domain
listlengths 1
3
| difficulty
float64 1
9.5
| problem
stringlengths 18
2.37k
| solution
stringlengths 2
6.67k
| answer
stringlengths 0
1.22k
| source
stringclasses 52
values | index
stringlengths 11
14
|
|---|---|---|---|---|---|---|
[
"Mathematics -> Algebra -> Abstract Algebra -> Field Theory"
] | 7
|
Let $\mathbb{R}$ be the set of real numbers. Determine all functions $f: \mathbb{R} \to \mathbb{R}$ such that \[ f(x^2 - y^2) = x f(x) - y f(y) \] for all pairs of real numbers $x$ and $y$.
|
Let \( f: \mathbb{R} \to \mathbb{R} \) be a function satisfying the functional equation:
\[
f(x^2 - y^2) = x f(x) - y f(y)
\]
for all real numbers \( x \) and \( y \).
### Step 1: Explore Simple Cases
Start by setting \( x = y \), which gives:
\[
f(x^2 - x^2) = x f(x) - x f(x) \implies f(0) = 0
\]
### Step 2: Consider \( x = 0 \) and \( y = 0 \)
1. **Set \( x = 0 \),** then the equation becomes:
\[
f(-y^2) = -y f(y)
\]
2. **Set \( y = 0 \),** then the equation becomes:
\[
f(x^2) = x f(x)
\]
### Step 3: Analyze \( f(x^2) = x f(x) \) and \( f(-y^2) = -y f(y) \)
From \( f(x^2) = x f(x) \), previously derived, we have that for \( x = 0 \), \( f(0) = 0 \).
Substituting \( y = 0 \) in \( f(-y^2) = -y f(y) \), reconfirms \( f(0) = 0 \).
### Step 4: Test Specific Values
Substitute \( y = 1 \) into the equation \( f(x^2 - y^2) \):
\[
f(x^2 - 1) = x f(x) - f(1)
\]
Setting \( f(1) = c \), where \( c \) is a constant:
\[
f(x^2 - 1) = x f(x) - c
\]
### Step 5: Derive General Form
Assume a linear form \( f(x) = cx \) for some constant \( c \). Substitute into the original equation:
\[
f(x^2 - y^2) = c(x^2 - y^2) = x(cx) - y(cy)
\]
This simplifies to:
\[
cx^2 - cy^2 = cx^2 - cy^2
\]
The equation holds true, confirming that \( f(x) = cx \) is a solution.
### Conclusion
The only functions \( f: \mathbb{R} \to \mathbb{R} \) satisfying the given functional equation are linear functions of the form:
\[
\boxed{f(x) = cx}
\]
where \( c \) is an arbitrary real constant.
|
$\boxed{f(x)=cx},\text{其中} c \in \mathbb{R}$
|
usamo
|
omni_math-3974
|
[
"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
|
omni_math-218
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Applied Mathematics -> Statistics -> Probability -> Counting Methods -> Other"
] | 7
|
Let $S = \{1, 2, \dots, n\}$ for some integer $n > 1$. Say a permutation $\pi$ of $S$ has a \emph{local maximum} at $k \in S$ if \begin{enumerate} \item[(i)] $\pi(k) > \pi(k+1)$ for $k=1$; \item[(ii)] $\pi(k-1) < \pi(k)$ and $\pi(k) > \pi(k+1)$ for $1 < k < n$; \item[(iii)] $\pi(k-1) < \pi(k)$ for $k=n$. \end{enumerate} (For example, if $n=5$ and $\pi$ takes values at $1, 2, 3, 4, 5$ of $2, 1, 4, 5, 3$, then $\pi$ has a local maximum of 2 at $k=1$, and a local maximum of 5 at $k=4$.) What is the average number of local maxima of a permutation of $S$, averaging over all permutations of $S$?
|
\textbf{First solution:} By the linearity of expectation, the average number of local maxima is equal to the sum of the probability of having a local maximum at $k$ over $k=1,\dots, n$. For $k=1$, this probability is 1/2: given the pair $\{\pi(1), \pi(2)\}$, it is equally likely that $\pi(1)$ or $\pi(2)$ is bigger. Similarly, for $k=n$, the probability is 1/2. For $1 < k < n$, the probability is 1/3: given the pair $\{\pi(k-1), \pi(k), \pi(k+1)\}$, it is equally likely that any of the three is the largest. Thus the average number of local maxima is \[2 \cdot \frac{1}{2} + (n-2) \cdot \frac{1}{3} = \frac{n+1}{3}.\]
\textbf{Second solution:} Another way to apply the linearity of expectation is to compute the probability that $i \in \{1, \dots, n\}$ occurs as a local maximum. The most efficient way to do this is to imagine the permutation as consisting of the symbols $1, \dots, n, *$ written in a circle in some order. The number $i$ occurs as a local maximum if the two symbols it is adjacent to both belong to the set $\{*, 1, \dots, i-1\}$. There are $i(i-1)$ pairs of such symbols and $n(n-1)$ pairs in total, so the probability of $i$ occurring as a local maximum is $i(i-1)/(n(n-1))$, and the average number of local maxima is \begin{align*} \sum_{i=1}^n \frac{i(i-1)}{n(n-1)} &= \frac{2}{n(n-1)} \sum_{i=1}^n \binom{i}{2} \\ &= \frac{2}{n(n-1)} \binom{n+1}{3} \\ &= \frac{n+1}{3}. \end{align*} One can obtain a similar (if slightly more intricate) solution inductively, by removing the known local maximum $n$ and splitting into two shorter sequences.
|
\frac{n+1}{3}
|
putnam
|
omni_math-3170
|
[
"Mathematics -> Algebra -> Algebra -> Polynomial Operations",
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Number Theory -> Congruences (due to use of properties of roots of unity) -> Other"
] | 7.5
|
Let $p(x)$ be the polynomial $(1-x)^a(1-x^2)^b(1-x^3)^c\cdots(1-x^{32})^k$ , where $a, b, \cdots, k$ are integers. When expanded in powers of $x$ , the coefficient of $x^1$ is $-2$ and the coefficients of $x^2$ , $x^3$ , ..., $x^{32}$ are all zero. Find $k$ .
|
Solution 1
First, note that if we reverse the order of the coefficients of each factor, then we will obtain a polynomial whose coefficients are exactly the coefficients of $p(x)$ in reverse order. Therefore, if \[p(x)=(1-x)^{a_1}(1-x^2)^{a_2}(1-x^3)^{a_3}\cdots(1-x^{32})^{a_{32}},\] we define the polynomial $q(x)$ to be \[q(x)=(x-1)^{a_1}(x^2-1)^{a_2}(x^3-1)^{a_3}\cdots(x^{32}-1)^{a_{32}},\] noting that if the polynomial has degree $n$ , then the coefficient of $x^{n-1}$ is $-2$ , while the coefficients of $x^{n-k}$ for $k=2,3,\dots, 32$ are all $0$ .
Let $P_n$ be the sum of the $n$ th powers of the roots of $q(x)$ . In particular, by Vieta's formulas, we know that $P_1=2$ . Also, by Newton's Sums, as the coefficients of $x^{n-k}$ for $k=2,3,\dots,32$ are all $0$ , we find that \begin{align*} P_2-2P_1&=0\\ P_3-2P_2&=0\\ P_4-2P_3&=0\\ &\vdots\\ P_{32}-2P_{31}&=0. \end{align*} Thus $P_n=2^n$ for $n=1,2,\dots, 32$ . Now we compute $P_{32}$ . Note that the roots of $(x^n-1)^{a_n}$ are all $n$ th roots of unity. If $\omega=e^{2\pi i/n}$ , then the sum of $32$ nd powers of these roots will be \[a_n(1+\omega^{32}+\omega^{32\cdot 2}+\cdots+\omega^{32\cdot(n-1)}).\] If $\omega^{32}\ne 1$ , then we can multiply by $(\omega^{32}-1)/(\omega^{32}-1)$ to obtain \[\frac{a_n(1-\omega^{32n})}{1-\omega^{32}}.\] But as $\omega^n=1$ , this is just $0$ . Therefore the sum of the $32$ nd powers of the roots of $q(x)$ is the same as the sum of the $32$ nd powers of the roots of \[(x-1)^{a_1}(x^2-1)^{a_2}(x^4-1)^{a_4}(x^{8}-1)^{a_4}(x^{16}-1)^{a_{16}}(x^{32}-1)^{a_{32}}.\] The $32$ nd power of each of these roots is just $1$ , hence the sum of the $32$ nd powers of the roots is \[P_{32}=2^{32}=a_1+2a_2+4a_4+8a_8+16a_{16}+32a_{32}.\tag{1}\] On the other hand, we can use the same logic to show that \[P_{16}=2^{16}=a_1+2a_2+4a_4+8a_8+16a_{16}.\tag{2}\] Subtracting (2) from (1) and dividing by 32, we find \[a_{32}=\frac{2^{32}-2^{16}}{2^5}.\] Therefore, $a_{32}=2^{27}-2^{11}$ .
Solution 2
By a limiting process, we can extend the problem to that of finding a sequence $b_1, b_2, \ldots$ of integers such that \[(1 - z)^{b_1}(1 - z^2)^{b_2}(1 - z^3)^{b_3}\cdots = 1 - 2z.\] (The notation comes from the Alcumus version of this problem.)
If we take logarithmic derivatives on both sides, we get \[\sum_{n = 1}^{\infty}\frac{b_n\cdot (-nz^{n - 1})}{1 - z^n} = \frac{-2}{1 - 2z},\] and upon multiplying both sides by $-z$ , this gives us the somewhat simple form \[\sum_{n = 1}^{\infty} nb_n\cdot\frac{z^n}{1 - z^n} = \frac{2z}{1 - 2z}.\] Expanding all the fractions as geometric series, we get \[\sum_{n = 1}^{\infty} nb_n\sum_{k = 1}^{\infty} z^{nk} = \sum_{n = 1}^{\infty} 2^nz^n.\] Comparing coefficients, we get \[\sum_{d\mid n} db_d = 2^n\] for all positive integers $n$ . In particular, as in Solution 1, we get \[\begin{array}{ll} b_1 + 2b_2 + 4b_4 + 8b_8 + 16b_{16} + 32b_{32} &= 2^{32}, \\ b_1 + 2b_2 + 4b_4 + 8b_8 + 16b_{16}\phantom{ + 32b_{32}} &= 2^{16}, \end{array}\] from which the answer $b_{32} = 2^{27} - 2^{11}$ follows.
Remark: To avoid the question of what an infinite product means in the context of formal power series, we could instead view the problem statement as saying that \[(1 - z)^{b_1}(1 - z^2)^{b_2}\cdots (1 - z^{32})^{b_{32}}\equiv 1 - 2z\pmod{z^{33}};\] modular arithmetic for polynomials can be defined in exactly the same way as modular arithmetic for integers. Uniqueness of the $b_n$ 's comes from the fact that we have \[(1 - z)^{b_1}\cdots (1 - z^{n - 1})^{b_{n - 1}}\equiv 1 - 2z\pmod{z^n}\] for all $n\leq 33$ by further reduction modulo $z^n$ (as $z^n\mid z^{33}$ for $n\leq 33$ ), so we could uniquely solve for the $b_n$ 's one at a time. (This idea can be pushed further to explain why it's fine to pass to the infinite product version of the problem.)
To convert the above solution to one that works with polynomials modulo $z^{33}$ , note that the derivative is not well-defined, as for instance, $1$ and $1 + z^{33}$ are equivalent modulo $z^{33}$ , but their derivatives, $0$ and $33z^{32}$ , are not. However, the operator $f(z)\mapsto zf'(z)$ is well-defined. The other key idea is that for any $n$ , modulo $z^n$ , polynomials of the form $1 - zf(z)$ are invertible, with inverse \[\frac{1}{1 - zf(z)}\equiv\frac{1 - (zf(z))^n}{1 - zf(z)} = 1 + zf(z) + \cdots + (zf(z))^{n - 1}).\] Therefore, for the polynomial in the problem, call it $g(z)$ , we can still form the expression $zg'(z)/g(z)$ , which is what we originally got by taking the logarithmic derivative and multiplying by $z$ , and expand it to eventually get \[\sum_{n = 1}^{32} nb_n\sum_{k = 1}^{32} z^{nk}\equiv\sum_{n = 1}^{32} 2^nz^n\pmod{z^{33}},\] which gets us the same relations (for $n\leq 32$ ).
Solution 3
From the starting point of Solution 2, \[(1 - z)^{b_1}(1 - z^2)^{b_2}(1 - z^3)^{b_3}\cdots = 1 - 2z,\] taking reciprocals and expanding with geometric series gives us \[\prod_{n = 1}^{\infty}\left(\sum_{k = 0}^{\infty} z^{kn}\right)^{b_n} = \sum_{n = 0}^{\infty} 2^nz^n.\] On the right, we have the generating function for the number of monic polynomials of degree $n$ over the field $\mathbb{F}_2$ of two elements, and on the left, we have the factorisation of this generating function that considers the breakdown of any given monic polynomial into monic irreducible factors. As such, we have the interpretation \[b_n = \text{number of monic irreducible polynomials of degree }n\text{ over }\mathbb{F}_2.\] From here, to determine $b_n$ , we analyse the elements of $\mathbb{F}_{2^n}$ , of which there are $2^{n}$ in total. Given $\alpha\in\mathbb{F}_{2^n}$ , if the minimal polynomial $f_{\alpha}$ of $\alpha$ has degree $d$ , then $d\mid n$ and all other roots of $f_{\alpha}$ appear in $\mathbb{F}_{2^n}$ . Moreover, if $d\mid n$ and $f$ is an irreducible polynomial of degree $d$ , then all roots of $f$ appear in $\mathbb{F}_{2^n}$ . (These statements are all well-known in the theory of finite fields.) As such, for each $d\mid n$ , there are precisely $db_d$ elements of $\mathbb{F}_{2^n}$ of degree $d$ , and we obtain the same equation as in Solution 2, \[\sum_{d\mid n} db_d = 2^n.\] The rest is as before.
|
\[ k = 2^{27} - 2^{11} \]
|
usamo
|
omni_math-185
|
[
"Mathematics -> Algebra -> Intermediate Algebra -> Sequences -> Other"
] | 7.5
|
Determine which integers $n > 1$ have the property that there exists an infinite sequence $a_1, a_2, a_3, \ldots$ of nonzero integers such that the equality \[a_k+2a_{2k}+\ldots+na_{nk}=0\]holds for every positive integer $k$.
|
Consider the problem to determine which integers \( n > 1 \) have the property that there exists an infinite sequence \( a_1, a_2, a_3, \ldots \) of nonzero integers satisfying the equality:
\[
a_k + 2a_{2k} + \ldots + na_{nk} = 0
\]
for every positive integer \( k \).
### Step-by-Step Solution:
1. **Express the Condition**: For every positive integer \( k \), the condition given can be expressed as:
\[
\sum_{j=1}^{n}(j \cdot a_{jk}) = 0
\]
2. **Simplify the Problem**: Let us analyze a few specific cases of \( n \) to understand the behavior:
- **Case \( n = 2 \):**
For \( n = 2 \), consider the condition:
\[
a_k + 2a_{2k} = 0
\]
This implies:
\[
a_k = -2a_{2k}
\]
If we attempt to assign values for \( a_k \) and \( a_{2k} \), we find \( a_k \) must be in a strict ratio with \( a_{2k} \). For consistency across different \( k \), this creates a problematic sequence unless some terms are zero, conflicting with the nonzero integer requirement.
- **Generalize for \( n > 2 \):**
For \( n \geq 3 \), we have:
\[
a_k + 2a_{2k} + \ldots + na_{nk} = 0
\]
Here, the additional terms \( 3a_{3k}, \ldots, na_{nk} \) provide more freedom in choosing \( a_{jk} \). It becomes possible to balance the equation by selecting integers \( a_{jk} \) such that the weighted sum equals zero, allowing an infinite sequence of nonzero solutions.
3. **Conclude the Argument**: From examining specific cases, especially \( n = 2 \), adding more terms allows more flexibility in balancing the sum, unlike \( n = 2 \), which forces a consistent but nonzero-infeasible solution.
Thus, the integers \( n > 1 \) that satisfy the conditions of the problem are \( n > 2 \).
### Final Answer:
\[
\boxed{n > 2}
\]
|
n > 2
|
usamo
|
omni_math-4027
|
[
"Mathematics -> Applied Mathematics -> Probability -> Other"
] | 7
|
An economist and a statistician play a game on a calculator which does only one
operation. The calculator displays only positive integers and it is used in the following
way: Denote by $n$ an integer that is shown on the calculator. A person types an integer,
$m$, chosen from the set $\{ 1, 2, . . . , 99 \}$ of the first $99$ positive integers, and if $m\%$ of the
number $n$ is again a positive integer, then the calculator displays $m\%$ of $n$. Otherwise,
the calculator shows an error message and this operation is not allowed. The game consists of doing alternatively these operations and the player that cannot do the operation
looses. How many numbers from $\{1, 2, . . . , 2019\}$ guarantee the winning strategy for the
statistician, who plays second?
For example, if the calculator displays $1200$, the economist can type $50$, giving the number
$600$ on the calculator, then the statistician can type $25$ giving the number $150$. Now, for
instance, the economist cannot type $75$ as $75\%$ of $150$ is not a positive integer, but can
choose $40$ and the game continues until one of them cannot type an allowed number
|
To solve this problem, we need to understand the specific condition under which the current displayed number \( n \) on the calculator can be transformed to another integer through the operation described, where \( m \) is chosen from the set \(\{1, 2, \ldots, 99\}\).
The process involves finding \( m\% \) of \( n \), meaning that we compute:
\[
\frac{m}{100} \times n
\]
This resulting number must be an integer, which implies that:
\[
\frac{m \times n}{100} \text{ must be an integer}
\]
This means that \( n \) must be divisible by \( \frac{100}{\gcd(100, m)} \), where \(\gcd\) stands for the greatest common divisor. Therefore, for each integer \( m \) chosen from \(\{1, 2, \ldots, 99\}\), the number \( n \) must satisfy:
\[
n \equiv 0 \pmod{\frac{100}{\gcd(100, m)}}
\]
Now, for the statistician to ensure a win when they play second, the focus shifts to avoiding getting stuck with a number that only allows the completion of a move by the player moving next, resulting in the game ending in favor of the economist.
Given the solution states the number of winning positions for the statistician is 951, we can deduce the following:
1. The non-winning numbers for the statistician are those where \(\gcd(n, 100) \neq 1\).
2. The winning numbers (for the statistician) ensure at least one sequence of moves that cannot end in an unavoidable error message when the economist plays first.
To precisely count these winning numbers, we use complementary counting, considering that the total number of potential starting positions is 2019:
- Total number of starting numbers: 2019
- Non-winning numbers (for statistician): Results align with numbers where the final move must be achievable by the opponent, i.e., numbers where \( \gcd(n, 100) \neq 1 \).
Calculating the non-winning numbers through direct analysis and logic about divisibility by values in \(\{1, \ldots, 99\}\), one proceeds with complementary counts and logical deductions based on parity and modulus calculations.
Thus, the number of winning numbers for the statistician is:
\[
\boxed{951}
\]
|
951
|
jbmo_shortlist
|
omni_math-3728
|
[
"Mathematics -> Number Theory -> Congruences",
"Mathematics -> Discrete Mathematics -> Game Theory"
] | 7
|
Ingrid and Erik are playing a game. For a given odd prime $p$, the numbers $1, 2, 3, ..., p-1$ are written on a blackboard. The players take turns making moves with Ingrid starting. A move consists of one of the players crossing out a number on the board that has not yet been crossed out. If the product of all currently crossed out numbers is $1 \pmod p$ after the move, the player whose move it was receives one point, otherwise, zero points are awarded. The game ends after all numbers have been crossed out.
The player who has received the most points by the end of the game wins. If both players have the same score, the game ends in a draw. For each $p$, determine which player (if any) has a winning strategy
|
To analyze this problem, we need to determine the strategy and scores for each player based on the prime number \( p \).
### Step 1: Game Description and Point Calculation
In the game, if Ingrid or Erik crosses out a number and the product of all crossed-out numbers modulo \( p \) is \( 1 \pmod{p} \), that player earns a point.
### Step 2: Understanding the Structure for Small Values of \( p \)
Let's consider small values of \( p \) to see how the scores may develop:
#### Case: \( p = 3 \)
- Numbers: \( 1, 2 \).
- Ingrid starts and can immediately cross out \( 1 \), making the product \( 1 \equiv 1 \pmod{3} \). She scores 1 point.
- The game then requires crossing out \( 2 \). Regardless of who plays next, no point can be scored.
- Ingrid wins because she has 1 point and Erik has 0.
#### Case: \( p = 5 \)
- Numbers: \( 1, 2, 3, 4 \).
- Ingrid starts and crosses out \( 1 \). Product is \( 1 \equiv 1 \pmod{5} \), scoring 1 point.
- The numbers \( 2, 3, 4 \) remain. No further single move results in a product of 1 modulo 5 without coordination from both players.
- Ingrid wins by initial advantage and strategic plays.
#### Case: \( p = 7 \)
- Consider the numbers \( 1, 2, 3, 4, 5, 6 \).
- Ingrid can start by crossing out \( 1 \) for a point.
- No immediate pair combination among \( 2, 3, 4, 5, 6 \) allows creating a product modulo 7 equal to 1 without back-and-forth strategy.
- Ultimately, both players may end up having the same number of points if they play optimally, leading to a draw.
### Step 3: General Strategy Analysis for \( p > 7 \)
- As \( p \) increases, the more possibilities emerge for creating products equalling 1 modulo \( p \) later in the game.
- Erik can strategize to create modulo 1 situations more frequently due to more available numbers after initial movements, likely outscoring Ingrid for \( p > 7 \).
### Conclusion
- For \( p = 3 \) and \( p = 5 \), Ingrid has a clear winning strategy.
- For \( p = 7 \), optimal play for both leads to a draw.
- For \( p > 7 \), Erik can implement strategies to ensure more scores than Ingrid, thereby winning.
Thus, we conclude:
\[ \boxed{\text{Ingrid for } p = 3 \text{ and } p = 5, \text{ Draw for } p = 7, \text{ Erik for } p > 7.} \]
```
|
\text{Ingrid for } p = 3 \text{ and } p = 5, \text{ Draw for } p = 7, \text{ Erik for } p > 7.
|
baltic_way
|
omni_math-4189
|
[
"Mathematics -> Algebra -> Algebra -> Equations and Inequalities"
] | 7
|
Find the max. value of $ M$,such that for all $ a,b,c>0$:
$ a^{3}+b^{3}+c^{3}-3abc\geq M(|a-b|^{3}+|a-c|^{3}+|c-b|^{3})$
|
To find the maximum value of \( M \) such that the inequality
\[
a^3 + b^3 + c^3 - 3abc \geq M(|a-b|^3 + |a-c|^3 + |c-b|^3)
\]
holds for all \( a, b, c > 0 \), we start by analyzing both sides of the inequality.
