" ]
7.5
Does there exist a two-variable polynomial $P(x, y)$ with real number coefficients such that $P(x, y)$ is positive exactly when $x$ and $y$ are both positive?
No. For any $\epsilon$ and positive $x, P(x, \epsilon)>0$ and $P(x,-\epsilon) \leq 0$. Thus by continuity/IVT, $P(x, 0)=0$ for all positive $x$. Similarly $P(0, y)=0$ for all positive $y$. This implies $x y \mid P(x, y)$, and so we can write $P(x, y)=x y Q(x, y)$. But then this same logic holds for $Q$, and this cannot continue infinitely unless $P$ is identically 0 - in which case the conditions do not hold. So no such polynomial exists.
No such polynomial exists.
HMMT_2