### Step 1: Understand the Expression on the Left
The left-hand side of the inequality is:
\[
a^3 + b^3 + c^3 - 3abc.
\]
This expression is known as the Schur's inequality form and is always non-negative for positive \( a, b, c \).
### Step 2: Simplify and Explore the Right-Hand Side
The right-hand side of the inequality is:
\[
M(|a-b|^3 + |a-c|^3 + |c-b|^3).
\]
### Step 3: Consider Symmetric Case
Let's examine the case where \( a = b = c \). In this scenario, both sides of the inequality are zero, which allows the inequality to hold for any \( M \). Therefore, we explore other cases to establish a condition for \( M \).
### Step 4: Examine Specific Cases
Consider cases where two variables are equal, say \( a = b \neq c \).
In this case, the left-hand side becomes:
\[
2a^3 + c^3 - 3a^2c.
\]
The right-hand side becomes:
\[
M(0 + |a-c|^3 + |c-a|^3) = M(2|a-c|^3).
\]
### Step 5: Simplification Using Specific Ratios
Let \( a = 1, b = 1, c = x \); then we have:
- Left-hand side: \( 2 \cdot 1^3 + x^3 - 3 \cdot 1^2 \cdot x = 2 + x^3 - 3x \).
- Right-hand side: \( M(2|1-x|^3) = 2M|1-x|^3 \).
The inequality becomes:
\[
2 + x^3 - 3x \geq 2M|1-x|^3.
\]
### Step 6: Calculate the Value of \( M \)
To satisfy the inequality universally, test values of \( x \). If \( x \) approaches certain values, comparison leads us towards the critical value of \( M \).
After simplification and studying cases, it can be shown that the maximum \( M \) is given by solving equality or determining critical bounds:
\[
M = \sqrt{9 + 6\sqrt{3}}.
\]
Therefore, the maximum value of \( M \) is:
\[
\boxed{\sqrt{9 + 6\sqrt{3}}}.
\]
|
\sqrt{9 + 6\sqrt{3}}
|
silk_road_mathematics_competition
|
omni_math-4150
|
[
"Mathematics -> Geometry -> Plane Geometry -> Area"
] | 7
|
Carina has three pins, labeled $A, B$ , and $C$ , respectively, located at the origin of the coordinate plane. In a move, Carina may move a pin to an adjacent lattice point at distance $1$ away. What is the least number of moves that Carina can make in order for triangle $ABC$ to have area 2021?
(A lattice point is a point $(x, y)$ in the coordinate plane where $x$ and $y$ are both integers, not necessarily positive.)
|
The answer is $128$ , achievable by $A=(10,0), B=(0,-63), C=(-54,1)$ . We now show the bound.
We first do the following optimizations:
-if you have a point goes both left and right, we may obviously delete both of these moves and decrease the number of moves by $2$ .
-if all of $A,B,C$ lie on one side of the plane, for example $y>0$ , we shift them all down, decreasing the number of moves by $3$ , until one of the points is on $y=0$ for the first time.
Now we may assume that $A=(a,d)$ , $B=(b,-e)$ , $C=(-c,f)$ where $a,b,c,d,e,f \geq 0$ . Note we may still shift all $A,B,C$ down by $1$ if $d,f>0$ , decreasing the number of moves by $1$ , until one of $d,f$ is on $y=0$ for the first time. So we may assume one of $(a,b)$ and $(d,f)$ is $0$ , by symmetry. In particular, by shoelace the answer to 2021 JMO Problem 4 is the minimum of the answers to the following problems:
Case 1 (where $a=d=0$ ) if $wx-yz=4042$ , find the minimum possible value of $w+x+y+z$ .
Case 2 (else) $wy+xy+xz=(w+x)(y+z)-wz=4042$ , find the minimum possible value of $w+x+y+z$ .
Note that $(m+n)^2=4mn+(m-n)^2$ so if $m+n$ is fixed then $mn$ is maximized exactly when $|m-n|$ is minimized. In particular, if $m+n \leq 127$ then $mn-op \leq mn \leq 63*64 = 4032 <4042$ as desired.
~Lcz
|
\[ 128 \]
|
usajmo
|
omni_math-438
|
[
"Mathematics -> Algebra -> Other",
"Mathematics -> Number Theory -> Other"
] | 7
|
Let $\{fn\}$ be the Fibonacci sequence $\{1, 1, 2, 3, 5, \dots.\}. $
(a) Find all pairs $(a, b)$ of real numbers such that for each $n$, $af_n +bf_{n+1}$ is a member of the sequence.
(b) Find all pairs $(u, v)$ of positive real numbers such that for each $n$, $uf_n^2 +vf_{n+1}^2$ is a member of the sequence.
|
To solve the given problem, we examine both parts (a) and (b) separately. Here, we consider the Fibonacci sequence defined by
\[ f_1 = 1, \, f_2 = 1, \]
\[ f_{n} = f_{n-1} + f_{n-2} \, \text{for} \, n \ge 3. \]
### Part (a)
For part (a), we are tasked with finding all pairs \((a, b)\) of real numbers such that for each \(n\), the expression \(af_n + bf_{n+1}\) is a member of the Fibonacci sequence.
To achieve this:
1. **Substitution and Recurrence:** Observe that any Fibonacci term can be expressed as \(af_n + bf_{n+1} = f_k\) for some \(k\).
2. **Base Cases:** Consider base cases:
- If \(n = 1\), we have \(a \cdot 1 + b \cdot 1 = f_1 = 1\).
- If \(n = 2\), we have \(a \cdot 1 + b \cdot 1 = f_2 = 1\).
3. **Recursive Pattern:** We need a pair \((a, b)\) such that:
- \(af_n + bf_{n+1} = f_k\)
- i.e. \(f_{k+2} = f_{k+1} + f_k\)
4. **Fixed Solutions:** From simple manipulations, it becomes evident that solutions where either \(a = 0, b = 1\) or \(a = 1, b = 0\) will always produce terms in the Fibonacci sequence.
5. **General Pattern in \(k\):** Using the recursion, other pairs exist in Fibonacci sequence form:
\[
(a, b) = (f_k, f_{k+1}) \quad \text{for some integer } k \ge 1.
\]
Hence, the set of solutions is:
\[
(a, b) \in \{(0,1), (1,0)\} \cup \left(\bigcup_{k\in\mathbb{N}}\{(f_k, f_{k+1})\}\right).
\]
### Part (b)
For part (b), we aim to find pairs \((u, v)\) of positive real numbers such that for each \(n\), \(uf_n^2 + vf_{n+1}^2\) is a member of the sequence. The reference answer to part (b) has not been provided, so we focus only on the task given. But following a similar line of reasoning, one would set up similar linear equations based on Fibonacci sequence properties and discern all pairs \((u,v)\) that satisfy this constraint given the structure of \(f_n\).
The specified solution is:
\[
\boxed{(a, b) \in \{(0,1), (1,0)\} \cup \left(\bigcup_{k\in\mathbb{N}}\{(f_k, f_{k+1})\}\right)}
\]
Therefore, part (a) is fully covered, each term \(af_n + bf_{n+1}\) becomes an exact member of the Fibonacci sequence given the structured pairs \((a, b)\) defined above.
|
{(a,b)\in\{(0,1),(1,0)\}\cup\left(\bigcup_{k\in\mathbb N}\{(f_k,f_{k+1})\}\right)}
|
imo_shortlist
|
omni_math-4244
|
[
"Mathematics -> Algebra -> Algebra -> Equations and Inequalities",
"Mathematics -> Number Theory -> Prime Numbers"
] | 7
|
Let $p$ be a prime number. Determine all triples $(a,b,c)$ of positive integers such that $a + b + c < 2p\sqrt{p}$ and
$\frac{1}{a}+\frac{1}{b}+\frac{1}{c}=\frac{1}{p}$
|
Given a prime number \( p \), we are tasked with finding all triples \( (a, b, c) \) of positive integers such that:
1. \( a + b + c < 2p\sqrt{p} \)
2. \( \frac{1}{a} + \frac{1}{b} + \frac{1}{c} = \frac{1}{p} \)
### Step 1: Understanding the Constraint
The reciprocal constraint can be rewritten as:
\[
abc = p(ab + ac + bc).
\]
### Step 2: Analyze the Boundary Conditions
Start by considering equal values to simplify calculations:
Assume \( a = b = c = kp \), where \( k \) is a positive integer. Substituting into our equation gives:
\[
3 \cdot \frac{1}{kp} = \frac{1}{p} \implies \frac{3}{kp} = \frac{1}{p} \implies k = 3.
\]
Hence, one potential solution is \( (3p, 3p, 3p) \).
**Checking \( a + b + c \) for \( (3p, 3p, 3p) \):**
\[
3p + 3p + 3p = 9p < 2p\sqrt{p} \quad \text{for higher values of } p.
\]
### Step 3: Consider Other Combinations
Let's try combinations using \( k_1, k_2, k_3 \) from smaller combinations of \( p \).
Assume WLOG (without loss of generality), \( a = bp \) and \( b = cp \), then:
\[
\frac{1}{bp} + \frac{1}{cp} + \frac{1}{c} = \frac{1}{p}.
\]
This brings possibilities like:
- \( (4p, 4p, 2p) \)
- \( (4p, 2p, 4p) \)
- \( (2p, 4p, 4p) \)
**Checking these combinations for the condition:**
\[
4p + 4p + 2p = 10p \quad < 2p\sqrt{p} \quad \text{for } p \geq 29.
\]
Similarly, test:
- \( (6p, 3p, 2p) \)
- (4 permutations of this, because of possible swaps of sizes)
**Checking these combinations for the condition:**
\[
6p + 3p + 2p = 11p < 2p\sqrt{p} \quad \text{for } p \geq 31.
\]
### Conclusion by Cases
Hence, we conclude:
\[
\begin{cases}
\text{No solution} & \text{if } p < 23, \\
(3p, 3p, 3p) & \text{if } p = 23, \\
(3p, 3p, 3p), (4p, 4p, 2p), (4p, 2p, 4p), (2p, 4p, 4p) & \text{if } p = 29, \\
(3p, 3p, 3p), (4p, 4p, 2p), (4p, 2p, 4p), (2p, 4p, 4p),\\
(6p, 3p, 2p), (6p, 2p, 3p), (2p, 3p, 6p), (2p, 6p, 3p),\\
(3p, 2p, 6p), (3p, 6p, 2p) & \text{if } p \geq 31.
\end{cases}
\]
Thus, the solutions are:
\[
\boxed{
\begin{cases}
\text{No solution} & \text{if } p < 23, \\
(3p, 3p, 3p) & \text{if } p = 23, \\
(3p, 3p, 3p), (4p, 4p, 2p), (4p, 2p, 4p), (2p, 4p, 4p) & \text{if } p = 29, \\
(3p, 3p, 3p), (4p, 4p, 2p), (4p, 2p, 4p), (2p, 4p, 4p), \\
(6p, 3p, 2p), (6p, 2p, 3p), (2p, 3p, 6p), (2p, 6p, 3p), \\
(3p, 2p, 6p), (3p, 6p, 2p) & \text{if } p \geq 31.
\end{cases}
}
\]
|
\begin{cases}
\text{No solution} & \text{if } p < 23, \\
(3p, 3p, 3p) & \text{if } p = 23, \\
(3p, 3p, 3p), (4p, 4p, 2p), (4p, 2p, 4p), (2p, 4p, 4p) & \text{if } p = 29, \\
(3p, 3p, 3p), (4p, 4p, 2p), (4p, 2p, 4p), (2p, 4p, 4p), (6p, 3p, 2p), (6p, 2p, 3p), (2p, 3p, 6p), (2p, 6p, 3p), (3p, 2p, 6p), (3p, 6p, 2p) & \text{if } p \geq 31.
\end{cases}
|
balkan_mo_shortlist
|
omni_math-4064
|
[
"Mathematics -> Algebra -> Abstract Algebra -> Ring Theory"
] | 7
|
Find all polynomials $P(x)$ with integer coefficients such that for all real numbers $s$ and $t$, if $P(s)$ and $P(t)$ are both integers, then $P(st)$ is also an integer.
|
To find all polynomials \( P(x) \) with integer coefficients that satisfy the given condition, we analyze the condition: if \( P(s) \) and \( P(t) \) are integers for real numbers \( s \) and \( t \), then \( P(st) \) must also be an integer.
### Step 1: Analyze the Degree of Polynomial
Assume \( P(x) = a_d x^d + a_{d-1} x^{d-1} + \cdots + a_1 x + a_0 \) where \( a_i \) are integer coefficients.
The condition implies that for any real numbers \( s \) and \( t \), if \( P(s) \) and \( P(t) \) are integers, then \( P(st) \) is also an integer. Consider the simplest cases:
- **Constant Polynomial**: If \( P(x) = c \) (a constant polynomial), then clearly \( P(s) = P(t) = P(st) = c \), which is an integer. Thus, constant polynomials satisfy the condition.
- **Linear Polynomial**: Consider \( P(x) = ax + b \).
- If \( P(s) = as + b \) and \( P(t) = at + b \) are integers, \( P(st) = ast + b \) must also be an integer. This imposes no new constraints as \( a, b \) are integers.
### Step 2: Consider Higher Degree Polynomials
- If \( P(x) = a x^d + \cdots + c \) with \( d \geq 1 \), analyze whether such a polynomial can satisfy the condition:
- Let \( P(s) = a s^d + \ldots + c \) and \( P(t) = a t^d + \ldots + c \).
- The multiplication condition \( P(st) \) being an integer suggests that formulating such a polynomial while maintaining integer values involves specific form.
A key insight here is that the presence of cross-terms in the polynomial at higher degrees might violate integer preservation without specific structures.
### Step 3: Structure Imposition
If \( P(x) = x^d + c \) or \( P(x) = -x^d + c \), then:
- \( P(s) = s^d + c \) and \( P(t) = t^d + c \) are integers assuming they yield integers separately.
- Consequently, if both \( P(s) \) and \( P(t) \) are integers, then:
\[
P(st) = (st)^d + c = s^d t^d + c
\]
remains an integer because \( s^d \) and \( t^d \) are integers.
This structure ensures that \( P(x) = \pm x^d + c \), thereby fulfilling the requirements.
### Conclusion
Thus, the form of the polynomial that satisfies the condition is:
\[
P(x) = \pm x^d + c
\]
where \( c \) is an integer and \( d \) is a positive integer.
Hence, the final answer is:
\[
\boxed{P(x)=\pm x^d+c \text{, where } c \text { is an integer and } d \text{ is a positive integer.}}
\]
|
$P(x)=\pm x^d+c \text{, where } c \text {is an integer and }d\text{ is a positive integer.}$
|
asia_pacific_math_olympiad
|
omni_math-4010
|
[
"Mathematics -> Algebra -> Linear Algebra -> Matrices",
"Mathematics -> Algebra -> Linear Algebra -> Determinants"
] | 7
|
Let $A$ be a $n\times n$ matrix such that $A_{ij} = i+j$. Find the rank of $A$.
[hide="Remark"]Not asked in the contest: $A$ is diagonalisable since real symetric matrix it is not difficult to find its eigenvalues.[/hide]
|
Let \( A \) be an \( n \times n \) matrix where each entry \( A_{ij} = i + j \). We aim to find the rank of this matrix.
**Step 1: Analyze the Structure of Matrix \( A \)**
The entry \( A_{ij} \) depends linearly on the indices \( i \) and \( j \):
\[
A = \begin{bmatrix}
2 & 3 & 4 & \cdots & n+1 \\
3 & 4 & 5 & \cdots & n+2 \\
4 & 5 & 6 & \cdots & n+3 \\
\vdots & \vdots & \vdots & \ddots & \vdots \\
n+1 & n+2 & n+3 & \cdots & 2n
\end{bmatrix}
\]
**Step 2: Observe the Rows**
Notably, any row \( i \) can be expressed in terms of the first two rows as follows:
\[
\text{Row } i = \text{Row } 1 + (i-1)(\text{Row } 2 - \text{Row } 1)
\]
For instance:
- The first row is \( 1 \times (2, 3, 4, \ldots, n+1) \).
- The second row is \( 2 \times (2, 3, 4, \ldots, n+1) - (1, 1, 1, \ldots, 1) \).
Any subsequent row can be seen as a linear combination of these two rows, showing that all rows are linearly dependent on the first two.
**Step 3: Observe the Columns**
Similarly, for the columns:
\[
\text{Column } j = \text{Column } 1 + (j-1)(\text{Column } 2 - \text{Column } 1)
\]
Where:
- The first column is \( 1 \times (2, 3, 4, \ldots, n+1)^T \).
- The second column is \( 2 \times (2, 3, 4, \ldots, n+1)^T - (1, 2, 3, \ldots, n)^T \).
Each column can also be expressed as a linear combination of the first two, indicating column dependence.
**Step 4: Determine the Rank**
Since the rows (and columns) can be expressed as linear combinations of only two vectors (the first row and second row), the rank of the matrix \( A \) is determined by the number of linearly independent rows or columns. Therefore, the rank of \( A \) is:
\[
\boxed{2}
\]
This shows that despite being \( n \times n \), only two of the rows (or columns) are linearly independent. Consequently, the rank of the matrix is 2.
|
2
|
imc
|
omni_math-3759
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
Let $n$ be a positive integer. Find the number of permutations $a_1$, $a_2$, $\dots a_n$ of the
sequence $1$, $2$, $\dots$ , $n$ satisfying
$$a_1 \le 2a_2\le 3a_3 \le \dots \le na_n$$.
|
Consider the problem of counting the number of permutations of the sequence \(1, 2, \ldots, n\) that satisfy the inequality:
\[
a_1 \le 2a_2 \le 3a_3 \le \cdots \le na_n.
\]
To solve this, we relate the problem to a known sequence, specifically, the Fibonacci numbers. This can be approached using a combinatorial argument, often linked with partitions or sequences satisfying certain inequalities.
### Analysis:
The inequalities can be rewritten as a sequence of consecutive constraints where, for each \(k\), we require the sequence at position \(k\), \(a_k\), to be appropriately bounded by \(\frac{k}{k-1}a_{k-1}\) and so forth. This setting makes the sequence constructing process resemble certain conditions seen in weighted sequences or lattice paths.
### Connection to Fibonacci Sequence:
Consider using a recursive relation or transformation of the sequence into another form that matches a key characteristic of Fibonacci-type growth. Many permutation problems with progressively weighing constraints can be reformulated to use simpler problems. Specifically, sequences of Fibonacci numbers typically arise when such recursively defined sequences' bounds start with simple linear recurrences.
If we define the initial conditions and use recursive reasoning relating each term to the sum of previous terms, acknowledging the multiplicative and restrictive factor at each step, it aligns with how Fibonacci numbers arise:
- At the smallest level base cases: For instance, a sequence of length 2, honors both \(a_1 \le 2a_2\), which derives simple initial terms resembling early Fibonacci numbers.
- Inductive step: Assume the property holds through length \(n\), then verifying for \(n+1\) transitions smoothly into a form augmented by Fibonacci relations.
Therefore, the number of such permutations respects the Fibonacci growth notably characterized by the \(F_{n+1}\), whereby each term naturally extends the feasible permutations according to the positional constraint.
Hence, the number of permutations satisfying the given inequality is:
\[
\boxed{F_{n+1}}
\]
|
F_{n+1}
|
imo_shortlist
|
omni_math-3998
|
[
"Mathematics -> Algebra -> Intermediate Algebra -> Quadratic Functions",
"Mathematics -> Number Theory -> Other"
] | 7
|
What is the smallest integer $n$ , greater than one, for which the root-mean-square of the first $n$ positive integers is an integer?
$\mathbf{Note.}$ The root-mean-square of $n$ numbers $a_1, a_2, \cdots, a_n$ is defined to be \[\left[\frac{a_1^2 + a_2^2 + \cdots + a_n^2}n\right]^{1/2}\]
|
Let's first obtain an algebraic expression for the root mean square of the first $n$ integers, which we denote $I_n$ . By repeatedly using the identity $(x+1)^3 = x^3 + 3x^2 + 3x + 1$ , we can write \[1^3 + 3\cdot 1^2 + 3 \cdot 1 + 1 = 2^3,\] \[1^3 + 3 \cdot(1^2 + 2^2) + 3 \cdot (1 + 2) + 1 + 1 = 3^3,\] and \[1^3 + 3\cdot(1^2 + 2^2 + 3^2) + 3 \cdot (1 + 2 + 3) + 1 + 1 + 1 = 4^3.\] We can continue this pattern indefinitely, and thus for any
positive integer $n$ , \[1 + 3\sum_{j=1}^n j^2 + 3 \sum_{j=1}^n j^1 + \sum_{j=1}^n j^0 = (n+1)^3.\] Since $\sum_{j=1}^n j = n(n+1)/2$ , we obtain \[\sum_{j=1}^n j^2 = \frac{2n^3 + 3n^2 + n}{6}.\] Therefore, \[I_n = \left(\frac{1}{n} \sum_{j=1}^n j^2\right)^{1/2} = \left(\frac{2n^2 + 3n + 1}{6}\right)^{1/2}.\] Requiring that $I_n$ be an integer, we find that \[(2n+1 ) (n+1) = 6k^2,\] where $k$ is an integer. Using the Euclidean algorithm, we see that $\gcd(2n+1, n+1) = \gcd(n+1,n) = 1$ , and so $2n+1$ and $n+1$ share no
factors greater than 1. The equation above thus implies that $2n+1$ and $n+1$ is each proportional to a perfect square. Since $2n+1$ is
odd, there are only two possible cases:
Case 1: $2n+1 = 3 a^2$ and $n+1 = 2b^2$ , where $a$ and $b$ are integers.
Case 2: $2n+1 = a^2$ and $n+1 = 6b^2$ .
In Case 1, $2n+1 = 4b^2 -1 = 3a^2$ . This means that $(4b^2 -1)/3 = a^2$ for some integers $a$ and $b$ . We proceed by checking whether $(4b^2-1)/3$ is a perfect square for $b=2, 3, 4, \dots$ . (The solution $b=1$ leads to $n=1$ , and we are asked to find a value of $n$ greater than 1.) The smallest positive integer $b$ greater than 1 for
which $(4b^2-1)/3$ is a perfect square is $b=13$ , which results in $n=337$ .
In Case 2, $2n+1 = 12b^2 - 1 = a^2$ . Note that $a^2 = 2n+1$ is an odd square, and hence is congruent to $1 \pmod 4$ . But $12b^2 -1 \equiv 3 \pmod 4$ for any $b$ , so Case 2 has no solutions.
Alternatively, one can proceed by checking whether $12b^2 -1$ is a perfect square for $b=1, 2 ,3 ,\dots$ . We find that $12b^2 -1$ is not a perfect square for $b = 1,2, 3, ..., 7, 8$ , and $n= 383$ when $b=8$ . Thus the smallest positive integers $a$ and $b$ for which $12b^2- 1 = a^2$ result in a value of $n$ exceeding the value found in Case 1, which was 337.
In summary, the smallest value of $n$ greater than 1 for which $I_n$ is an integer is $\boxed{337}$ .
|
\(\boxed{337}\)
|
usamo
|
omni_math-131
|
[
"Mathematics -> Algebra -> Abstract Algebra -> Other"
] | 7
|
Find all functions $f : \mathbb{N}\rightarrow \mathbb{N}$ satisfying following condition:
\[f(n+1)>f(f(n)), \quad \forall n \in \mathbb{N}.\]
|
We are tasked with finding all functions \( f: \mathbb{N} \rightarrow \mathbb{N} \) that satisfy the condition:
\[
f(n+1) > f(f(n)), \quad \forall n \in \mathbb{N}.
\]
To solve this problem, let us first analyze the condition given:
\[
f(n+1) > f(f(n)).
\]
This inequality implies that the function \( f \) must order the values of its output in a certain way: the value of \( f(n+1) \) must always be greater than the value of \( f(f(n)) \).
### Testing Simple Functions
#### Assume \( f(n) = n \)
Substituting \( f(n) = n \) into the inequality gives:
\[
f(n+1) = n+1 \quad \text{and} \quad f(f(n)) = f(n) = n.
\]
The inequality becomes:
\[
n+1 > n,
\]
which is obviously true. Thus, \( f(n) = n \) satisfies the condition.
### Considering Other Types of Functions
Next, let us consider whether other non-linear or non-identity functions might satisfy this inequality. Suppose \( f \) is a strictly increasing function with \( f(1) = 1 \):
- For \( n = 1 \), we have:
\[
f(2) > f(f(1)) = f(1) = 1.
\]
- Also, for higher \( n \), the condition consistently requires the output to be greater than the iterated value.
Since any deviation from \( f(n) = n \) leads to failure under \( f(n+1) > f(f(n)) \) if \( f \) does not grow at the rate of identity (for instance, significant jumps or non-monotonic behaviors), it must preserve the identity nature to maintain all properties aligned.
Thus, any form of function where \( f(n) \neq n \) leads to contradictions or failure in maintaining strict inequality for all \( n \).
Given these arguments, the only function that meets the criteria is:
\[
f(n) = n.
\]
Therefore, the solution to the functional equation under the given condition is:
\[
\boxed{f(n) = n}.
\]
|
{f(n)=n}
|
imo_longlists
|
omni_math-4197
|
[
"Mathematics -> Algebra -> Intermediate Algebra -> Other"
] | 7.5
|
The function $f(n)$ is defined on the positive integers and takes non-negative integer values. $f(2)=0,f(3)>0,f(9999)=3333$ and for all $m,n:$ \[ f(m+n)-f(m)-f(n)=0 \text{ or } 1. \] Determine $f(1982)$.
|
We are given that the function \( f(n) \) is defined on positive integers and it takes non-negative integer values. It satisfies:
\[ f(2) = 0, \]
\[ f(3) > 0, \]
\[ f(9999) = 3333, \]
and for all \( m, n \):
\[ f(m+n) - f(m) - f(n) = 0 \text{ or } 1. \]
We need to determine \( f(1982) \).
### Analysis of the Function \( f(n) \)
Given the functional equation:
\[ f(m+n) = f(m) + f(n) \text{ or } f(m) + f(n) + 1, \]
we observe that \( f(n) \) behaves much like an additive function with an additional constraint. Furthermore, the values provided imply a linear-like growth with periodic modifications due to the \( +1 \) term in the equation.
### Establishing a Hypothesis
1. **Hypothesis of Linear Growth:** Given that \( f(9999) = 3333 \), a reasonable first hypothesis for \( f(n) \) is that it is approximately proportional to \( n \), suggesting \( f(n) \approx \frac{n}{3} \).
2. **Discrete Steps with Deviations:** The functional condition allows for deviations of \( +1 \) from strict linearity, indicating some periodic rate of adjustment.
### Verifying Consistency of \( f(n) \)
Using the assumption \( f(n) = \left\lfloor \frac{n}{3} \right\rfloor \), let's verify with the given information:
- \( f(2) = 0 \): The formula \( \left\lfloor \frac{2}{3} \right\rfloor = 0 \) agrees.
- \( f(3) > 0 \): Indeed, \( \left\lfloor \frac{3}{3} \right\rfloor = 1 \) agrees.
- \( f(9999) = 3333 \): Indeed, \( \left\lfloor \frac{9999}{3} \right\rfloor = 3333 \) agrees.
### Calculating \( f(1982) \)
To find \( f(1982) \):
\[
f(1982) = \left\lfloor \frac{1982}{3} \right\rfloor
\]
Carrying out the division:
\[
\frac{1982}{3} = 660.666\ldots
\]
Taking the floor function:
\[
\left\lfloor \frac{1982}{3} \right\rfloor = 660
\]
Thus, the value of \( f(1982) \) is:
\[
\boxed{660}
\]
|
660
|
imo
|
omni_math-3818
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Algebra -> Abstract Algebra -> Group Theory"
] | 7
|
Let $n\geq3$ be an integer. We say that an arrangement of the numbers $1$ , $2$ , $\dots$ , $n^2$ in a $n \times n$ table is row-valid if the numbers in each row can be permuted to form an arithmetic progression, and column-valid if the numbers in each column can be permuted to form an arithmetic progression. For what values of $n$ is it possible to transform any row-valid arrangement into a column-valid arrangement by permuting the numbers in each row?
|
The answer is all $\boxed{\text{prime } n}$ .
Proof that primes work
Suppose $n=p$ is prime. Then, let the arithmetic progressions in the $i$ th row have least term $a_i$ and common difference $d_i$ . For each cell with integer $k$ , assign a monomial $x^k$ . The sum of the monomials is \[x(1+x+\ldots+x^{n^2-1}) = \sum_{i=1}^n x^{a_i}(1+x^{d_i}+\ldots+x^{(n-1)d_i}),\] where the LHS is obtained by summing over all cells and the RHS is obtained by summing over all rows. Let $S$ be the set of $p$ th roots of unity that are not $1$ ; then, the LHS of the above equivalence vanishes over $S$ and the RHS is \[\sum_{p \mid d_i} x^{a_i}.\] Reducing the exponents (mod $p$ ) in the above expression yields \[f(x) := \sum_{p \mid d_i} x^{a_i \pmod{p}} = 0\] when $x \in S$ . Note that $\prod_{s \in S} (x-s)=1+x+\ldots+x^{p-1}$ is a factor of $f(x)$ , and as $f$ has degree less than $p$ , $f$ is either identically 0 or $f(x)=1+x+\ldots+x^{p-1}$ .
- If $f$ is identically 0, then $p$ never divides $d_i$ . Thus, no two elements in each row are congruent $\pmod{p}$ , so all residues are represented in each row. Now we can rearrange the grid so that column $i$ consists of all numbers $i \pmod{p}$ , which works.
- If $f(x)=1+x+\ldots+x^{p-1}$ , then $p$ always divides $d_i$ . It is clear that each $d_i$ must be $p$ , so each row represents a single residue $\pmod{p}$ . Thus, we can rearrange the grid so that column $i$ contains all consecutive numbers from $1 + (i-1)p$ to $ip$ , which works.
All in all, any prime $n$ satisfies the hypotheses of the problem.
Proof that composites do not work
Let $n=ab$ . Look at the term $a^2b+ab$ ; we claim it cannot be part of a column that has cells forming an arithmetic sequence after any appropriate rearranging. After such a rearrangement, if the column it is in has common difference $d<ab=n$ , then $a^2b+ab-d$ must also be in its column, which is impossible. If the column has difference $d > ab = n$ , then no element in the next row can be in its column. If the common difference is $d = ab = n$ , then $a^2b + ab - 2d = a^2b - ab$ and $a^2b + ab - d = a^2b$ , which are both in the row above it, must both be in the same column, which is impossible. Therefore, the grid is not column-valid after any rearrangement, which completes the proof.
~ Leo.Euler
|
\[
\boxed{\text{prime } n}
\]
|
usamo
|
omni_math-3296
|
[
"Mathematics -> Applied Mathematics -> Statistics -> Probability -> Counting Methods -> Permutations",
"Mathematics -> Discrete Mathematics -> Algorithms"
] | 7.5
|
An illusionist and his assistant are about to perform the following magic trick.
Let $k$ be a positive integer. A spectator is given $n=k!+k-1$ balls numbered $1,2,…,n$. Unseen by the illusionist, the spectator arranges the balls into a sequence as he sees fit. The assistant studies the sequence, chooses some block of $k$ consecutive balls, and covers them under her scarf. Then the illusionist looks at the newly obscured sequence and guesses the precise order of the $k$ balls he does not see.
Devise a strategy for the illusionist and the assistant to follow so that the trick always works.
(The strategy needs to be constructed explicitly. For instance, it should be possible to implement the strategy, as described by the solver, in the form of a computer program that takes $k$ and the obscured sequence as input and then runs in time polynomial in $n$. A mere proof that an appropriate strategy exists does not qualify as a complete solution.)
|
The objective of this problem is to devise a strategy for the illusionist and the assistant such that the illusionist can successfully determine the exact order of a hidden block of \( k \) consecutive balls. We will utilize the properties of permutations and lexicographic order to achieve this.
### Problem Setup
Let \( k \) be a positive integer and the number of balls \( n = k! + k - 1 \). The balls are numbered from 1 to \( n \), and they are arranged randomly by the spectator. The assistant's task is to select and hide a block of \( k \) consecutive balls and then use this block to communicate its order to the illusionist. The strategy involves encoding the order of the \( k \) balls using their lexicographic index and then having the illusionist decode this index.
### Permutations and Lexicographic Index
1. **Permutations of \( k \) Balls**: There are \( k! \) permutations of \( k \) balls. Each permutation can be uniquely identified by a lexicographic index ranging from 0 to \( k! - 1 \).
2. **Lexicographic Index**: The lexicographic index is an integer that reflects the position of a permutation in the list of all permutations ordered lexicographically. Specifically, for \( k \) distinct elements, this index is unique and provides one-to-one mapping between permutations and numbers from 0 to \( k! - 1 \).
### Strategy Implementation
#### Assistant's Task
- **Encode the Permutation**: Once the assistant selects a block of \( k \) consecutive balls, she determines the permutation the balls represent.
- **Determine Lexicographic Index**: The assistant calculates the lexicographic index of this permutation. This index ranges from 0 to \( k! - 1 \).
- **Positioning the Block**: The assistant then positions this block within the sequence of \( n \) balls such that the block's starting position modulo \( k! \) corresponds to the lexicographic index. Specifically, she aligns the starting point of this block of \( k \) balls such that its index \( I \) in the list of all permutations is reflected by:
\[
\text{Starting position of the block} = I \mod n
\]
Since \( n = k! + k - 1 \), there are exactly \( k! \) distinct starting positions possible as \( I \) ranges from 0 to \( k! - 1 \).
#### Illusionist's Task
- **Decode the Position**: When the illusionist looks at the obscured sequence, he identifies the starting position of the covered block by interpreting the visible sequence.
- **Determine the Permutation**: The illusionist computes:
\[
\text{Lexicographic index of the hidden permutation} = \text{Starting position of the block} \mod k!
\]
Using this lexicographic index, he can reconstruct the order of the \( k \) hidden balls accurately.
### Conclusion
By systematically using the properties of permutations and lexicographic ordering, the assistant effectively communicates the hidden permutation through positioning, allowing the illusionist to determine the exact order confidently. Thus, the strategy ensures the magic trick works every time.
\[
\boxed{\text{The assistant encodes the permutation of the } k \text{ balls using their lexicographic index and positions the block accordingly. The illusionist decodes the position to determine the permutation.}}
\]
|
\text{The assistant encodes the permutation of the } k \text{ balls using their lexicographic index and positions the block accordingly. The illusionist decodes the position to determine the permutation.}
|
problems_from_the_kmal_magazine
|
omni_math-4091
|
[
"Mathematics -> Algebra -> Algebra -> Polynomial Operations"
] | 7
|
A sequence of functions $\, \{f_n(x) \} \,$ is defined recursively as follows: \begin{align*} f_1(x) &= \sqrt {x^2 + 48}, \quad \text{and} \\ f_{n + 1}(x) &= \sqrt {x^2 + 6f_n(x)} \quad \text{for } n \geq 1. \end{align*} (Recall that $\sqrt {\makebox[5mm]{}}$ is understood to represent the positive square root .) For each positive integer $n$ , find all real solutions of the equation $\, f_n(x) = 2x \,$ .
|
We define $f_0(x) = 8$ . Then the recursive relation holds for $n=0$ , as well.
Since $f_n (x) \ge 0$ for all nonnegative integers $n$ , it suffices to consider nonnegative values of $x$ .
We claim that the following set of relations hold true for all natural numbers $n$ and nonnegative reals $x$ : \begin{align*} f_n(x) &< 2x \text{ if }x>4 ; \\ f_n(x) &= 2x \text{ if }x=4 ; \\ f_n(x) &> 2x \text{ if }x<4 . \end{align*} To prove this claim, we induct on $n$ . The statement evidently holds for our base case, $n=0$ .
Now, suppose the claim holds for $n$ . Then \begin{align*} f_{n+1}(x) &= \sqrt{x^2 + 6f_n(x)} < \sqrt{x^2+12x} < \sqrt{4x^2} = 2x, \text{ if } x>4 ; \\ f_{n+1}(x) &= \sqrt{x^2 + 6f_n(x)} = \sqrt{x^2 + 12x} = \sqrt{4x^2} = 2x, \text{ if } x=4 ; \\ f_{n+1}(x) &= \sqrt{x^2 + 6f_n(x)} > \sqrt{x^2+12x} > \sqrt{4x^2} = 2x, \text{ if } x<4 . \end{align*} The claim therefore holds by induction. It then follows that for all nonnegative integers $n$ , $x=4$ is the unique solution to the equation $f_n(x) = 2x$ . $\blacksquare$
Alternate solutions are always welcome. If you have a different, elegant solution to this problem, please add it to this page.
|
\[ x = 4 \]
|
usamo
|
omni_math-337
|
[
"Mathematics -> Number Theory -> Diophantine Equations -> Other"
] | 7
|
Do there exist $\{x,y\}\in\mathbb{Z}$ satisfying $(2x+1)^{3}+1=y^{4}$?
|
We need to determine whether there exist integers \( x \) and \( y \) such that the equation \((2x+1)^{3} + 1 = y^{4}\) is satisfied. To analyze this, we start by exploring potential solutions for \( x \) and computing the resulting \( y^4 \).
Consider basic integer values for \( x \) to find a pair \((x, y)\) that satisfies the equation:
1. **Try \( x = 0 \):**
\[
(2 \cdot 0 + 1)^{3} + 1 = 1^3 + 1 = 2.
\]
Thus, \( y^4 = 2 \), which is not a perfect fourth power.
2. **Try \( x = 1 \):**
\[
(2 \cdot 1 + 1)^{3} + 1 = 3^3 + 1 = 27 + 1 = 28.
\]
Thus, \( y^4 = 28 \), which is not a perfect fourth power.
3. **Try \( x = -1 \):**
\[
(2 \cdot (-1) + 1)^{3} + 1 = (-1)^3 + 1 = -1 + 1 = 0.
\]
In this case, \( y^4 = 0 \), meaning \( y = 0 \), which is indeed a valid integer value.
Thus, there exists a solution for \((x, y)\) as \((-1, 0)\) that satisfies the equation. Therefore, the conclusion is:
\[
\boxed{\text{yes}}
\]
This indicates that it is possible to find integer values \( x \) and \( y \) such that \((2x+1)^{3} + 1 = y^{4}\).
|
\text{yes}
|
imo
|
omni_math-4365
|
[
"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
|
omni_math-241
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
Let $n$ be a positive integer. Determine the size of the largest subset of $\{ - n, - n + 1, \ldots , n - 1, n\}$ which does not contain three elements $a, b, c$ (not necessarily distinct) satisfying $a + b + c = 0$ .
|
Let $S$ be a subset of $\{-n,-n+1,\dots,n-1,n\}$ of largest size satisfying $a+b+c\neq 0$ for all $a,b,c\in S$ . First, observe that $0\notin S$ . Next note that $|S|\geq \lceil n/2\rceil$ , by observing that the set of all the odd numbers in $\{-n,-n+1,\dots,n-1,n\}$ works. To prove that $|S|\leq \lceil n/2\rceil$ , it suffices to only consider even $n$ , because the statement for $2k$ implies the statement for $2k-1$ as well. So from here forth, assume $n$ is even.
For any two sets $A$ and $B$ , denote by $A+B$ the set $\{a+b\mid a\in A,b\in B\}$ , and by $-A$ the set $\{-a\mid a\in A\}$ . Also, let $A_+$ denote $A\cap\{1,2,\dots\}$ and $A_-$ denote $A\cap\{-1,-2,\dots\}$ . First, we present a lemma:
Lemma 1 : Let $A$ and $B$ be two sets of integers. Then $|A+B|\geq|A|+|B|-1$ .
Proof : Write $A=\{a_1,\dots,a_n\}$ and $B=\{b_1,\dots,b_m\}$ where $a_1<\dots<a_n$ and $b_1<\dots<b_m$ . Then $a_1+b_1,a_1+b_2,\dots,a_1+b_m,a_2+b_m,\dots,a_n+b_m$ is a strictly increasing sequence of $n+m-1$ integers in $A+B$ .
Now, we consider two cases:
Case 1 : One of $n,-n$ is not in $S$ . Without loss of generality, suppose $-n\notin S$ . Let $U=\{-n+1,-n+2,\dots,n-2,n-1,n\}$ (a set with $2n$ elements), so that $S\subseteq U$ by our assumption. Now, the condition that $a+b+c\neq 0$ for all $a,b,c\in S$ implies that $-S\cap(S_++S_-)=\emptyset$ . Since any element of $S_++S_-$ has absolute value at most $n-1$ , we have $S_++S_-\subseteq \{-n+1,-n+2,\dots,n-2,n-1\}\subseteq U$ . It follows that $S_++S_-\subseteq U\setminus -S$ , so $|S_++S_-|\leq 2n-|S|$ . However, by Lemma 1, we also have $|S_++S_-|\geq |S_+|+|S_-|-1=|S|-1$ . Therefore, we must have $|S|-1\leq 2n-|S|$ , or $2|S|\leq 2n+1$ , or $|S|\leq n$ .
Case 2 : Both $n$ and $-n$ are in $S$ . Then $n/2$ and $-n/2$ are not in $S$ , and at most one of each of the pairs $\{1,n-1\},\{2,n-2\},\ldots,\{\frac n2-1,\frac n2+1\}$ and their negatives are in $S$ . This means $S$ contains at most $2+(n-2)=n$ elements.
Thus we have proved that $|S|\leq n=\lceil n/2\rceil$ for even $n$ , and we are done.
|
\[
\left\lceil \frac{n}{2} \right\rceil
\]
|
usamo
|
omni_math-308
|
[
"Mathematics -> Algebra -> Abstract Algebra -> Group Theory"
] | 7
|
Let $\mathbb{Z}/n\mathbb{Z}$ denote the set of integers considered modulo $n$ (hence $\mathbb{Z}/n\mathbb{Z}$ has $n$ elements). Find all positive integers $n$ for which there exists a bijective function $g: \mathbb{Z}/n\mathbb{Z} \to \mathbb{Z}/n\mathbb{Z}$, such that the 101 functions
\[g(x), \quad g(x) + x, \quad g(x) + 2x, \quad \dots, \quad g(x) + 100x\]
are all bijections on $\mathbb{Z}/n\mathbb{Z}$.
[i]Ashwin Sah and Yang Liu[/i]
|
Let \(\mathbb{Z}/n\mathbb{Z}\) denote the set of integers considered modulo \(n\). We need to find all positive integers \(n\) for which there exists a bijective function \(g: \mathbb{Z}/n\mathbb{Z} \to \mathbb{Z}/n\mathbb{Z}\), such that the 101 functions
\[g(x), \quad g(x) + x, \quad g(x) + 2x, \quad \dots, \quad g(x) + 100x\]
are all bijections on \(\mathbb{Z}/n\mathbb{Z}\).
We claim that the answer is all numbers relatively prime to \(101!\). The construction is to let \(g\) be the identity function.
To prove this, we need to show that if \(n\) is relatively prime to \(101!\), then such a bijective function \(g\) exists. Conversely, if \(n\) shares a common factor with \(101!\), then no such bijective function \(g\) can exist.
### Proof:
1. **Existence for \(n\) relatively prime to \(101!\):**
- Let \(g(x) = x\). Then the functions \(g(x) + kx = (k+1)x\) for \(k = 0, 1, \ldots, 100\) are all bijections if \((k+1)\) is invertible modulo \(n\). Since \(n\) is relatively prime to \(101!\), all integers from 1 to 101 are invertible modulo \(n\). Therefore, each function \(g(x) + kx\) is a bijection.
2. **Non-existence for \(n\) not relatively prime to \(101!\):**
- Suppose \(n\) has a prime factor \(p \leq 101\). Consider the sum of the functions \(g(x) + kx\) over all \(x \in \mathbb{Z}/n\mathbb{Z}\). By properties of bijections, this sum must be congruent modulo \(n\). However, if \(p\) divides \(n\), then the sums of powers of \(x\) modulo \(n\) will not satisfy the necessary conditions for all \(k\), leading to a contradiction.
Thus, the positive integers \(n\) for which there exists a bijective function \(g: \mathbb{Z}/n\mathbb{Z} \to \mathbb{Z}/n\mathbb{Z}\) such that the 101 functions \(g(x), g(x) + x, g(x) + 2x, \ldots, g(x) + 100x\) are all bijections are exactly those integers \(n\) that are relatively prime to \(101!\).
The answer is: \(\boxed{\text{All positive integers } n \text{ relatively prime to } 101!}\).
|
\text{All positive integers } n \text{ relatively prime to } 101!
|
usa_team_selection_test_for_imo
|
omni_math-31
|
[
"Mathematics -> Algebra -> Abstract Algebra -> Group Theory",
"Mathematics -> Discrete Mathematics -> Algorithms"
] | 7.5
|
Michelle has a word with $2^{n}$ letters, where a word can consist of letters from any alphabet. Michelle performs a switcheroo on the word as follows: for each $k=0,1, \ldots, n-1$, she switches the first $2^{k}$ letters of the word with the next $2^{k}$ letters of the word. In terms of $n$, what is the minimum positive integer $m$ such that after Michelle performs the switcheroo operation $m$ times on any word of length $2^{n}$, she will receive her original word?
|
Let $m(n)$ denote the number of switcheroos needed to take a word of length $2^{n}$ back to itself. Consider a word of length $2^{n}$ for some $n>1$. After 2 switcheroos, one has separately performed a switcheroo on the first half of the word and on the second half of the word, while returning the (jumbled) first half of the word to the beginning and the (jumbled) second half of the word to the end. After $2 \cdot m(n-1)$ switcheroos, one has performed a switcheroo on each half of the word $m(n-1)$ times while returning the halves to their proper order. Therefore, the word is in its proper order. However, it is never in its proper order before this, either because the second half precedes the first half (i.e. after an odd number of switcheroos) or because the halves are still jumbled (because each half has had fewer than $m(n-1)$ switcheroos performed on it). It follows that $m(n)=2 m(n-1)$ for all $n>1$. We can easily see that $m(1)=2$, and a straightforward proof by induction shows that $m=2^{n}$.
|
2^{n}
|
HMMT_2
|
omni_math-1205
|
[
"Mathematics -> Applied Mathematics -> Statistics -> Probability -> Counting Methods -> Permutations",
"Mathematics -> Number Theory -> Congruences"
] | 7
|
During a break, $n$ children at school sit in a circle around their teacher to play a game. The teacher walks clockwise close to the children and hands out candies to some of them according to the following rule. He selects one child and gives him a candy, then he skips the next child and gives a candy to the next one, then he skips 2 and gives a candy to the next one, then he skips 3, and so on. Determine the values of $n$ for which eventually, perhaps after many rounds, all children will have at least one candy each.
|
Number the children from 0 to $n-1$. Then the teacher hands candy to children in positions $f(x)=1+2+\cdots+x \bmod n=\frac{x(x+1)}{2} \bmod n$. Our task is to find the range of $f: \mathbb{Z}_{n} \rightarrow \mathbb{Z}_{n}$, and to verify whether the range is $\mathbb{Z}_{n}$, that is, whether $f$ is a bijection. If $n=2^{a}m, m>1$ odd, look at $f(x)$ modulo $m$. Since $m$ is odd, $m|f(x) \Longleftrightarrow m| x(x+1)$. Then, for instance, $f(x) \equiv 0(\bmod m)$ for $x=0$ and $x=m-1$. This means that $f(x)$ is not a bijection modulo $m$, and there exists $t$ such that $f(x) \not \equiv t(\bmod m)$ for all $x$. By the Chinese Remainder Theorem, $$f(x) \equiv t \quad(\bmod n) \Longleftrightarrow \begin{cases}f(x) \equiv t & \left(\bmod 2^{a}\right) \\ f(x) \equiv t & (\bmod m)\end{cases}$$ Therefore, $f$ is not a bijection modulo $n$. If $n=2^{a}$, then $$f(x)-f(y)=\frac{1}{2}(x(x+1)-y(y+1))=\frac{1}{2}(x^{2}-y^{2}+x-y)=\frac{(x-y)(x+y+1)}{2}.$$ and $$f(x) \equiv f(y) \quad\left(\bmod 2^{a}\right) \Longleftrightarrow(x-y)(x+y+1) \equiv 0 \quad\left(\bmod 2^{a+1}\right) \tag{*}$$ If $x$ and $y$ have the same parity, $x+y+1$ is odd and $(*)$ is equivalent to $x \equiv y\left(\bmod 2^{a+1}\right)$. If $x$ and $y$ have different parity, $$(*) \Longleftrightarrow x+y+1 \equiv 0 \quad\left(\bmod 2^{a+1}\right)$$ However, $1 \leq x+y+1 \leq 2(2^{a}-1)+1=2^{a+1}-1$, so $x+y+1$ is not a multiple of $2^{a+1}$. Therefore $f$ is a bijection if $n$ is a power of 2.
|
All powers of 2
|
apmoapmo_sol
|
omni_math-1793
|
[
"Mathematics -> Number Theory -> Congruences",
"Mathematics -> Number Theory -> Prime Numbers"
] | 7
|
Determine all the pairs $ (p , n )$ of a prime number $ p$ and a positive integer $ n$ for which $ \frac{ n^p + 1 }{p^n + 1} $ is an integer.
|
To solve the problem of finding all pairs \((p, n)\) of a prime number \(p\) and a positive integer \(n\) for which \(\frac{n^p + 1}{p^n + 1}\) is an integer, we start by analyzing the expression:
\[
\frac{n^p + 1}{p^n + 1}.
\]
**Step 1: Initial observation**
We need to determine when this ratio is an integer. Clearly, \((p^n + 1)\) must divide \((n^p + 1)\).
**Step 2: Consider special cases**
- **Case 1: \(n = p\)**
Substituting \(n = p\) into the expression gives:
\[
\frac{p^p + 1}{p^p + 1} = 1,
\]
which is indeed an integer. Thus, \((p, n) = (p, p)\) is a valid solution for any prime \(p\).
- **Case 2: \(p = 2\)**
Substituting \(p = 2\) into the expression gives:
\[
\frac{n^2 + 1}{2^n + 1}.
\]
Testing small values of \(n\), we have:
- For \(n = 4\):
\[
\frac{4^2 + 1}{2^4 + 1} = \frac{16 + 1}{16 + 1} = \frac{17}{17} = 1,
\]
which is an integer. Therefore, \((p, n) = (2, 4)\) is a valid solution.
- **Step 3: Verify uniqueness of solutions**
Consider other values and general arguments. By testing small pairs or using congruences, it becomes apparent that the expression does not become an integer for \(n \neq p\) or other simple substitutions, except in potentially specific adjusted conditions or constructions that do not yield another simple, universal form like the ones found.
Thus, the complete solution set is:
\[
\boxed{(p, p), (2, 4)}.
\]
This includes all pairs \((p, p)\) for any prime number \(p\) and the specific pair \((2, 4)\).
|
$(p,n)=(p,p),(2,4)$
|
apmo
|
omni_math-3925
|
[
"Mathematics -> Discrete Mathematics -> Algorithms"
] | 7
|
Given three letters $X, Y, Z$, we can construct letter sequences arbitrarily, such as $XZ, ZZYXYY, XXYZX$, etc. For any given sequence, we can perform following operations:
$T_1$: If the right-most letter is $Y$, then we can add $YZ$ after it, for example, $T_1(XYZXXY) =
(XYZXXYYZ).$
$T_2$: If The sequence contains $YYY$, we can replace them by $Z$, for example, $T_2(XXYYZYYYX) =
(XXYYZZX).$
$T_3$: We can replace $Xp$ ($p$ is any sub-sequence) by $XpX$, for example, $T_3(XXYZ) = (XXYZX).$
$T_4$: In a sequence containing one or more $Z$, we can replace the first $Z$ by $XY$, for example,
$T_4(XXYYZZX) = (XXYYXYZX).$
$T_5$: We can replace any of $XX, YY, ZZ$ by $X$, for example, $T_5(ZZYXYY) = (XYXX)$ or $(XYXYY)$ or $(ZZYXX).$
Using above operations, can we get $XYZZ$ from $XYZ \ ?$
|
To determine whether we can transform the sequence "XYZ" into "XYZZ" using the operations \( T_1, T_2, T_3, T_4, \) and \( T_5 \), we systematically examine how these operations affect the sequence:
### Initial Sequence
The starting sequence is:
\[
\text{XYZ}
\]
### Available Operations and Their Effects
1. **Operation \( T_1 \)**: This operation applies only if the sequence ends with "Y":
- Transform \(\text{Y} \to \text{YYZ}\).
Applying \( T_1 \) to \(\text{XYZ}\), we have:
\[
\text{XYZ} \to \text{XYZYZ}
\]
2. **Operation \( T_2 \)**: This can be applied if the sequence contains "YYY":
- Replace \(\text{YYY} \to \text{Z}\).
However, \(\text{XYZ}\) and sequences obtained from it, like \(\text{XYZYZ}\), do not contain "YYY".
3. **Operation \( T_3 \)**: This can be applied to any sequence:
- Transform \( Xp \to XpX \) (append "X" to the end).
Applying \( T_3 \) to \(\text{XYZ}\), we have:
\[
\text{XYZ} \to \text{XYZX}
\]
Applying \( T_3 \) to \(\text{XYZYZ}\), we have:
\[
\text{XYZYZ} \to \text{XYZYZX}
\]
4. **Operation \( T_4 \)**: Replace the first "Z" with "XY", if any "Z" is present:
- Change \(\ldots Z \ldots \to \ldots XY \ldots \).
Applying \( T_4 \) to \(\text{XYZ}\), we have:
\[
\text{XYZ} \to \text{XYXY}
\]
5. **Operation \( T_5 \)**: Replace any "XX", "YY", or "ZZ" with "X":
- If these sequences appear, they can be replaced.
However, in sequences like \(\text{XYZ}\), "ZZ" isn't present.
### Analysis
Throughout the available transformations from "XYZ", the creation of a double "ZZ" instead of individual "Z" appears problematic:
- **ZZZ Creation**: To use \( T_5 \) to replace with "X", you need "ZZ" specifically. Given our operations, generating a sequence like "ZZ" from transformations and reductions seems out of reach without a replicator, as creating additional "Z" via operators like \( T_4 \) do not introduce double "Z":
- \( T_1 \) and \( T_3 \) extend sequences but do not produce consecutive "ZZ".
- \( T_4 \) reduces "Z" by \( XY \).
- **No operation explicitly compounds consecutive "Z"**.
Thus, using these operations, the composition of "XYZZ" from "XYZ" is unachievable, concluding:
\[
\boxed{\text{no}}
\]
|
\text{no}
|
imo_longlists
|
omni_math-4378
|
[
"Mathematics -> Applied Mathematics -> Statistics -> Probability -> Counting Methods -> Other",
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Algebra -> Intermediate Algebra -> Complex Numbers"
] | 7
|
Kelvin and 15 other frogs are in a meeting, for a total of 16 frogs. During the meeting, each pair of distinct frogs becomes friends with probability $\frac{1}{2}$. Kelvin thinks the situation after the meeting is cool if for each of the 16 frogs, the number of friends they made during the meeting is a multiple of 4. Say that the probability of the situation being cool can be expressed in the form $\frac{a}{b}$, where $a$ and $b$ are relatively prime. Find $a$.
|
Consider the multivariate polynomial $$\prod_{1 \leq i<j \leq 16}\left(1+x_{i} x_{j}\right)$$ We're going to filter this by summing over all $4^{16} 16$-tuples $\left(x_{1}, x_{2}, \ldots, x_{16}\right)$ such that $x_{j}= \pm 1, \pm i$. Most of these evaluate to 0 because $i^{2}=(-i)^{2}=-1$, and $1 \cdot-1=-1$. If you do this filtering, you get the following 4 cases: Case 1: Neither of $i$ or $-i$ appears. Then the only cases we get are when all the $x_{j}$ are 1, or they're all -1. Total is $2^{121}$. Case 2: $i$ appears, but $-i$ does not. Then all the remaining $x_{j}$ must be all 1 or all -1. This contributes a sum of $(1+i)^{15} \cdot 2^{105}+(1-i)^{15} \cdot 2^{105}=2^{113}$. $i$ can be at any position, so we get $16 \cdot 2^{113}$. Case 3: $-i$ appears, but $i$ does not. Same contribution as above. $16 \cdot 2^{113}$. Case 4: Both $i$ and $-i$ appear. Then all the rest of the $x_{j}$ must be all 1 or all -1. This contributes a sum of $2 \cdot(1+i(-i)) \cdot(1+i)^{14} \cdot(1-i)^{14} \cdot 2^{91}=2^{107}$. $i$ and $-i$ can appear in $16 \cdot 15$ places, so we get $240 \cdot 2^{107}$. So the final answer is this divided a factor for our filter. $\left(4^{16}=2^{32}\right.$.) So our final answer is $\frac{2^{89}+16 \cdot 2^{82}+240 \cdot 2^{75}}{2^{120}}=\frac{1167}{2^{41}}$. Therefore, the answer is 1167.
|
1167
|
HMMT_2
|
omni_math-1526
|
[
"Mathematics -> Geometry -> Plane Geometry -> Other"
] | 7
|
One day, there is a Street Art Show at somewhere, and there are some spectators around. We consider this place as an Euclidean plane. Let $K$ be the center of the show. And name the spectators by $A_{1}, A_{2}, \ldots, A_{n}, \ldots$ They pick their positions $P_{1}, P_{2}, \ldots, P_{n}, \ldots$ one by one. The positions need to satisfy the following three conditions simultaneously. (i) The distance between $K$ and $A_{n}$ is no less than 10 meters, that is, $K P_{n} \geq 10 \mathrm{~m}$ holds for any positive integer $n$. (ii) The distance between $A_{n}$ and any previous spectator is no less than 1 meter, that is, $P_{m} P_{n} \geq 1 \mathrm{~m}$ holds for any $n \geq 2$ and any $1 \leq m \leq n-1$. (iii) $A_{n}$ always choose the position closest to $K$ that satisfies (i) and (ii), that is, $K P_{n}$ reaches its minimum possible value. If there are more than one point that satisfy (i) and (ii) and have the minimum distance to $K, A_{n}$ may choose any one of them. For example, $A_{1}$ is not restricted by (ii), so he may choose any point on the circle $C$ which is centered at $K$ with radius 10 meters. For $A_{2}$, since there are lots of points on $C$ which are at least 1 meter apart from $P_{1}$, he may choose anyone of them. (1) Which of the following statement is true? (A) There exist positive real numbers $c_{1}, c_{2}$ such that for any positive integer $n$, no matter how $A_{1}, A_{2}, \ldots, A_{n}$ choose their positions, $c_{1} \leq K P_{n} \leq c_{2}$ always hold (unit: meter); (B) There exist positive real numbers $c_{1}, c_{2}$ such that for any positive integer $n$, no matter how $A_{1}, A_{2}, \ldots, A_{n}$ choose their positions, $c_{1} \sqrt{n} \leq K P_{n} \leq c_{2} \sqrt{n}$ always hold (unit: meter); (C) There exist positive real numbers $c_{1}, c_{2}$ such that for any positive integer $n$, no matter how $A_{1}, A_{2}, \ldots, A_{n}$ choose their positions, $c_{1} n \leq K P_{n} \leq c_{2} n$ always hold (unit: meter); (D) There exist positive real numbers $c_{1}, c_{2}$ such that for any positive integer $n$, no matter how $A_{1}, A_{2}, \ldots, A_{n}$ choose their positions, $c_{1} n^{2} \leq K P_{n} \leq c_{2} n^{2}$ always hold (unit: meter).
|
The answer is B. Suppose the length of $K P_{n}$ is $d_{n}$ meters. We consider the discs centered at $P_{1}, P_{2}, \ldots, P_{n-1}$ with radius 1 meter. Use the property of $P_{n}$ we get that these discs and the interior of $C$ cover the disc centered at $K$ with radius $d_{n}$, so $$ \pi \cdot d_{n}^{2} \leq(n-1) \cdot \pi \cdot 1^{2}+\pi \cdot 10^{2} $$ It follows that $$ d_{n} \leq \sqrt{n+99} \leq \sqrt{100 n}=10 \sqrt{n} $$ On the other hand, we consider the discs centered at $P_{1}, P_{2}, \ldots, P_{n}$ with radius $\frac{1}{2}$ meter. Since the distance between any two of $P_{1}, P_{2}, \ldots, P_{n}$ is not less than 1 meter, all these discs do not intersect. Note that every length of $K P_{1}, K P_{2}, \ldots, K P_{n}$ is not more than $d_{n}$ meter (If the length of $K P_{m}$ is more than $d_{n}$ meter, then $A_{m}$ can choose $P_{n}$, which is closer to $K$, contradiction.) So all these discs are inside the circle centered at $K$ with radius $d_{n}+\frac{1}{2}$, and $$ \pi\left(d_{n}+\frac{1}{2}\right)^{2} \geq n \cdot \pi \cdot\left(\frac{1}{2}\right)^{2} $$ So $$ d_{n} \geq \frac{\sqrt{n}}{2}-\frac{1}{2} $$ For $n=1, d_{1}=10$; For $n \geq 2$, we have that $\frac{1}{2}<\frac{2 \sqrt{n}}{5}$, so $$ d_{n}>\frac{\sqrt{n}}{2}-\frac{2 \sqrt{n}}{5}=\frac{\sqrt{n}}{10} $$ Therefore, $\frac{\sqrt{n}}{10} \leq d_{n} \leq 10 \sqrt{n}$, (B) is correct.
|
c_{1} \sqrt{n} \leq K P_{n} \leq c_{2} \sqrt{n}
|
alibaba_global_contest
|
omni_math-3383
|
[
"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
|
omni_math-200
|
[
"Mathematics -> Algebra -> Abstract Algebra -> Other"
] | 7
|
Find all functions $f:\mathbb Z\rightarrow \mathbb Z$ such that, for all integers $a,b,c$ that satisfy $a+b+c=0$, the following equality holds:
\[f(a)^2+f(b)^2+f(c)^2=2f(a)f(b)+2f(b)f(c)+2f(c)f(a).\]
(Here $\mathbb{Z}$ denotes the set of integers.)
[i]
|
To solve the functional equation, we are given that for any integers \(a\), \(b\), and \(c\) such that \(a+b+c=0\), the following must hold:
\[
f(a)^2 + f(b)^2 + f(c)^2 = 2f(a)f(b) + 2f(b)f(c) + 2f(c)f(a).
\]
Let's rewrite the equation by transferring all terms to one side:
\[
f(a)^2 + f(b)^2 + f(c)^2 - 2f(a)f(b) - 2f(b)f(c) - 2f(c)f(a) = 0.
\]
The left-hand side can be factored as:
\[
(f(a) - f(b))^2 + (f(b) - f(c))^2 + (f(c) - f(a))^2 = 0.
\]
For the sum of squares to equal zero, each individual square must be zero. Thus, we have:
\[
f(a) - f(b) = 0, \quad f(b) - f(c) = 0, \quad f(c) - f(a) = 0.
\]
This implies:
\[
f(a) = f(b) = f(c).
\]
Since this must hold for all integers \(a\), \(b\), and \(c\) such that \(a+b+c=0\), it indicates that \(f\) is a constant function. However, we also consider other potential periodic behaviors based on the symmetries inherent in integers.
### Exploring Possible Solutions
1. **Constant Function:** If \(f(t) = k\) for all \(t\), then clearly the given equality holds for any integers \(a, b,\) and \(c\) since all terms are equal.
**Solution:** \( f(t) = 0 \) for all integers \(t\).
2. **Piecewise Linear Staircase Function:**
- Consider solutions where \(f(t)\) has different values for even and odd integers.
- Let \(f(t) = 0\) for even \(t\) and \(f(t) = f(1)\) for odd \(t\).
- Check the condition:
- If \(a, b, c\) are such that \(a+b+c=0\) and have alternating parity, the original equation holds.
**Solution:** \( f(t) = 0 \) for even \(t\) and \( f(t) = f(1) \) for odd \(t\).
3. **Scalar Multiple Function:**
- Assume \(f(t) = kv(t)\) for a linear homogeneous function \(v(t)\).
- Testing simple forms: \( f(t) = 4f(1) \) for even \(t\) and \(f(t) = f(1)\) for odd \(t\).
- Verifying symmetry, this satisfies the correlation when substituted.
**Solution:** \( f(t) = 4f(1) \) for even \(t\) and \( f(t) = f(1) \) for odd \(t\).
4. **Quadratic Function:**
- A general form of solutions might be quadratic in terms of an initial term: \(f(t) = t^2 f(1)\).
- Substituting back and expanding verifies that the symmetry holds.
**Solution:** \( f(t) = t^2 f(1) \) for any \(f(1)\).
Given the symmetry and periodic characteristics within this problem structure, the possible solutions, considering \(f(1)\) is any integer, are:
\[
\boxed{f(t) = 0 \text{ for all } t.}
\]
\[
\boxed{f(t) = 0 \text{ for } t \text{ even and } f(t) = f(1) \text{ for } t \text{ odd.}}
\]
\[
\boxed{f(t) = 4f(1) \text{ for } t \text{ even and } f(t) = f(1) \text{ for } t \text{ odd.}}
\]
\[
\boxed{f(t) = t^2 f(1) \text{ for any } f(1).}
\]
|
f(t) = 0 \text{ for all } t.
\text{ OR }
f(t) = 0 \text{ for } t \text{ even and } f(t) = f(1) \text{ for } t \text{ odd}
\text{ OR }
f(t) = 4f(1) \text{ for } t \text{ even and } f(t) = f(1) \text{ for } t \text{ odd}
\text{ OR }
f(t) = t^2 f(1) \text{ for any } f(1).
|
imo
|
omni_math-4268
|
[
"Mathematics -> Number Theory -> Divisibility -> Other",
"Mathematics -> Algebra -> Algebra -> Equations and Inequalities"
] | 7.5
|
Find all integers $\,a,b,c\,$ with $\,1<a<b<c\,$ such that \[ (a-1)(b-1)(c-1) \] is a divisor of $abc-1.$
|
We are tasked with finding all integers \( a, b, c \) with \( 1 < a < b < c \) such that
\[
(a-1)(b-1)(c-1)
\]
is a divisor of
\[
abc - 1.
\]
Let's first express \( abc - 1 \) in terms of potential divisors' expressions:
1. We want \((a-1)(b-1)(c-1) \mid abc - 1\), meaning \((a-1)(b-1)(c-1)\) divides \(abc - 1\).
Since \((a-1), \ (b-1),\) and \((c-1)\) are all positive integers greater than 1, we explore integer solutions systematically by substitution and testing constraints.
### Consider the case \(a = 2\):
1. \(b\) and \(c\) must satisfy \(abc - 1 \equiv 0 \pmod{(a-1)(b-1)(c-1)}\). When \(a = 2\), the expression simplifies to:
\[
bc - 1 \equiv 0 \pmod{(1)(b-1)(c-1)}.
\]
2. Simplifying:
\[
(b-1)(c-1) \mid 2bc - 1 \implies bc \equiv 1 \pmod{(b-1)(c-1)}.
\]
Testing small integers \(b\) and \(c\) subject to \(1 < 2 < b < c\):
- For \(b = 4\) and \(c = 8\):
- \(bc = 32\) and \((b-1)(c-1) = 3 \times 7 = 21\).
Checking divisibility:
- \( 2bc - 1 = 64 - 1 = 63\),
- Since \(63 \equiv 0 \pmod{21}\), this implies \((b, c) = (4, 8)\) is valid for \(a = 2\).
Thus, \((2, 4, 8)\) is one solution.
### Consider the case \(a = 3\):
1. For \(a = 3\), we have:
\[
(b-1)(c-1) \mid 3bc - 1.
\]
Exploring possible values of \(b\) and \(c\) given \(1 < 3 < b < c\):
- For \(b = 5\) and \(c = 15\):
- \(bc = 75\) and \((b-1)(c-1) = 4 \times 14 = 56\).
Calculating:
- \(3bc - 1 = 225 - 1 = 224\),
- Since \(224 \equiv 0 \pmod{56}\), this satisfies the condition.
Thus, \((3, 5, 15)\) is another solution.
### No Larger Values:
Looking for additional combinations of \((a, b, c)\) where \(1 < a < b < c\), any further increase in \(a, b, c\) leads to values of \((a-1)(b-1)(c-1)\) that no longer satisfy the division condition when checked against new \(abc - 1\) values under these constraints, given \(a \leq 3\).
Thus, the two valid triplet solutions found are:
\[
\boxed{(2, 4, 8)} \quad \text{and} \quad \boxed{(3, 5, 15)}
\]
Both satisfy \((a-1)(b-1)(c-1) \mid abc - 1\) and adhere to \(1 < a < b < c\).
|
(2, 4, 8) \text{ and } (3, 5, 15)
|
imo
|
omni_math-3891
|
[
"Mathematics -> Number Theory -> Congruences",
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
Let $ p$ be an odd prime number. How many $ p$-element subsets $ A$ of $ \{1,2,\dots,2p\}$ are there, the sum of whose elements is divisible by $ p$?
|
Let \( p \) be an odd prime number. We are tasked with finding the number of \( p \)-element subsets \( A \) of the set \(\{1, 2, \dots, 2p\}\) such that the sum of the elements in \( A \) is divisible by \( p \).
### Step 1: Representation of Subsets
The set \(\{1, 2, \dots, 2p\}\) contains \( 2p \) elements. We want to choose \( p \) elements from these \( 2p \) elements. The number of ways to choose \( p \) elements from \( 2p \) is given by \(\binom{2p}{p}\).
### Step 2: Applying Properties of Subsets and Congruences
When we consider a subset \( A \) with \( p \) elements, the sum of its elements, denoted as \( \sum_{a \in A} a \), needs to satisfy:
\[
\sum_{a \in A} a \equiv 0 \pmod{p}.
\]
By properties of combinatorial numbers, when subsets are considered modulo a prime, symmetry and cancellation often play roles. Specifically, we will use the properties of binomial coefficients and modulo arithmetic.
### Step 3: Symmetry and Combinatorial Argument
Given the symmetry and periodicity modulo \( p \), every element's contribution modulo \( p \) can be considered as additive symmetric pairs \((x, 2p+1-x)\) whose sums modulo \( p \) cover the full range as the whole set is considered.
### Step 4: Exploiting Known Theorems
Using Lucas' theorem or known properties about binomial coefficients modulo an odd prime \( p \), it can be derived that:
\[
\binom{2p}{p} \equiv 2 \pmod{p}.
\]
This relation outlines that, up to a multiple of \( p \), there are:
\[
\binom{2p}{p} = 2p \times k + 2
\]
possible selections of sets, where the residue class modulo \( p \), i.e., remainder when divided by \( p \), is \( 2 \).
### Conclusion
Since we are dividing the total number of such subsets by \( p \) to get those subsets where the sum is specifically zero modulo \( p \), the formula for the number of such subsets is:
\[
\frac{1}{p} \left( \binom{2p}{p} - 2 \right) + 2.
\]
This simplifies to:
\[
\boxed{2 + \frac{1}{p} \left( \binom{2p}{p} - 2 \right)}.
\]
Thus, the number of \( p \)-element subsets \( A \) such that the sum of the elements is divisible by \( p \) is given by the expression inside the boxed formula.
|
\boxed{2 + \frac{1}{p} \left(\binom{2p}{p} - 2 \right)}
|
imo
|
omni_math-4230
|
[
"Mathematics -> Algebra -> Algebra -> Equations and Inequalities",
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7.5
|
Let $n \ge 4$ be an integer. Find all functions $W : \{1, \dots, n\}^2 \to \mathbb R$ such that for every partition $[n] = A \cup B \cup C$ into disjoint sets, \[ \sum_{a \in A} \sum_{b \in B} \sum_{c \in C} W(a,b) W(b,c) = |A| |B| |C|. \]
|
Let \( n \ge 4 \) be an integer. We need to find all functions \( W : \{1, \dots, n\}^2 \to \mathbb{R} \) such that for every partition \([n] = A \cup B \cup C\) into disjoint sets, the following condition holds:
\[
\sum_{a \in A} \sum_{b \in B} \sum_{c \in C} W(a,b) W(b,c) = |A| |B| |C|.
\]
To solve this, we denote the function \( W \) by \( f \) for simplicity. We start by considering specific partitions of \([n]\).
First, consider the partition \( P(\{1\}, \{2\}, \{3, 4, 5, \ldots\}) \):
\[
f(1,2)f(2,4) + f(1,2)f(2,5) + \cdots + f(1,2)f(2,n) = (n-2) - f(1,2)f(2,3).
\]
Next, consider the partition \( P(\{1\}, \{3\}, \{2, 4, 5, \ldots\}) \):
\[
f(1,3)f(3,4) + f(1,3)f(3,5) + \cdots + f(1,3)f(3,n) = (n-2) - f(1,3)f(3,2).
\]
Now, consider the partition \( P(\{1\}, \{2, 3\}, \{4, 5, \ldots\}) \):
\[
f(1,2)f(2,4) + f(1,2)f(2,5) + \cdots + f(1,2)f(2,n) + f(1,3)f(3,4) + f(1,3)f(3,5) + \cdots + f(1,3)f(3,n) = 2(n-3).
\]
This simplifies to:
\[
(n-2) - f(1,2)f(2,3) + (n-2) - f(1,3)f(3,2) = 2(n-3) \implies f(1,2)f(2,3) + f(1,3)f(3,2) = 2.
\]
Similarly, for any distinct \( a, b, c \):
\[
f(a,b)f(b,c) + f(a,c)f(c,b) = 2.
\]
Considering \( P(\{3, 4, 5, \ldots\}, \{2\}, \{1\}) \), \( P(\{2, 4, 5, \ldots\}, \{3\}, \{1\}) \), and \( P(\{4, 5, \ldots\}, \{2, 3\}, \{1\}) \), we get:
\[
f(3,2)f(2,1) + f(2,3)f(3,1) = 2,
\]
which generalizes to:
\[
f(a,b)f(b,c) + f(b,a)f(a,c) = 2.
\]
Thus, we see that:
\[
f(a,c)f(c,b) = f(b,a)f(a,c) \implies f(a,c) = 0 \quad \text{or} \quad f(c,b) = f(b,a).
\]
Suppose \( f(a,c) = 0 \) for some \( a \neq c \). Considering \( P(\{a\}, \{c\}, [n] - \{a,c\}) \), we get \( 0 = n-2 \), a contradiction. Hence, \( f(c,b) = f(b,a) \) for all distinct \( a, b, c \).
This implies that \( f(x,b) \) and \( f(b,x) \) are constants for each \( b \) and all \( x \neq b \). Consequently, \( f(x,y) \) is a constant \( k \) if \( x \neq y \). From the condition \( 2k^2 = 2 \), we find \( k = 1 \) or \( k = -1 \).
Therefore, all solutions are of the form where \( W(a,a) \) can be any value, and for all distinct \( a, b \in [n] \), \( W(a,b) \) equals a constant \( k \), where \( k = 1 \) or \( k = -1 \).
The answer is: \boxed{W(a,b) = k \text{ for all distinct } a, b \text{ and } k = 1 \text{ or } k = -1.}
|
W(a,b) = k \text{ for all distinct } a, b \text{ and } k = 1 \text{ or } k = -1.
|
usa_team_selection_test
|
omni_math-60
|
[
"Mathematics -> Geometry -> Differential Geometry -> Curvature",
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
Let $P_1,P_2,\dots,P_n$ be $n$ distinct points over a line in the plane ($n\geq2$). Consider all the circumferences with diameters $P_iP_j$ ($1\leq{i,j}\leq{n}$) and they are painted with $k$ given colors. Lets call this configuration a ($n,k$)-cloud.
For each positive integer $k$, find all the positive integers $n$ such that every possible ($n,k$)-cloud has two mutually exterior tangent circumferences of the same color.
|
Consider \( n \) distinct points \( P_1, P_2, \ldots, P_n \) arranged on a line in the plane, and we define circumferences using these points as diameters \( P_iP_j \) for \( 1 \leq i < j \leq n \). Each circumference is colored using one of \( k \) colors, forming a configuration called an \((n, k)\)-cloud.
The objective is to identify all positive integers \( n \) such that every possible \((n, k)\)-cloud has two mutually exterior tangent circumferences of the same color.
#### Step-by-step Solution
1. **Understanding Exterior Tangency:**
Two circumferences are mutually exterior tangent if they touch each other at exactly one point and do not intersect otherwise. For two circumferences with diameters \( P_iP_j \) and \( P_kP_l \) (where \( i, j, k, l \) are distinct), mutual exterior tangency occurs if one circumference is completely outside but just touching the other at exactly one point.
2. **Color Distribution:**
Given \( k \) colors, we want at least two circumferences of the same color to be mutually exterior tangent. To ensure this, symmetry and distribution principles become pivotal.
3. **Pigeonhole Principle Application:**
The total number of distinct circumferences that can be formed from \( n \) points is \( \binom{n}{2} \). We are coloring these with \( k \) colors. According to the pigeonhole principle, to guarantee at least one pair of circumferences sharing the same color, we require:
\[
\binom{n}{2} > k \implies \frac{n(n-1)}{2} > k
\]
4. **Bonding of Circles:**
To ensure exterior tangency in every \((n, k)\)-cloud, we derive conditions to minimize the arrangements that could bypass the condition of having two such circles of the same color. An engineered design would suggest that \( n \) must be large enough that non-tangency configurations do not span all color choices. It ensures overlap within any chosen \( k \).
5. **Deriving the Bound on \( n \):**
Observing the need for flexibility in color choices while assuring tangencies prompts us to use efficient coloring that avoids creating overlaps with very small \( n \).
The derived threshold \( n \ge 2^k + 1 \) follows from detailed combinatorial exploration, ensuring that no simple division among \( k \) colors for fewer points avoids mutual tangency due to their geometric and symmetric nature on the line.
Thus, the solution deduces that to guarantee the existence of two mutually exterior tangent circumferences of the same color in any \((n, k)\)-cloud, the minimum \( n \) should satisfy:
\[
\boxed{n \geq 2^k + 1}
\]
|
n \geq 2^k + 1
|
bero_American
|
omni_math-3720
|
[
"Mathematics -> Geometry -> Plane Geometry -> Polygons"
] | 7
|
For what real values of $k>0$ is it possible to dissect a $1 \times k$ rectangle into two similar, but noncongruent, polygons?
|
Given a \(1 \times k\) rectangle, we want to determine for which real values of \(k > 0\) it is possible to dissect the rectangle into two similar, but noncongruent, polygons.
First, let's understand the requirements: two polygons are similar if their corresponding angles are equal and their corresponding sides are in proportion. They are noncongruent if they are not identical in shape and size.
### Step-by-Step Analysis
1. **Understanding Rectangle Dissection**:
- A \(1 \times k\) rectangle has dimensions of length \(k\) and width \(1\).
2. **Dissecting into Two Polygons**:
- Suppose we can dissect this rectangle into two polygons, \(P_1\) and \(P_2\), where both polygons are similar to each other but not congruent.
3. **Properties of Similar Polygons**:
- If two polygons are similar, then their corresponding side lengths are in the same ratio. Thus, if polygon \(P_1\) has sides \(a_1, b_1, \ldots\), then polygon \(P_2\) has sides proportional to these, say \(ka_1, kb_1, \ldots\), where \(k\) is the factor of similarity.
- Since \(P_1\) and \(P_2\) are not congruent, the similarity ratio \(k\) cannot be equal to 1.
4. **Special Condition**:
- Assume initially that the rectangles are congruent (\(k = 1\)), implying that \(P_1\) and \(P_2\) would be identical, contradicting the requirement of being noncongruent.
- Thus, for these polygons to be noncongruent and similar, the similarity factor \(k\) cannot be 1.
5. **Conclusion**:
- The only restriction for \(k\) comes from the prohibition of \(P_1\) and \(P_2\) being congruent. Therefore, the possible values for \(k\) are all positive real numbers except 1.
Thus, the solution to the problem, where \(k\) allows a \(1 \times k\) rectangle to be dissected into two similar but noncongruent polygons, is:
\[
\boxed{k \ne 1}
\]
===
report over
===
|
$k \ne 1$
|
usamo
|
omni_math-4379
|
[
"Mathematics -> Geometry -> Plane Geometry -> Triangulations",
"Mathematics -> Geometry -> Plane Geometry -> Other"
] | 7
|
For each $P$ inside the triangle $ABC$, let $A(P), B(P)$, and $C(P)$ be the points of intersection of the lines $AP, BP$, and $CP$ with the sides opposite to $A, B$, and $C$, respectively. Determine $P$ in such a way that the area of the triangle $A(P)B(P)C(P)$ is as large as possible.
|
Let \( \triangle ABC \) be a given triangle. For any point \( P \) inside this triangle, define the intersections \( A(P), B(P), C(P) \) as follows:
- \( A(P) \) is the intersection of line \( AP \) with side \( BC \).
- \( B(P) \) is the intersection of line \( BP \) with side \( CA \).
- \( C(P) \) is the intersection of line \( CP \) with side \( AB \).
We aim to determine the position of \( P \) such that the area of \( \triangle A(P)B(P)C(P) \) is maximized.
### Analyzing the Geometry
The area of \( \triangle A(P)B(P)C(P) \) is closely tied to the location of \( P \). Specifically, this area is maximized when \( P \) is the centroid of \( \triangle ABC \). This conclusion can be drawn by considering the specific properties of the centroid:
- The centroid divides each median in a 2:1 ratio.
- It is the point within \( \triangle ABC \) where the triangle is divided into smaller triangles of equal area.
### Area Calculation
For the maximal area condition, consider \( P \) to be the centroid \( G \) of \( \triangle ABC \). The area of the triangle \( \triangle A(P)B(P)C(P) \) formed by the cevians (medians) is known from the properties of centroids:
\[
\text{Area of } \triangle A(P)B(P)C(P) = \frac{1}{4} \times \text{Area of } \triangle ABC
\]
This formula arises from the fact that the centroid divides the triangle into smaller triangles each having equal area, resulting in four smaller triangles each having one-fourth the area of \( \triangle ABC \).
### Conclusion
Thus, when \( P \) is placed at the centroid of the triangle \( \triangle ABC \), the area of triangle \( \triangle A(P)B(P)C(P) \) becomes:
\[
\boxed{\frac{S_{\triangle ABC}}{4}}
]
This completes the solution for determining \( P \) such that the area of \( \triangle A(P)B(P)C(P) \) is maximized.
|
\frac{S_{\triangle ABC}}{4}
|
imo_longlists
|
omni_math-4156
|
[
"Mathematics -> Number Theory -> Prime Numbers",
"Mathematics -> Algebra -> Algebra -> Polynomial Operations"
] | 7
|
Find all prime numbers $p$ for which there exists a unique $a \in\{1,2, \ldots, p\}$ such that $a^{3}-3 a+1$ is divisible by $p$.
|
We show that $p=3$ is the only prime that satisfies the condition. Let $f(x)=x^{3}-3 x+1$. As preparation, let's compute the roots of $f(x)$. By Cardano's formula, it can be seen that the roots are $2 \operatorname{Re} \sqrt[3]{\frac{-1}{2}+\sqrt{\left(\frac{-1}{2}\right)^{2}-\left(\frac{-3}{3}\right)^{3}}}=2 \operatorname{Re} \sqrt[3]{\cos \frac{2 \pi}{3}+i \sin \frac{2 \pi}{3}}=\left\{2 \cos \frac{2 \pi}{9}, 2 \cos \frac{4 \pi}{9}, 2 \cos \frac{8 \pi}{9}\right\}$ where all three values of the complex cubic root were taken. Notice that, by the trigonometric identity $2 \cos 2 t=(2 \cos t)^{2}-2$, the map $\varphi(x)=x^{2}-2$ cyclically permutes the three roots. We will use this map to find another root of $f$, when it is considered over $\mathbb{F}_{p}$. Suppose that $f(a)=0$ for some $a \in \mathbb{F}_{p}$ and consider $g(x)=\frac{f(x)}{x-a}=\frac{f(x)-f(a)}{x-a}=x^{2}+a x+\left(a^{2}-3\right)$. We claim that $b=a^{2}-2$ is a root of $g(x)$. Indeed, $g(b)=\left(a^{2}-2\right)^{2}+a\left(a^{2}-2\right)+\left(a^{2}-3\right)=(a+1) \cdot f(a)=0$. By Vieta's formulas, the other root of $g(x)$ is $c=-a-b=-a^{2}-a+2$. If $f$ has a single root then the three roots must coincide, so $a=a^{2}-2=-a^{2}-a+2$. Here the quadratic equation $a=a^{2}-2$, or equivalently $(a+1)(a-2)=0$, has two solutions, $a=-1$ and $a=2$. By $f(-1)=f(2)=3$, in both cases we have $0=f(a)=3$, so the only choice is $p=3$. Finally, for $p=3$ we have $f(1)=-1, f(2)=3$ and $f(3)=19$, from these values only $f(2)$ is divisible by 3, so $p=3$ satisfies the condition.
|
3
|
imc
|
omni_math-2329
|
[
"Mathematics -> Algebra -> Algebra -> Polynomial Operations",
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
Let $ n$ be a positive integer. Find the number of odd coefficients of the polynomial
\[ u_n(x) \equal{} (x^2 \plus{} x \plus{} 1)^n.
\]
|
Given the polynomial \( u_n(x) = (x^2 + x + 1)^n \), we are tasked with finding the number of odd coefficients in its expansion.
Firstly, let's expand \( (x^2 + x + 1)^n \) and observe that the coefficients of the resulting polynomial can be represented in terms of binomial coefficients. By the Binomial Theorem, we have
\[
u_n(x) = \sum_{k=0}^{n} \binom{n}{k} (x^2)^k (x+1)^{n-k}.
\]
Now, each term \((x^2)^k (x+1)^{n-k}\) can be further expanded using the Binomial Theorem on \((x+1)^{n-k}\):
\[
(x+1)^{n-k} = \sum_{m=0}^{n-k} \binom{n-k}{m} x^m.
\]
Thus, the full expansion of \(u_n(x)\) is
\[
u_n(x) = \sum_{k=0}^{n} \sum_{m=0}^{n-k} \binom{n}{k} \binom{n-k}{m} x^{2k+m}.
\]
The coefficient of \(x^j\) in this double sum is thus
\[
\sum_{2k+m=j} \binom{n}{k} \binom{n-k}{m}.
\]
We must count the number of these coefficients that are odd, which is equivalent to evaluating them modulo 2.
To determine which coefficients are odd, consider the expression \((x^2 + x + 1)^n\) modulo 2. Notice,
\[
(x^2 + x + 1) \equiv x^2 + x + 1 \pmod{2}.
\]
Observing that \(x^2 + x + 1\) has no real roots when considered over the field GF(2) (a finite field with 2 elements), it forms a cyclic group of order 3. Whenever the binary equivalent form repeats, it affects the parity of the coefficients.
A concept from finite fields can be used, specifically properties about roots of unity over GF(2) and dynamic properties of polynomial expansion to inform us about odd coefficients.
For odd coefficients, observe pattern periodicity over roots: let \( \omega \) be a cube root of unity. Then, the polynomial:
\[
u_n(1) + u_n(\omega) + u_n(\omega^2)
\]
helps in determining those patterns. By calculating the expression in finite fields:
\[
(x^2+x+1)^n = \prod_{i=0}^{n-1} (x-\omega^i),
\]
an odd coefficient arises exactly from positions defined by binary Hamming weights being reduced under a modulus of similar roots.
Analyzing these values, the specific oddness arises from coefficients characterized by this combinatorial context.
Finally, the number of odd coefficients is given as the product:
\[
\boxed{\prod f(a_i)}
\]
where each \(f(a_i)\) describes characteristically necessary evaluations for each polynomial regime. We can explore each coefficient reference tailored to field qualities, but fundamentally respects established partitions and their roots, complete by proving explicit criteria over set generator values, reducing convolution simply. In essence, solve number relations matching field cycles across polynomial symmetry.
|
\prod f(a_i)
|
imo_shortlist
|
omni_math-4311
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
Let $n$ be a positive integer. Anna and Beatrice play a game with a deck of $n$ cards labelled with the numbers $1, 2,...,n$. Initially, the deck is shuffled. The players take turns, starting with Anna. At each turn, if $k$ denotes the number written on the topmost card, then the player first looks at all the cards and then rearranges the $k$ topmost cards. If, after rearranging, the topmost card shows the number k again, then the player has lost and the game ends. Otherwise, the turn of the other player begins. Determine, depending on the initial shuffle, if either player has a winning strategy, and if so, who does.
|
Consider a deck with \( n \) cards labeled \( 1, 2, \ldots, n \) arranged in some initial order. We need to determine under what circumstances Anna, who starts the game, has a winning strategy. The strategy depends on the number \( k \) on the topmost card at each player's turn.
### Game Description:
1. At each turn, the player observes the topmost card, which has the number \( k \). The allowed move is to rearrange the top \( k \) cards in any order.
2. If the top card remains as \( k \) after rearrangement the same player loses, and the game ends.
### Strategy Analysis:
Let's consider what gives Anna a winning strategy:
- Anna examines the number \( k \) on the topmost card. She has the freedom to rearrange the \( k \) topmost cards.
- If \( k \) is the smallest among these \( k \) cards, any rearrangement will necessarily keep \( k \) as the topmost card. This results in Anna losing immediately because the topmost card after her rearrangement is still \( k \).
However, if \( k \) is **not** the smallest card among the top \( k \) cards, Anna can always rearrange these cards such that a card smaller than \( k \) becomes the topmost card, hence avoiding losing:
- This move changes the topmost card to a number smaller than \( k \), passing control to Beatrice without losing right away. With each player's optimal play, Beatrice faces the same situation: if the number on the topmost card during her turn is the smallest among the cards she can rearrange, she will lose.
Thus, Anna has a winning strategy if and only if the number \( k \) on the topmost card is **not** the smallest of the \( k \) topmost cards:
\[
\boxed{\text{Anna has a winning strategy if and only if } k \text{ is not the smallest of the } k \text{ topmost cards.}}
\]
```
|
\text{Anna has a winning strategy if and only if } k \text{ is not the smallest of the } k \text{ topmost cards.}
|
middle_european_mathematical_olympiad
|
omni_math-3735
|
[
"Mathematics -> Algebra -> Algebra -> Polynomial Operations"
] | 7
|
For each integer $m$, consider the polynomial \[P_m(x)=x^4-(2m+4)x^2+(m-2)^2.\] For what values of $m$ is $P_m(x)$ the product of two non-constant polynomials with integer coefficients?
|
By the quadratic formula, if $P_m(x)=0$, then $x^2=m\pm 2\sqrt{2m}+2$, and hence the four roots of $P_m$ are given by $S = \{\pm\sqrt{m}\pm\sqrt{2}\}$. If $P_m$ factors into two nonconstant polynomials over the integers, then some subset of $S$ consisting of one or two elements form the roots of a polynomial with integer coefficients.
First suppose this subset has a single element, say $\sqrt{m} \pm \sqrt{2}$; this element must be a rational number. Then $(\sqrt{m} \pm \sqrt{2})^2 = 2 + m \pm 2 \sqrt{2m}$ is an integer, so $m$ is twice a perfect square, say $m = 2n^2$. But then $\sqrt{m} \pm \sqrt{2} = (n\pm 1)\sqrt{2}$ is only rational if $n=\pm 1$, i.e., if $m = 2$.
Next, suppose that the subset contains two elements; then we can take it to be one of $\{\sqrt{m} \pm \sqrt{2}\}$, $\{\sqrt{2} \pm \sqrt{m}\}$ or $\{\pm (\sqrt{m} + \sqrt{2})\}$. In all cases, the sum and the product of the elements of the subset must be a rational number. In the first case, this means $2\sqrt{m} \in \QQ$, so $m$ is a perfect square. In the second case, we have $2 \sqrt{2} \in \QQ$, contradiction. In the third case, we have $(\sqrt{m} + \sqrt{2})^2 \in \QQ$, or $m + 2 + 2\sqrt{2m} \in \QQ$, which means that $m$ is twice a perfect square.
We conclude that $P_m(x)$ factors into two nonconstant polynomials over the integers if and only if $m$ is either a square or twice a square.
Note: a more sophisticated interpretation of this argument can be given using Galois theory. Namely, if $m$ is neither a square nor twice a square, then the number fields $\QQ(\sqrt{m})$ and $\QQ(\sqrt{2})$ are distinct quadratic fields, so their compositum is a number field of degree 4, whose Galois group acts transitively on $\{\pm \sqrt{m} \pm \sqrt{2}\}$. Thus $P_m$ is irreducible.
|
m is either a square or twice a square.
|
putnam
|
omni_math-3550
|
[
"Mathematics -> Algebra -> Intermediate Algebra -> Inequalities",
"Mathematics -> Geometry -> Plane Geometry -> Other"
] | 7
|
Let $k$ be an arbitrary natural number.
Let $\{m_1,m_2,\ldots{},m_k\}$ be a permutation of $\{1,2,\ldots{},k\}$ such that $a_{m_1} < a_{m_2} < \cdots{} < a_{m_k}$.
Note that we can never have equality since $|a_{m_i} - a_{m_{i+1}}| \ge \frac{1}{m_i+m_{i+1}}$.
Let $\overline{a_ia_j} = |a_i-a_j|$.
By looking at the $a_i$ as a set of intervals on $[0,c]$, it makes sense that $\overline{a_{m_1}a_{m_k}} = \sum \limits_{i=1}^{k-1} \overline{a_{m_i}a_{m_{i+1}}}$.
$\overline{a_{m_i}a_{m_k}} \ge \sum\limits_{i=1}^{k-1} \frac{1}{m_i+m_{i+1}}$.
By the Arithmetic Mean Harmonic Mean Inequality,
$\frac{(a_1+a_2) + (a_2+a_3) + \ldots{} + (m_{k-1}+m_k)}{k-1} \ge \frac{k-1}{\frac{1}{m_1+m_2} + \ldots{} + \frac{1}{m_{k-1}+m_k}}$.
$(m_1+2m_2+\ldots{}+2m_{k-1}+2m_k)\left(\frac{1}{m_1+m_2} + \ldots{} + \frac{1}{m_{k-1}+m_k}\right) \ge (k-1)^2$.
$(\overline{a_{m_1}a_{m_k}})(m_1+2m_2+\ldots{}+2m_{k-1}+m_k) \ge (k-1)^2$.
The right term of the left-hand side is less than $2(m_1+m_2+\ldots{}+m_k)$:
$2\overline{a_{m_1}a_{m_k}}(m_1+m_2+\ldots{}+m_k) > (k-1)^2$
Since $\{m_1,m_2,\ldots{},m_k\}$ is a permutation of $\{1,2,\ldots{},k\}$,
$2\overline{a_{m_1}a_{m_k}} \cdot \frac{k(k+1)}{2} > (k-1)^2$.
$\overline{a_{m_1}a_{m_k}} > \frac{(k-1)^2}{k(k+1)} = \frac{k-1}{k} \cdot \frac{k-1}{k+1} > \left(\frac{k-1}{k+1}\right)^2 = \left(1-\frac{2}{k+1}\right)^2$.
If $\overline{a_{m_1}a_{m_k}} < 1$ for all $k \in \mathbb N$, we can easily find a $k$ such that $\left(1-\frac{2}{k+1}\right)^2 > \overline{a_{m_1}a_{m_k}}$, causing a contradiction.
So $\overline{a_{m_1}a_{m_k}} \ge 1$ for some integers $m_1$, $m_k$.
$|a_{m_1}-a_{m_k}| \ge 1$.
Since both terms are positive, it is clear that at least one of them is greater than or equal to $1$.
So $c \ge 1$, as desired.
|
Consider the permutation \(\{m_1, m_2, \ldots, m_k\}\) of \(\{1, 2, \ldots, k\}\) such that \(a_{m_1} < a_{m_2} < \cdots < a_{m_k}\), and note that:
\[
|a_{m_i} - a_{m_{i+1}}| \ge \frac{1}{m_i + m_{i+1}}
\]
Based on this permutation, the total distance \(\overline{a_{m_1}a_{m_k}} = |a_{m_1} - a_{m_k}|\) can be interpreted as a sum of the smaller intervals:
\[
\overline{a_{m_1}a_{m_k}} = \sum_{i=1}^{k-1} \overline{a_{m_i}a_{m_{i+1}}}
\]
Applying the inequality given, we have:
\[
\overline{a_{m_1}a_{m_k}} \ge \sum_{i=1}^{k-1} \frac{1}{m_i + m_{i+1}}
\]
The Arithmetic Mean-Harmonic Mean Inequality (AM-HM Inequality) gives us:
\[
\frac{(m_1 + m_2) + (m_2 + m_3) + \cdots + (m_{k-1} + m_k)}{k-1} \ge \frac{k-1}{\frac{1}{m_1 + m_2} + \cdots + \frac{1}{m_{k-1} + m_k}}
\]
Simplifying, this implies:
\[
(m_1 + 2m_2 + \cdots + 2m_{k-1} + m_k) \left( \frac{1}{m_1 + m_2} + \cdots + \frac{1}{m_{k-1} + m_k} \right) \ge (k-1)^2
\]
Since \(m_1 + 2m_2 + \ldots + 2m_{k-1} + m_k\) is less than or equal to \(2(m_1 + m_2 + \ldots + m_k)\), we find:
\[
2\overline{a_{m_1}a_{m_k}} (m_1 + m_2 + \cdots + m_k) \ge (k-1)^2
\]
And knowing that \(\{m_1, m_2, \ldots, m_k\}\) is a permutation of \(\{1, 2, \ldots, k\}\), we find:
\[
2\overline{a_{m_1}a_{m_k}} \cdot \frac{k(k+1)}{2} \ge (k-1)^2
\]
This implies:
\[
\overline{a_{m_1}a_{m_k}} \ge \frac{(k-1)^2}{k(k+1)} = \frac{k-1}{k} \cdot \frac{k-1}{k+1}
\]
Further simplifying, we get:
\[
\overline{a_{m_1}a_{m_k}} > \left(\frac{k-1}{k+1}\right)^2 = \left(1 - \frac{2}{k+1}\right)^2
\]
Finally, if \(\overline{a_{m_1}a_{m_k}} < 1\) for all \(k \in \mathbb{N}\), a contradiction arises, suggesting that:
\[
\overline{a_{m_1}a_{m_k}} \ge 1
\]
Therefore, since \(|a_{m_1} - a_{m_k}|\) is at least 1, it follows that \(c \ge 1\).
Hence, the conclusion is:
\[
\boxed{c \ge 1}
\]
|
imo_shortlist
|
omni_math-4160
|
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Discrete Mathematics -> Logic"
] | 7
|
$A$ and $B$ play a game, given an integer $N$, $A$ writes down $1$ first, then every player sees the last number written and if it is $n$ then in his turn he writes $n+1$ or $2n$, but his number cannot be bigger than $N$. The player who writes $N$ wins. For which values of $N$ does $B$ win?
[i]
|
To determine for which values of \( N \) player \( B \) wins, we need to analyze the structure of the game and identify a strategy that ensures victory for player \( B \).
### Game Analysis
Given the rules of the game:
- Player \( A \) starts by writing the number \( 1 \).
- Each player alternates turns writing either \( n+1 \) or \( 2n \), where \( n \) is the last number written.
- The number written cannot exceed \( N \).
- The player who writes \( N \) wins.
To solve the problem, we must deduce for which values of \( N \), player \( B \) is guaranteed a win regardless of how optimally player \( A \ plays.
### Winning and Losing Positions
In combinatorial game theory, we identify "winning" and "losing" positions:
- **Winning Position:** A position where the player whose turn it is can force a win with optimal play.
- **Losing Position:** A position where the player whose turn it is will lose with optimal play from the opponent.
The strategy involves determining the losing positions. Player \( B \) will win if and only if \( A \) begins his turn in a losing position.
### Characterizing Losing Positions
1. **Base Case:**
- \( N = 1 \) is a losing position because player \( A \) starts and immediately wins.
2. **Recursive Analysis:**
- If writing \( n \) is a losing position, then both \( n+1 \) and \( 2n \) must be winning positions.
- Conversely, \( n \) is a winning position if either \( n+1 \) or \( 2n \) is a losing position.
By iteratively applying these conditions, we can deduce:
- A position \( n \) is losing if \( n \) is the sum of distinct odd powers of 2. These are numbers whose binary representation consists of zeros and a single one in an odd position.
### Examples and Verification
For example, consider \( N = 3 \).
- \( N = 1 \) is losing (as defined above).
- Starting from 1, player \( A \) can choose 2.
- Then, player \( B \) can choose 3 and win. Hence, 3 was a winning position for \( B \).
Values of \( N \) being sums of distinct odd powers of 2 include \( N = 1, 3, 5, 9, 17, \ldots \). In binary, these are 1, 11, 101, 1001, 10001, corresponding to \( 2^0, 2^1 + 2^0, 2^2 + 2^0, 2^3 + 2^0, 2^4 + 2^0, \ldots \).
### Final Conclusion
Player \( B \) wins for values of \( N \) that are sums of distinct odd powers of 2. Therefore, the answer is:
\[
\boxed{N = \text{the sum of distinct odd powers of } 2}
\]
This strategy ensures that player \( B \) will always find themselves in a winning position when \( N \) fits this pattern. The correct identification of such \( N \) solidifies player \( B \)'s chances of winning.
|
N = \text{the sum of distinct odd powers of }2
|
imo_shortlist
|
omni_math-4400
|
[
"Mathematics -> Discrete Mathematics -> Graph Theory"
] | 7.5
|
The following operation is allowed on a finite graph: Choose an arbitrary cycle of length 4 (if there is any), choose an arbitrary edge in that cycle, and delete it from the graph. For a fixed integer ${n\ge 4}$, find the least number of edges of a graph that can be obtained by repeated applications of this operation from the complete graph on $n$ vertices (where each pair of vertices are joined by an edge).
[i]
|
Consider the complete graph \( K_n \) on \( n \) vertices, where \( n \geq 4 \). The graph initially contains \(\binom{n}{2} = \frac{n(n-1)}{2}\) edges. We want to find the least number of edges that can be left in the graph by repeatedly applying the following operation: choose an arbitrary cycle of length 4, then choose an arbitrary edge in that cycle, and delete it.
### Strategy:
The goal is to minimize the number of edges in the final graph, avoiding any cycles of length 4. A graph without cycles of length 4 is known as a *triangle-free* graph for \( n \geq 4 \).
### Analysis:
1. **Initial Observation**: Removing edges from cycles of length 4 reduces the number of edges, but the goal is to minimize the number of edges left, ensuring no 4-cycles remain.
2. **Example of a Target Graph**: A simple graph structure that has no 4-cycles is a *star* graph \( S_n \), which is obtained by selecting one vertex to be the center and connecting it to all other \( n-1 \) vertices. The star graph is acyclic and clearly contains exactly \( n-1 \) edges.
3. **Verification**:
- The operation directly targets 4-cycles, which a star graph cannot have.
- After removing edges from all cycles length 4 in the complete graph, a possible structure similar to a star graph or any other tree structure emerges with \( n-1 \) edges and no 4-cycles.
4. **Lower Bound Justification**:
- Consider Turan's theorem for extremal graph theory:
- For a graph without cycles of length 4, known as \( C_4 \)-free, the number of edges \( e \) satisfies:
\[
e \leq \frac{n^2}{4}.
\]
- If \( e \leq n \) is achievable while ensuring no 4-cycles, it's optimal.
5. **Constructing the Final Graph**:
- On achieving the goal where edges left correspond to a linear or star configuration, having \( n \) edges is plausible as each vertex connects to a distinct vertex linearly.
Therefore, with persistent deletion of edges in cycles of length 4, we aim to settle at a graph where only a minimal set of edges corresponding to a line or star remains, typically \( n \).
### Conclusion:
Therefore, the least number of edges that remains, ensuring no further 4-cycles can be formed, is:
\[
\boxed{n}.
\]
Thus, the reference answer is confirmed.
|
n
|
imo_shortlist
|
omni_math-4111
|
[
"Mathematics -> Algebra -> Abstract Algebra -> Other"
] | 7
|
Find all functions $f: \mathbb{R}^{+} \rightarrow \mathbb{R}^{+}$ such that $$(z+1) f(x+y)=f(x f(z)+y)+f(y f(z)+x)$$ for all positive real numbers $x, y, z$.
|
The identity function $f(x)=x$ clearly satisfies the functional equation. Now, let $f$ be a function satisfying the functional equation. Plugging $x=y=1$ into (3) we get $2 f(f(z)+1)=(z+1)(f(2))$ for all $z \in \mathbb{R}^{+}$. Hence, $f$ is not bounded above. Lemma. Let $a, b, c$ be positive real numbers. If $c$ is greater than $1, a / b$ and $b / a$, then the system of linear equations $$c u+v=a \quad u+c v=b$$ has a positive real solution $u, v$. Proof. The solution is $$u=\frac{c a-b}{c^{2}-1} \quad v=\frac{c b-a}{c^{2}-1}$$ The numbers $u$ and $v$ are positive if the conditions on $c$ above are satisfied. We will now prove that $$f(a)+f(b)=f(c)+f(d) \text { for all } a, b, c, d \in \mathbb{R}^{+} \text {with } a+b=c+d$$ Consider $a, b, c, d \in \mathbb{R}^{+}$such that $a+b=c+d$. Since $f$ is not bounded above, we can choose a positive number $e$ such that $f(e)$ is greater than $1, a / b, b / a, c / d$ and $d / c$. Using the above lemma, we can find $u, v, w, t \in \mathbb{R}^{+}$satisfying $$\begin{aligned} & f(e) u+v=a, \quad u+f(e) v=b \\ & f(e) w+t=c, \quad w+f(e) t=d . \end{aligned}$$ Note that $u+v=w+t$ since $(u+v)(f(e)+1)=a+b$ and $(w+t)(f(e)+1)=c+d$. Plugging $x=u, y=v$ and $z=e$ into (3) yields $f(a)+f(b)=(e+1) f(u+v)$. Similarly, we have $f(c)+f(d)=(e+1) f(w+t)$. The claim follows immediately. We then have $$y f(x)=f(x f(y)) \quad \text { for all } x, y \in \mathbb{R}^{+}$$ since by $(3)$ and $(4)$, $$(y+1) f(x)=f\left(\frac{x}{2} f(y)+\frac{x}{2}\right)+f\left(\frac{x}{2} f(y)+\frac{x}{2}\right)=f(x f(y))+f(x)$$ Now, let $a=f(1 / f(1))$. Plugging $x=1$ and $y=1 / f(1)$ into (5) yields $f(a)=1$. Hence $a=a f(a)$ and $f(a f(a))=f(a)=1$. Since $a f(a)=f(a f(a))$ by (5), we have $f(1)=a=1$. It follows from (5) that $$f(f(y))=y \quad \text { for all } y \in \mathbb{R}^{+}$$ Using (4) we have for all $x, y \in \mathbb{R}^{+}$that $$\begin{aligned} & f(x+y)+f(1)=f(x)+f(y+1), \quad \text { and } \\ & f(y+1)+f(1)=f(y)+f(2) \end{aligned}$$ Therefore $$f(x+y)=f(x)+f(y)+b \text { for all } x, y \in \mathbb{R}^{+}$$ where $b=f(2)-2 f(1)=f(2)-2$. Using (5), (7) and (6), we get $$4+2 b=2 f(2)=f(2 f(2))=f(f(2)+f(2))=f(f(2))+f(f(2))+b=4+b$$ This shows that $b=0$ and thus $$f(x+y)=f(x)+f(y) \text { for all } x, y \in \mathbb{R}^{+}$$ In particular, $f$ is strictly increasing. We conclude as follows. Take any positive real number $x$. If $f(x)>x$, then $f(f(x))>$ $f(x)>x=f(f(x))$, a contradiction. Similarly, it is not possible that $f(x)<x$. This shows that $f(x)=x$ for all positive real numbers $x$.
|
f(x)=x \text{ for all positive real numbers } x
|
apmoapmo_sol
|
omni_math-1582
|
[
"Mathematics -> Algebra -> Intermediate Algebra -> Other",
"Mathematics -> Applied Mathematics -> Statistics -> Probability -> Counting Methods -> Other"
] | 7
|
Let $a_1,a_2,a_3,\cdots$ be a non-decreasing sequence of positive integers. For $m\ge1$ , define $b_m=\min\{n: a_n \ge m\}$ , that is, $b_m$ is the minimum value of $n$ such that $a_n\ge m$ . If $a_{19}=85$ , determine the maximum value of $a_1+a_2+\cdots+a_{19}+b_1+b_2+\cdots+b_{85}$ .
|
We create an array of dots like so: the array shall go out infinitely to the right and downwards, and at the top of the $i$ th column we fill the first $a_i$ cells with one dot each. Then the $19$ th row shall have 85 dots. Now consider the first 19 columns of this array, and consider the first 85 rows. In row $j$ , we see that the number of blank cells is equal to $b_j-1$ . Therefore the number of filled cells in the first 19 columns of row $j$ is equal to $20-b_j$ .
We now count the number of cells in the first 19 columns of our array, but we do it in two different ways. First, we can sum the number of dots in each column: this is simply $a_1+\cdots+a_{19}$ . Alternatively, we can sum the number of dots in each row: this is $(20-b_1)+\cdots +(20-b_{85})$ . Since we have counted the same number in two different ways, these two sums must be equal. Therefore \[a_1+\cdots +a_{19}+b_1+\cdots +b_{85}=20\cdot 85=\boxed{1700}.\] Note that this shows that the value of the desired sum is constant.
|
\boxed{1700}
|
usamo
|
omni_math-179
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
Let $n$ be a positive integer. Determine the size of the largest subset of $\{ -n, -n+1, \dots, n-1, n\}$ which does not contain three elements $a$, $b$, $c$ (not necessarily distinct) satisfying $a+b+c=0$.
|
Consider the set \( S = \{-n, -n+1, \ldots, n-1, n\} \). We want to find the size of the largest subset of \( S \) such that no three elements \( a, b, c \) within the subset satisfy \( a + b + c = 0 \).
To solve this problem, it is useful to evaluate the properties of numbers that sum to zero. For each positive integer \( k \), the triplet \((-k, 0, k)\) automatically sums to zero. In our problem, we need to avoid selecting any three numbers summing to zero, which implies avoiding such typical triplets or any rearrangement that sums to zero.
### Strategy
1. **Splitting the Set**: Consider the set of numbers in \( S \). One approach is to select either all negative numbers up to zero or positive numbers including zero such that their absolute values don't lead to zero-summing triplets. A typical choice revolves around balancing positive and negative numbers while avoiding zero where possible.
2. **Constructing the Subset**: Consider selecting negatives and zero, or negatives paired with positives in a way that zero-summing is avoided:
- **Case 1**: Select numbers are covering as much as possible while preventing zero-sum. E.g., all negatives and zero when avoiding balancing.
- **Case 2**: Pair each negative \( -k \) with a positive number \( k \) beyond what \( -k \) can sum to zero with (avoiding \((k, 0, -k)\)).
Consider splitting \( S \) into parts:
- **Negative Set**: \(\{ -n, -n+1, \ldots, -1, 0 \}\)
- **Positive Set**: \(\{ 1, 2, \ldots, n\}\)
Attempt to construct subsets avoiding zero-summing triplet selection.
3. **Maximum Balanced Set**: To be most inclusive without a zero triplet:
- Include negative numbers to zero without their reverse \( k \) (in a size balanced between odd and even size adjustments).
- Uses a selection reliant on sequence patterns in integers that if \( k \) is an extent, selection is symmetrical or extended to ensure balance without triplet sums.
### Counting the Optimal Case
By carefully selecting and avoiding elements:
- The ideal number of elements in the subset is twice the greatest positive round number limited by ceiling division:
\[
\text{size} = 2 \left\lceil \frac{n}{2} \right\rceil
\]
Thus, the size of the largest subset that does not include any three elements summing to zero is:
\[
\boxed{2 \left\lceil \frac{n}{2} \right\rceil}
\]
|
2 \left\lceil \frac{n}{2} \right\rceil
|
usamo
|
omni_math-3566
|
[
"Mathematics -> Number Theory -> Factorization",
"Mathematics -> Number Theory -> Prime Numbers"
] | 7
|
For any positive integer $d$, prove there are infinitely many positive integers $n$ such that $d(n!)-1$ is a composite number.
|
For any positive integer \( d \), we aim to prove that there are infinitely many positive integers \( n \) such that \( d(n!) - 1 \) is a composite number.
### Case 1: \( d = 1 \)
Assume for the sake of contradiction that for all sufficiently large \( n \in \mathbb{N} \), \( n! - 1 \) is prime. Define \( p_n = n! - 1 \) to be this prime for all \( n \geq N \) for some \( N \in \mathbb{N} \). Notice that \( p_n > n \). Consider \( n = p_n - k_n \) for some \( k_n \in \mathbb{N} \).
In \( \mathbb{F}_{p_n} \), by Wilson's Theorem, we have:
\[
1 = (p_n - k_n)! = \frac{(p_n - 1)!}{(p_n - 1)(p_n - 2) \cdots (p_n - k_n + 1)} = \frac{-1}{(-1)^{k_n - 1} \cdot (k_n - 1)!} = (-1)^{k_n} \cdot \frac{1}{(k_n - 1)!}
\]
Thus,
\[
(k_n - 1)! \equiv (-1)^{k_n} \pmod{p_n}
\]
If we pick \( n \equiv 1 \pmod{2} \), then \( k_n \equiv p_n - n \equiv 0 \pmod{2} \) as \( p_n \) is odd, and therefore \( p_n \mid (k_n - 1)! - 1 \).
Since \( k_n > C \) for all sufficiently large \( n \in \mathbb{N} \) for any \( C \in \mathbb{N} \), and considering the pigeonhole principle, we conclude that \( p_n = (k_n - 1)! - 1 \). Hence, \( p_n = n! - 1 \) implies \( p_n - n - 1 = k_n - 1 = n \), which means \( p_n = 2n + 1 \). This implies \( 2n + 1 = n! - 1 \) for infinitely many \( n \in \mathbb{N} \), which is impossible for large \( n \). Therefore, \( n! - 1 \) cannot be prime for all sufficiently large \( n \).
### General Case: \( d \geq 2 \)
Assume for the sake of contradiction that there exists some \( N \in \mathbb{N} \) such that for all \( n \in \mathbb{N}, n \geq N \), \( d \cdot n! - 1 \) is prime. Define the function \( f: \mathbb{N} \setminus [d] \to \mathbb{N} \) as:
\[
f(m) = (d-1)! \cdot (d+1)(d+2) \cdots (m-1) \cdot m + (-1)^m
\]
Let \( p_m \) be the smallest prime divisor of \( f(m) \). For \( m \geq 2d \), \( p_m > m \) because \( \gcd(m!, f(m) - 1) \mid d \) but \( d \mid f(m) \) implies \( \gcd(f(m) - 1, d) = 1 \). Thus, \( p_m \mid f(m) + (-1)^m \), meaning:
\[
0 \equiv d(f(m) + (-1)^m) \equiv m! + d(-1)^m \pmod{p_m}
\]
Using Wilson's Theorem in \( \mathbb{F}_{p_m} \):
\[
(p_m - m - 1)! = \frac{(p_m - 1)!}{(p_m - 1)(p_m - 2) \cdots (p_m - m)} = \frac{-1}{(-1)^m \cdot m!} = \frac{-1}{(-1)^m (-(-1)^m) \cdot d} = \frac{1}{d}
\]
Thus,
\[
p_m \mid d(p_m - m - 1)! - 1
\]
For sufficiently large \( m \), \( p_m = d \cdot (p_m - m - 1)! - 1 \). By the pigeonhole principle, \( k_m = p_m - m \) can appear finitely many times. Therefore, \( k_m \geq N + 1 \) implies \( d \cdot (p_m - m - 1)! - 1 \) is prime, leading to a contradiction.
Hence, for any positive integer \( d \), there are infinitely many positive integers \( n \) such that \( d(n!) - 1 \) is a composite number.
The answer is: \boxed{\text{There are infinitely many positive integers } n \text{ such that } d(n!) - 1 \text{ is a composite number.}}
|
\text{There are infinitely many positive integers } n \text{ such that } d(n!) - 1 \text{ is a composite number.}
|
china_team_selection_test
|
omni_math-132
|
[
"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
|
omni_math-255
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Algebra -> Intermediate Algebra -> Inequalities"
] | 7
|
Do there exist two bounded sequences $a_{1}, a_{2}, \ldots$ and $b_{1}, b_{2}, \ldots$ such that for each positive integers $n$ and $m > n$ at least one of the two inequalities $|a_{m} - a_{n}| > \frac{1}{\sqrt{n}}, |b_{m} - b_{n}| > \frac{1}{\sqrt{n}}$ holds?
|
Suppose such sequences $(a_{n})$ and $(b_{n})$ exist. For each pair $(x, y)$ of real numbers we consider the corresponding point $(x, y)$ in the coordinate plane. Let $P_{n}$ for each $n$ denote the point $(a_{n}, b_{n})$. The condition in the problem requires that the square $\{(x, y): |x - a_{n}| \leq \frac{1}{\sqrt{n}}, |y - b_{n}| \leq \frac{1}{\sqrt{n}}\}$ does not contain $P_{m}$ for $m \neq n$. For each point $A_{n}$ we construct its private square $\{(x, y): |x - a_{n}| \leq \frac{1}{2\sqrt{n}}, |y - b_{n}| \leq \frac{1}{2\sqrt{n}}\}$. The condition implies that private squares of points $A_{n}$ and $A_{m}$ are disjoint when $m \neq n$. Let $|a_{n}| < C, |b_{n}| < C$ for all $n$. Then all private squares of points $A_{n}$ lie in the square $\{(x, y): |x| \leq C + \frac{1}{2}, |y| \leq C + \frac{1}{2}\}$ with area $(2C + 1)^{2}$. However private squares do not intersect, and the private square of $P_{n}$ has area $\frac{1}{n}$. The series $1 + \frac{1}{2} + \frac{1}{3} + \cdots$ diverges; in particular, it contains some finite number of terms with sum greater than $(2C + 1)^{2}$, which is impossible if the respective private square lie inside a square with area $(2C + 1)^{2}$ and do not intersect. This contradiction shows that the desired sequences $(a_{n})$ and $(b_{n})$ do not exist.
|
No, such sequences do not exist.
|
izho
|
omni_math-3367
|
[
"Mathematics -> Number Theory -> Prime Numbers"
] | 7
|
Find all positive integers $a,n\ge1$ such that for all primes $p$ dividing $a^n-1$, there exists a positive integer $m<n$ such that $p\mid a^m-1$.
|
We are tasked with finding all positive integers \(a, n \ge 1\) such that for all primes \(p\) dividing \(a^n - 1\), there exists a positive integer \(m < n\) such that \(p \mid a^m - 1\).
By Zsigmondy's theorem, for any \(a > 1\) and \(n > 1\), there exists a primitive prime divisor of \(a^n - 1\) except for the cases \((a, n) = (2, 6)\) and \((a, n) = (2^k - 1, 2)\).
1. **Case \((a, n) = (2, 6)\):**
- For \(a = 2\) and \(n = 6\), we have \(2^6 - 1 = 63\).
- The prime divisors of 63 are 3 and 7.
- We need to check if these primes divide \(2^m - 1\) for some \(m < 6\).
- The values of \(2^m - 1\) for \(m < 6\) are: \(1, 3, 7, 15, 31\).
- Both 3 and 7 appear in this list, so this case holds.
2. **Case \((a, n) = (2^k - 1, 2)\):**
- For \(a = 2^k - 1\) and \(n = 2\), we have \((2^k - 1)^2 - 1 = (2^k - 1)(2^k + 1)\).
- Any prime divisor of \((2^k - 1)^2 - 1\) must divide either \(2^k - 1\) or \(2^k + 1\).
- The only possible \(m\) is \(m = 1\), which gives us \(2^k - 1 - 1 = 2^k - 2 = 2(2^{k-1} - 1)\).
- Thus, all prime divisors of \((2^k - 1)^2 - 1\) divide \(2(2^{k-1} - 1)\), satisfying the condition.
3. **Case \(a = 1\):**
- For \(a = 1\), \(1^n - 1 = 0\) for any \(n\), which trivially satisfies the condition.
Combining these results, the complete solution set is:
\[
(a, n) = (2, 6), (2^k - 1, 2), (1, n) \text{ for any } n \ge 1.
\]
The answer is: \boxed{(2, 6), (2^k - 1, 2), (1, n) \text{ for any } n \ge 1}.
|
(2, 6), (2^k - 1, 2), (1, n) \text{ for any } n \ge 1
|
usa_team_selection_test
|
omni_math-27
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
A social club has $2k+1$ members, each of whom is fluent in the same $k$ languages. Any pair of members always talk to each other in only one language. Suppose that there were no three members such that they use only one language among them. Let $A$ be the number of three-member subsets such that the three distinct pairs among them use different languages. Find the maximum possible value of $A$.
|
Let \( A \) be the number of three-member subsets such that the three distinct pairs among them use different languages. We aim to find the maximum possible value of \( A \).
Given that the social club has \( 2k+1 \) members, each fluent in \( k \) languages, and that no three members use only one language among them, we can classify the triplets into three categories: all same color, two colors, and all different colors. Since the "all same color" category is precluded, we let \( B \) be the number of two-color triples. Thus, we have:
\[
A + B = \binom{2k+1}{3}.
\]
To find the maximum \( A \), we need to show that \( B \geq k(2k+1) \). For each of the \( 2k+1 \) vertices, let \( c_i \) be the number of edges of color \( i \). The number of \( B \) triplets centered at a vertex is:
\[
\sum_{i=1}^{k} \binom{c_i}{2} \geq k \cdot \binom{\frac{\sum c_i}{k}}{2} = k \cdot \binom{2}{2} = k.
\]
Since there are \( 2k+1 \) vertices, we have:
\[
B \geq (2k+1) \cdot k.
\]
Thus, the maximum possible value of \( A \) is:
\[
A \leq \binom{2k+1}{3} - k(2k+1).
\]
For the construction, we can number all \( 2k+1 \) points and color the edges such that vertices \( P_m \) and \( P_{m+i} \) (mod \( 2k+1 \)) are colored with color \( i \). This satisfies the condition of no monochromatic triples and ensures \( B = (2k+1) \cdot k \).
Therefore, the maximum possible value of \( A \) is:
\[
A = \binom{2k+1}{3} - k(2k+1).
\]
The answer is: \boxed{\binom{2k+1}{3} - k(2k+1)}.
|
\binom{2k+1}{3} - k(2k+1)
|
usa_team_selection_test
|
omni_math-74
|
[
"Mathematics -> Geometry -> Plane Geometry -> Other",
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Discrete Mathematics -> Graph Theory"
] | 7
|
For a pair $ A \equal{} (x_1, y_1)$ and $ B \equal{} (x_2, y_2)$ of points on the coordinate plane, let $ d(A,B) \equal{} |x_1 \minus{} x_2| \plus{} |y_1 \minus{} y_2|$. We call a pair $ (A,B)$ of (unordered) points [i]harmonic[/i] if $ 1 < d(A,B) \leq 2$. Determine the maximum number of harmonic pairs among 100 points in the plane.
|
Given a set of 100 points in the plane, we want to determine the maximum number of harmonic pairs, where a pair \((A, B)\) of points is considered harmonic if \(1 < d(A, B) \leq 2\) and \(d(A, B) = |x_1 - x_2| + |y_1 - y_2|\).
To solve this problem, we can transform the distance function to make it easier to handle. By rotating the plane by 45 degrees, we change the coordinates of a point \(P = (x, y)\) to \(P' = (x - y, x + y)\). Under this transformation, the Manhattan distance \(d(P, Q)\) becomes \(d'(P', Q') = \max \{ |P'_x - Q'_x|, |P'_y - Q'_y| \}\).
We claim that the maximum number of harmonic pairs is \(\frac{3 \times 100^2}{4 \times 2} = 3750\). To achieve this bound, we can place 25 points each in small neighborhoods around the four points \((\pm \frac{1.0201082102011209}{2}, \pm \frac{1.0201082102011209}{2})\).
To prove that this is the maximum number, we construct a graph \(G\) with 100 vertices, where two vertices are connected if the corresponding points are harmonic. We need to show that \(G\) has no \(K_5\) (complete graph on 5 vertices).
**Claim:** \(G\) has no \(K_5\).
**Proof:** Consider the following two facts:
1. If a coloring of the edges of \(K_5\) with two colors does not produce a monochromatic triangle, then it must have a monochromatic cycle of length 5.
2. It is impossible to find three real numbers \(A, B, C\) such that all points \((A, 0), (B, 0), (C, 0)\) are mutually harmonic.
For each edge \(PQ\) in \(G\), color the edge red if \(\max \{ |P_x - Q_x|, |P_y - Q_y| \} = |P_x - Q_x|\), or blue otherwise. Suppose, for contradiction, that there is a \(K_5\) in \(G\) with points \(A, B, C, D, E\). By fact 2, it has no monochromatic triangle, so by fact 1, it has a monochromatic cycle of length 5. Without loss of generality, assume the cycle is red, and let it be \(A \rightarrow B \rightarrow \cdots \rightarrow E\).
If \(\max(A_y, B_y, C_y, D_y, E_y) - \min(A_y, B_y, C_y, D_y, E_y) > 2\), we have a contradiction because the extreme points would not be harmonic. Therefore, \(\max(A_y, B_y, C_y, D_y, E_y) - \min(A_y, B_y, C_y, D_y, E_y) \leq 2\). Assume \(\min(A_y, B_y, C_y, D_y, E_y) = A_y = 0\), so \(\max(A_y, B_y, C_y, D_y, E_y) \leq 2\). Thus, \(A_y, B_y, C_y, D_y, E_y \in [0, 2]\).
Color the vertices with ordinate in \([0, 1]\) black and those in \((1, 2]\) white. Traversing \(A \rightarrow B \rightarrow \cdots \rightarrow E\) changes the color of the interval each time, implying the odd cycle is bipartite, which is a contradiction.
By Turan's theorem, the strictest bound possible for the number of edges in \(G\) without a \(K_5\) is \(\frac{3 \times 100^2}{4 \times 2} = 3750\).
The answer is \(\boxed{3750}\).
|
3750
|
usa_team_selection_test
|
omni_math-8
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7.5
|
Let $n > 0$ be an integer. We are given a balance and $n$ weights of weight $2^0, 2^1, \cdots, 2^{n-1}$. We are to place each of the $n$ weights on the balance, one after another, in such a way that the right pan is never heavier than the left pan. At each step we choose one of the weights that has not yet been placed on the balance, and place it on either the left pan or the right pan, until all of the weights have been placed.
Determine the number of ways in which this can be done.
[i]
|
Consider an integer \( n > 0 \) and a balance with \( n \) weights of weights \( 2^0, 2^1, \ldots, 2^{n-1} \). Our task is to place each of these weights on the balance, one by one, so that the right pan is never heavier than the left pan. We aim to determine the number of ways to achieve this.
### Understanding the Problem
The setup involves choosing at each step one of the \( n \) weights that has not yet been placed on the balance and deciding whether it should be placed on the left pan or the right pan. This continues until all weights are placed.
### Key Constraints
1. At every step during the placement of the weights, the total weight in the right pan must not exceed the total weight in the left pan.
2. Once a weight is placed, it cannot be moved again.
### Solution Approach
To solve this problem, we consider each weight \( 2^k \) as a decision point: each weight can either be placed on the left or right pan, constrained by the requirement that the right pan never becomes heavier.
#### Combinatorial Enumeration using Catalan Paths
This is combinatorially equivalent to finding the number of ways to arrange the sequence of weights, where each step of adding a weight to the left is analogous to taking an upward step (\( U \)), and adding a weight to the right is analogous to a downward step (\( D \)). For the configuration to satisfy the condition (i.e., right pan never heavier than the left), it is essentially a "path" problem where paths never fall below the "starting level".
The number of distinct configurations achievable with these constraints is closely related to Catalan numbers, calculated in terms of "factorial double" or semifactorials, which specifically articulate the number of valid parenthesis combinations for a sequence of terms.
#### Calculation
The correct formula for the number of valid sequences like described above where the sequence never "falls below ground" is given by the formula:
\[
(2n-1)!!
\]
where \((2n-1)!!\) denotes the product of all odd integers up to \( 2n-1 \).
Thus, the number of ways the weights can be placed on the balance so that the right pan is never heavier than the left pan is:
\[
\boxed{(2n-1)!!}
\]
This result reflects the combinatorial counting of valid balanced arrangements, revealing the complexity and richness of the constraining arrangement task akin to classic path and matching problems in combinatorics.
|
(2n-1)!!
|
imo
|
omni_math-3947
|
[
"Mathematics -> Geometry -> Plane Geometry -> Triangulations",
"Mathematics -> Geometry -> Plane Geometry -> Angles"
] | 7
|
In the triangle $\triangle ABC$, let $G$ be the centroid, and let $I$ be the center of the inscribed circle. Let $\alpha$ and $\beta$ be the angles at the vertices $A$ and $B$, respectively. Suppose that the segment $IG$ is parallel to $AB$ and that $\beta = 2 \tan^{-1} (1/3)$. Find $\alpha$.
|
Let $M$ and $D$ denote the midpoint of $AB$ and the foot of the altitude from $C$ to $AB$, respectively, and let $r$ be the inradius of $\bigtriangleup ABC$. Since $C,G,M$ are collinear with $CM = 3GM$, the distance from $C$ to line $AB$ is $3$ times the distance from $G$ to $AB$, and the latter is $r$ since $IG \parallel AB$; hence the altitude $CD$ has length $3r$. By the double angle formula for tangent, $\frac{CD}{DB} = \tan\beta = \frac{3}{4}$, and so $DB = 4r$. Let $E$ be the point where the incircle meets $AB$; then $EB = r/\tan(\frac{\beta}{2}) = 3r$. It follows that $ED = r$, whence the incircle is tangent to the altitude $CD$. This implies that $D=A$, $ABC$ is a right triangle, and $\alpha = \frac{\pi}{2}$.
|
\frac{\pi}{2}
|
putnam
|
omni_math-3219
|
[
"Mathematics -> Geometry -> Solid Geometry -> 3D Shapes"
] | 7
|
Let $A,B,C,D$ denote four points in space such that at most one of the distances $AB,AC,AD,BC,BD,CD$ is greater than $1$ . Determine the maximum value of the sum of the six distances.
|
Suppose that $AB$ is the length that is more than $1$ . Let spheres with radius $1$ around $A$ and $B$ be $S_A$ and $S_B$ . $C$ and $D$ must be in the intersection of these spheres, and they must be on the circle created by the intersection to maximize the distance. We have $AC + BC + AD + BD = 4$ .
In fact, $CD$ must be a diameter of the circle. This maximizes the five lengths $AC$ , $BC$ , $AD$ , $BD$ , and $CD$ . Thus, quadrilateral $ACBD$ is a rhombus.
Suppose that $\angle CAD = 2\theta$ . Then, $AB + CD = 2\sin{\theta} + 2\cos{\theta}$ . To maximize this, we must maximize $\sin{\theta} + \cos{\theta}$ on the range $0^{\circ}$ to $90^{\circ}$ . However, note that we really only have to solve this problem on the range $0^{\circ}$ to $45^{\circ}$ , since $\theta > 45$ is just a symmetrical function.
For $\theta < 45$ , $\sin{\theta} \leq \cos{\theta}$ . We know that the derivative of $\sin{\theta}$ is $\cos{\theta}$ , and the derivative of $\cos{\theta}$ is $-\sin{\theta}$ . Thus, the derivative of $\sin{\theta} + \cos{\theta}$ is $\cos{\theta} - \sin{\theta}$ , which is nonnegative between $0^{\circ}$ and $45^{\circ}$ . Thus, we can conclude that this is an increasing function on this range.
It must be true that $2\sin{\theta} \leq 1$ , so $\theta \leq 30^{\circ}$ . But, because $\sin{\theta} + \cos{\theta}$ is increasing, it is maximized at $\theta = 30^{\circ}$ . Thus, $AB = \sqrt{3}$ , $CD = 1$ , and our sum is $5 + \sqrt{3}$ .
~mathboy100
|
\[ 5 + \sqrt{3} \]
|
usamo
|
omni_math-166
|
[
"Mathematics -> Algebra -> Intermediate Algebra -> Sequences -> Other"
] | 7
|
Given a pair $(a_0, b_0)$ of real numbers, we define two sequences $a_0, a_1, a_2,...$ and $b_0, b_1, b_2, ...$ of real numbers by $a_{n+1}= a_n + b_n$ and $b_{n+1}=a_nb_n$ for all $n = 0, 1, 2,...$. Find all pairs $(a_0, b_0)$ of real numbers such that $a_{2022}= a_0$ and $b_{2022}= b_0$.
|
Given a pair \((a_0, b_0)\) of real numbers, we define two sequences \(a_0, a_1, a_2, \ldots\) and \(b_0, b_1, b_2, \ldots\) of real numbers by the recurrence relations:
\[
a_{n+1} = a_n + b_n
\]
\[
b_{n+1} = a_n b_n
\]
for all \(n = 0, 1, 2, \ldots\).
We are tasked with finding all pairs \((a_0, b_0)\) such that \(a_{2022} = a_0\) and \(b_{2022} = b_0\).
Let's analyze the dynamics of the sequences:
### Step 1: Investigate Special Cases
1. **Case \(b_0 = 0\):**
- For \(b_0 = 0\), the recurrence does not depend on the value of \(a_0\):
\[
a_{n+1} = a_n + b_n = a_n + 0 = a_n
\]
\[
b_{n+1} = a_n b_n = a_n \cdot 0 = 0
\]
- Thus, both sequences are constant, \(a_n = a_0\) and \(b_n = 0\), for all \(n \geq 1\).
- Specifically, \(a_{2022} = a_0\) and \(b_{2022} = 0 = b_0\).
### Step 2: Existence of Other Solutions
2. **Case \(b_0 \neq 0\):**
- Assume \(b_0 \neq 0\), causing non-trivial changes:
- The sequence \(b_n\) follows: \(b_1 = a_0 b_0\), \(b_2 = a_1 b_1 = (a_0 + b_0)(a_0 b_0)\), which generally leads to a more complex pattern.
- As \(b_{n+1} = a_n b_n\), without further specifics, these sequences become complex, typically returning to the initial condition is non-trivial and requires \(b_0 = 0\).
Based on reasoned evaluation, for \((a_{2022}, b_{2022}) = (a_0, b_0)\), we conclude:
### Conclusion
- The only solution that allows \(a_{2022} = a_0\) and \(b_{2022} = b_0\) is when the sequence doesn't change from its initial conditions. This is satisfied only if \(b_0 = 0\).
- Thus, for any real number \(a \), the pairs that satisfy the condition are \((a, 0)\).
Hence, the solution is:
\[
\boxed{(a, 0) \text{ for any real number } a.}
\]
This concludes that \((a, 0)\) is the only valid pair satisfying the equation for the given recursive sequence across the mentioned iteration.
|
(a, 0) \text{ for any real number } a.
|
middle_european_mathematical_olympiad
|
omni_math-3729
|
[
"Mathematics -> Applied Mathematics -> Statistics -> Probability -> Counting Methods -> Combinations",
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
A certain state issues license plates consisting of six digits (from 0 through 9). The state requires that any two plates differ in at least two places. (Thus the plates $\boxed{027592}$ and $\boxed{020592}$ cannot both be used.) Determine, with proof, the maximum number of distinct license plates that the state can use.
|
Consider license plates of $n$ digits, for some fixed $n$ , issued with the same criteria.
We first note that by the pigeonhole principle, we may have at most $10^{n-1}$ distinct plates. Indeed, if we have more, then there must be two plates which agree on the first $n-1$ digits; these plates thus differ only on one digit, the last one.
We now show that it is possible to issue $10^{n-1}$ distinct license plates which satisfy the problem's criteria. Indeed, we issue plates with all $10^{n-1}$ possible combinations for the first $n-1$ digit, and for each plate, we let the last digit be the sum of the preceding digits taken mod 10. This way, if two plates agree on the first $n-1$ digits, they agree on the last digit and are thus the same plate, and if two plates differ in only one of the first $n-1$ digits, they must differ as well in the last digit.
It then follows that $10^{n-1}$ is the greatest number of license plates the state can issue. For $n=6$ , as in the problem, this number is $10^5$ . $\blacksquare$
|
\[ 10^5 \]
|
usamo
|
omni_math-311
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Geometry -> Plane Geometry -> Polygons"
] | 7
|
We say that a finite set $\mathcal{S}$ of points in the plane is [i]balanced[/i] if, for any two different points $A$ and $B$ in $\mathcal{S}$, there is a point $C$ in $\mathcal{S}$ such that $AC=BC$. We say that $\mathcal{S}$ is [i]centre-free[/i] if for any three different points $A$, $B$ and $C$ in $\mathcal{S}$, there is no points $P$ in $\mathcal{S}$ such that $PA=PB=PC$.
(a) Show that for all integers $n\ge 3$, there exists a balanced set consisting of $n$ points.
(b) Determine all integers $n\ge 3$ for which there exists a balanced centre-free set consisting of $n$ points.
|
Consider a finite set \(\mathcal{S}\) of points in the plane. The problem involves two specific definitions: a set is **balanced** if, for any two different points \(A\) and \(B\) in \(\mathcal{S}\), there is a point \(C\) in \(\mathcal{S}\) such that \(AC = BC\). Additionally, the set is **centre-free** if for any three different points \(A\), \(B\), and \(C\) in \(\mathcal{S}\), there is no point \(P\) in \(\mathcal{S}\) such that \(PA = PB = PC\).
### Part (a)
To show that for all integers \(n \geq 3\), there exists a balanced set consisting of \(n\) points, consider placing the points equally spaced on a circle. This configuration is symmetrical, and for any pair of points \(A\) and \(B\), the perpendicular bisector of the segment \(AB\) will contain a point \(C\) on the circle such that \(AC = BC\). This holds for any pair of points on the circle.
1. Place \(n\) points on a circle such that each point is evenly spaced.
2. For any two points \(A\) and \(B\), there exists a point \(C\) on the circle because of the symmetrical nature of the circle (specifically, the point diametrically opposite to the midpoint of arc \(AB\)).
Thus, this configuration ensures that the set is balanced for any number of points \(n \geq 3\).
### Part (b)
To determine all integers \(n \geq 3\) for which there exists a balanced centre-free set consisting of \(n\) points, analyze the geometric properties of balanced configurations.
1. Consider the property of being centre-free: if any three points \(A\), \(B\), and \(C\) are considered, no fourth point \(P\) can exist such that \(PA = PB = PC\).
2. In a balanced configuration using an even number \(n\) of points, symmetry can lead to points \(P\) equidistant from any three others, which violates the centre-free condition.
By the above consideration, a balanced centre-free set cannot exist if \(n\) is even, because having evenly spaced points on a circle for an even \(n\) results in symmetry that allows for a central point equidistant from multiple others.
Thus, a balanced centre-free set consisting of \(n\) points is only possible for:
\[
\boxed{\text{All odd integers } n \geq 3}
\]
|
\text{All odd integers } n \geq 3.
|
imo
|
omni_math-4161
|
[
"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
|
omni_math-297
|
[
"Mathematics -> Number Theory -> Congruences"
] | 7
|
Find all pairs of positive integers $m,n\geq3$ for which there exist infinitely many positive integers $a$ such that \[ \frac{a^m+a-1}{a^n+a^2-1} \] is itself an integer.
[i]Laurentiu Panaitopol, Romania[/i]
|
We are tasked with finding all pairs of positive integers \( m, n \geq 3 \) such that there exist infinitely many positive integers \( a \) making the expression
\[
\frac{a^m + a - 1}{a^n + a^2 - 1}
\]
an integer. To solve this problem, we aim to explore potential values of \( m \) and \( n \) and identify conditions that would make the expression an integer for infinitely many values of \( a \).
### Analysis
1. **Expression as a Polynomial Division**: Consider the expression given:
\[
\frac{a^m + a - 1}{a^n + a^2 - 1}
\]
2. **Degree Comparison**: Notice that the numerator \( a^m + a - 1 \) and the denominator \( a^n + a^2 - 1 \) are polynomials in \( a \). For the ratio to be an integer for large values of \( a \), the degree of the numerator should be at least the degree of the denominator. Therefore, we initially require:
\[
m \geq n
\]
3. **Specific integers \( m \) and \( n \)**:
- We seek pairs \((m, n)\) such that the difference \( m - n \) compensates for the linear offset in the numerator, allowing division without remainder.
4. **Case Analysis**:
- Suppose \( m = n+1 \). The degrees barely align, meaning significant constraints must exist on the linear coefficients or possible reductions.
- Substitute \( m = n+2 \) into our testing. Check for \( (m, n) = (5, 3) \).
5. **Checking Specific Case**:
- Consider the pair \( (m, n) = (5, 3) \):
\[
\frac{a^5 + a - 1}{a^3 + a^2 - 1}
\]
- Verify when this becomes an integer for infinitely many \( a \):
- Perform polynomial long division or factoring to examine whether this expression simplifies for large \( a \).
6. **Verification**:
- Confirm through substitution or theoretical check using algebraic identities or modular arithmetic that certain values hold the integrity needed.
- Suppose \( m = 5 \) and \( n = 3 \), then the expression approaches a scenario where the numerator and denominator balance out naturally due to polynomial degrees and composition.
### Conclusion
Through a structured polynomial analysis and checking cases, it becomes evident that the pair \( (m, n) = (5, 3) \) is a suitable solution allowing the fraction to reduce to an integer for infinitely many integers \( a \).
Thus, the solution is:
\[
\boxed{(5, 3)}
\]
This outcome indicates that no other pair of integers \( m, n \geq 3 \) fits unless they similarly satisfy the structural requirements of polynomial division for infinitely many values of \( a \).
|
(5, 3)
|
imo
|
omni_math-3823
|
[
"Mathematics -> Algebra -> Algebra -> Equations and Inequalities",
"Mathematics -> Algebra -> Algebra -> Polynomial Operations"
] | 7
|
Find all of the positive real numbers like $ x,y,z,$ such that :
1.) $ x \plus{} y \plus{} z \equal{} a \plus{} b \plus{} c$
2.) $ 4xyz \equal{} a^2x \plus{} b^2y \plus{} c^2z \plus{} abc$
Proposed to Gazeta Matematica in the 80s by VASILE C?RTOAJE and then by Titu Andreescu to IMO 1995.
|
We are given the following system of equations for positive real numbers \( x, y, z \):
1. \( x + y + z = a + b + c \)
2. \( 4xyz = a^2x + b^2y + c^2z + abc \)
We want to find all positive solutions \((x, y, z)\).
### Step 1: Substituting and Manipulating
To solve these equations, we first analyze the second equation:
\[
4xyz = a^2x + b^2y + c^2z + abc
\]
Rearrange the equation as:
\[
4xyz - a^2x - b^2y - c^2z = abc
\]
### Step 2: Symmetry and Easy Cases
Notice that the equations are symmetric in \( x, y, z \) when considered in conjunction with their corresponding coefficients \( a, b, c \).
We first try to find a symmetric solution relying on the symmetry, specifically using the linear condition:
\[
x = \frac{b+c}{2}, \quad y = \frac{a+c}{2}, \quad z = \frac{a+b}{2}
\]
### Step 3: Verify the Proposed Solution
Verify this proposed solution by plugging into the original equations:
#### Checking the First Equation
\[
x + y + z = \frac{b+c}{2} + \frac{a+c}{2} + \frac{a+b}{2} = a + b + c
\]
This satisfies the first condition.
#### Checking the Second Equation
Calculate the left side:
\[
4xyz = 4 \cdot \frac{b+c}{2} \cdot \frac{a+c}{2} \cdot \frac{a+b}{2}
\]
Calculate the right side:
\[
a^2x + b^2y + c^2z + abc = a^2 \cdot \frac{b+c}{2} + b^2 \cdot \frac{a+c}{2} + c^2 \cdot \frac{a+b}{2} + abc
\]
This simplifies through algebraic manipulation; consider the uniformity and symmetry of the solution and matching terms:
Assuming the equality holds by symmetry and assuming simple algebra without loss of generality as the terms are balanced due to the choice of \(x, y, z\), a more detailed expansion and simplification process would verify the solution indeed satisfies:
\[
4 \cdot \frac{b+c}{2} \cdot \frac{a+c}{2} \cdot \frac{a+b}{2} = a^2 \cdot \frac{b+c}{2} + b^2 \cdot \frac{a+c}{2} + c^2 \cdot \frac{a+b}{2} + abc
\]
Thus, the solution satisfies both original given equations.
### Final Solution
The positive real numbers \((x, y, z)\) that satisfy the system of equations are:
\[
\boxed{\left(\frac{b+c}{2}, \frac{a+c}{2}, \frac{a+b}{2}\right)}
\]
This completes the solving process, and no further solutions are possible within the symmetric setup given by the problem conditions.
|
(x,y,z)=\left(\frac{b+c}{2},\frac{a+c}{2},\frac{a+b}{2}\right)
|
imo_shortlist
|
omni_math-3901
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
Let $n \geq 1$ be an integer. What is the maximum number of disjoint pairs of elements of the set $\{ 1,2,\ldots , n \}$ such that the sums of the different pairs are different integers not exceeding $n$?
|
Let \( n \geq 1 \) be an integer. We want to find the maximum number of disjoint pairs from the set \( \{ 1, 2, \ldots, n \} \) such that the sums of these different pairs are different integers not exceeding \( n \).
To solve this problem, consider the set \( S = \{ 1, 2, \ldots, n \} \). We will form pairs \((a, b)\) where \( 1 \leq a < b \leq n \), and the sum of each pair \( a + b \leq n \). Let's denote by \( k \) the number of such disjoint pairs.
**Analysis:**
1. **Sum Constraints:**
For a pair \( (a, b) \) to be valid, we require:
\[
a + b \leq n.
\]
2. **Disjoint Pairs:**
Each number in the set can be used at most once across all pairs. Therefore, \( 2k \) numbers are used to form \( k \) disjoint pairs.
3. **Maximum Sum of Pairs:**
The largest sum that can be created with any two distinct numbers from the set is \( n - 1 \), i.e., when we consider the smallest and largest number available in \( S \).
4. **Expression for Maximum Number of Pairs Given Constraints:**
To achieve different sums all below or equal to \( n \), we want to pair numbers such that:
\[
a + b = s, \quad \text{for each } s \text{ as } 3, 4, \ldots, n.
\]
Since \( a \leq b \) always holds, and pairs must be disjoint and sums \( \leq n \), the maximum feasible number of pairs is achieved by utilizing the formula, derived from counting all integer sums not exceeding \( n \):
\[
\left\lfloor \frac{2n-1}{5} \right\rfloor.
\]
Thus, using this examination as supported by the conditions, the maximum number of disjoint pairs sums with distinct sums not exceeding \( n \) is given by:
\[
\boxed{\left\lfloor \frac{2n-1}{5} \right\rfloor}.
\]
|
\left \lfloor \frac{2n-1}{5} \right \rfloor
|
imo_shortlist
|
omni_math-4196
|
[
"Mathematics -> Number Theory -> Integer Solutions -> Other"
] | 7
|
Solve in $ \mathbb{Z}^2 $ the equation: $ x^2\left( 1+x^2 \right) =-1+21^y. $
|
Consider the equation in integers \( \mathbb{Z}^2 \):
\[
x^2 (1 + x^2) = -1 + 21^y.
\]
First, rewrite the equation as:
\[
x^2 + x^4 = -1 + 21^y.
\]
Thus, we have:
\[
x^4 + x^2 + 1 = 21^y.
\]
We're tasked with finding integer solutions \((x, y)\).
### Step-by-step Analysis:
1. **Case \( x = 0 \):**
Substituting \( x = 0 \) gives:
\[
0^4 + 0^2 + 1 = 1.
\]
Thus:
\[
21^y = 1.
\]
This implies:
\[
y = 0.
\]
Therefore, one solution is:
\[
(x, y) = (0, 0).
\]
2. **Case \( x \neq 0 \):**
Simplify and rearrange the equation:
\[
x^2(x^2 + 1) = -1 + 21^y.
\]
This suggests testing small values of \( x \).
3. **Trial for \( x = 1 \):**
Substituting \( x = 1 \) gives:
\[
1^2(1 + 1) + 1 = 3.
\]
\[
21^y = 2.
\]
No integer solution for \( y \).
4. **Trial for \( x = 2 \):**
Substituting \( x = 2 \) gives:
\[
2^2(4 + 1) + 1 = 17.
\]
\[
21^y = 17.
\]
No integer solution for \( y \).
5. **Trial for \( x = \pm 2 \):**
Substituting \( x = 2 \) gives:
\[
2^4 + 2^2 + 1 = 21.
\]
Thus:
\[
21^y = 21.
\]
This implies:
\[
y = 1.
\]
Therefore, two solutions are:
\[
(x, y) = (2, 1) \text{ and } (-2, 1).
\]
To conclude, the integer solutions are:
\[
\boxed{(0, 0), (2, 1), (-2, 1)}.
\]
These steps demonstrate how \((x, y)\) values satisfy the equation \(x^2(x^2 + 1) = -1 + 21^y\) in \( \mathbb{Z}^2 \).
|
(0, 0), (2, 1), (-2, 1)
|
danube_mathematical_competition
|
omni_math-4286
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Algebra -> Prealgebra -> Integers"
] | 7
|
Let $n$ be a positive integer. There are $\tfrac{n(n+1)}{2}$ marks, each with a black side and a white side, arranged into an equilateral triangle, with the biggest row containing $n$ marks. Initially, each mark has the black side up. An operation is to choose a line parallel to the sides of the triangle, and flipping all the marks on that line. A configuration is called admissible if it can be obtained from the initial configuration by performing a finite number of operations. For each admissible configuration $C$ , let $f(C)$ denote the smallest number of operations required to obtain $C$ from the initial configuration. Find the maximum value of $f(C)$ , where $C$ varies over all admissible configurations.
|
This problem needs a solution. If you have a solution for it, please help us out by adding it .
The problems on this page are copyrighted by the Mathematical Association of America 's American Mathematics Competitions .
|
The problem does not have a provided solution, so the final answer cannot be extracted.
|
usamo
|
omni_math-194
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Discrete Mathematics -> Graph Theory"
] | 7
|
A ten-level 2-tree is drawn in the plane: a vertex $A_{1}$ is marked, it is connected by segments with two vertices $B_{1}$ and $B_{2}$, each of $B_{1}$ and $B_{2}$ is connected by segments with two of the four vertices $C_{1}, C_{2}, C_{3}, C_{4}$ (each $C_{i}$ is connected with one $B_{j}$ exactly); and so on, up to 512 vertices $J_{1}, \ldots, J_{512}$. Each of the vertices $J_{1}, \ldots, J_{512}$ is coloured blue or golden. Consider all permutations $f$ of the vertices of this tree, such that (i) if $X$ and $Y$ are connected with a segment, then so are $f(X)$ and $f(Y)$, and (ii) if $X$ is coloured, then $f(X)$ has the same colour. Find the maximum $M$ such that there are at least $M$ permutations with these properties, regardless of the colouring.
|
The answer is $2^{2^{7}}$. First we need a suitable terminology. Similarly to 10-level 2-tree we can define a $k$-level 2-tree for $k \geq 1$. For convenience we suppose that all the segments between vertices are directed from a letter to the next one. The number of the letter marking a vertex we call the level of this vertex; thus $A_{1}$ is the only vertex of level $1, B_{1}$ and $B_{2}$ belong to level 2 and so on). We will also call descendants of a vertex $X$ all vertices which can be reached from $X$ by directed segments. Let $T_{1}$ and $T_{2}$ be two $k$-level 2-trees with coloured leaves. We call a bijection $f: T_{1} \rightarrow T_{2}$ isomorphism when two conditions are satisfied: (i) if two vertices $X$ and $Y$ are connected by an edge in $T_{1}$, then $f(X)$ and $f(Y)$ are connected by an edge in $T_{2}$, and (ii) if $X$ has some colour in $T_{1}$, then $f(X)$ has the same colour in $T_{2}$. When $T_{1}=T_{2}$, we call $f$ automorphism of the tree. By $\chi(k)$ we denote the minimal number of automorphisms a $k$-level 2-tree with coloured leaves can have (the minimum is over all colourings). Our problem is to find $\chi(10)$. We start with almost obvious Lemma 1. Isomorphism of trees preserves the level of a vertex. Proof. Isomorphism $f$ cannot diminish the degree of a vertex. Indeed, neighbours of each vertex $X$ become neighbours of $f(X)$, therefore the degree of $f(X)$ is not less than the degree of $X$. By pigeonhole principle it also means that the degree can not increase. It follows that the last level vertices go to the last level vertices. Therefore vertices of the previous level go to the same level, since they remain neighbours of the last-level vertices, and so on. Now we are ready to solve the problem. Proposition 1. For each $k \geq 2$ we have $\chi(k) \geq(\chi(k-1))^{2}$. Proof. In a $k$-level tree the descendants of $B_{1}$ (including $B_{1}$ ) form a $k$-1-level tree $T_{1}$. This graph has at least $\chi(k-1)$ different automorphisms. The same is true for tree $T_{2}$ formed by the descendants of $B_{2}$. Let $g$ and $h$ be automorphisms of $T_{1}$ and $T_{2}$ respectively. Now we can define mapping $f$ of the whole tree applying $g$ to descendants of $B_{1}, h$ to descendants of $B_{2}$ and $A$ to itself. Obviously $f$ is an automorphism: for $X=A$ the condition holds since $B_{1}$ and $B_{2}$ were mapped to themselves (by Lemma 1 ), and for $X$ in $T_{1}$ or $T_{2}$ because $g$ and $h$ are automorphisms. Thus for each pair $(g, h)$ there is an automorphism $f$, different pairs produce different $f$, and the number of pairs is at least $(\chi(k-1))^{2}$. Corollary. For $k \geq 3$ we have $\chi(k) \geq 2^{2^{k-3}}$. Proof. This inequality is proved by induction, with Proposition 1 as induction step. It remains to check it for $k=3$. If in a 3 -level 2 -tree at least one of the vertices $B_{1}, B_{2}$ has two descendants of the same colour, there is an automorphism exchanging these two vertices and preserving the rest. If each of $B_{1}, B_{2}$ has one blue and one golden descendant, there is an automorphism exchanging $B_{1}$ and $B_{2}$ and preserving colours of their descendant. In both cases the number of automorphisms (including the identical one) is at least 2. We already know that every 3-level 2-tree with (four) coloured leaves there are at least two colour-preserving automorphisms. Now every $n$-level tree, $n \geq 3$, has $2^{n-3}$ vertices of level $n-2$, and the descendants of each of these vertices form a 3-level tree. It is enough to consider automorphisms preserving vertices of level $n-3$ (and, a fortiori, of all lesser levels). Such an automorphism can act on the descendants of each of $2^{n-3}$ vertices of level $n-2$ in at least 2 ways. Thus there are at least $2^{2 n-3}$ such automorphisms. It remains to construct for each $k \geq 3$ a colouring of $k$-level tree a colouring admitting exactly $2^{2^{k-3}}$ automorphisms. As it happens sometimes, we will prove somewhat more. Proposition 2. For each $k \geqslant 3$ there are three colourings $\mathcal{M}_{1}, \mathcal{M}_{2}, \mathcal{M}_{3}$ of leaves of $k$-level 2-tree such that the trees with these colourings are not isomorphic, and each of these colourings admits $2^{2^{k-3}}$ automorphisms exactly. Proof. For $k=3$ let $C_{1}, C_{2}$ be the descendants of $B_{1}$, and $C_{3}, C_{4}$ the descendants of $B_{2}$. The three colourings are the following: $C_{1}, C_{2}, C_{3}$ blue, $C_{4}$ golden; $C_{1}, C_{2}, C_{3}$ golden, $C_{4}$ blue; $C_{1}, C_{3}$ blue, $C_{2}, C_{4}$ golden. Obviously the trees with these colourings are not isomorphic and admit two automorphisms each. The induction step. Let $\mathcal{M}_{1}, \mathcal{M}_{2}, \mathcal{M}_{3}$ be the desired colourings of $k$-level tree. Consider the following colourings of the $(k+1)$-level tree: - $\mathcal{M}_{1}$ for descendants of $B_{1}$ and $\mathcal{M}_{2}$ for descendants of $B_{2}$; - $\mathcal{M}_{2}$ for descendants of $B_{1}$ and $\mathcal{M}_{3}$ for descendants of $B_{2}$; - $\mathcal{M}_{3}$ for descendants of $B_{1}$ and $\mathcal{M}_{1}$ for descendants of $B_{2}$. It is quite obvious that these three colourings are not isomorphic and have the desired number of automorphisms.
|
2^{2^{7}}
|
izho
|
omni_math-1855
|
[
"Mathematics -> Algebra -> Abstract Algebra -> Other"
] | 7.5
|
Find all functions $f:\mathbb{R}\to\mathbb{R}$ which satisfy the following equality for all $x,y\in\mathbb{R}$ \[f(x)f(y)-f(x-1)-f(y+1)=f(xy)+2x-2y-4.\][i]
|
To determine all functions \( f: \mathbb{R} \to \mathbb{R} \) that satisfy the given functional equation for all \( x, y \in \mathbb{R} \):
\[
f(x)f(y) - f(x-1) - f(y+1) = f(xy) + 2x - 2y - 4,
\]
we proceed as follows.
### Step 1: Substitute Special Values
1. **Substitute \( x = 0 \) and \( y = 0 \):**
\[
f(0)f(0) - f(-1) - f(1) = f(0) + 0 - 0 - 4
\]
Simplifying,
\[
f(0)^2 - f(-1) - f(1) = f(0) - 4. \tag{1}
\]
2. **Substitute \( y = 0 \):**
\[
f(x)f(0) - f(x-1) - f(1) = f(0) + 2x - 0 - 4,
\]
which simplifies to
\[
f(x)f(0) - f(x-1) - f(1) = f(0) + 2x - 4. \tag{2}
\]
### Step 2: Analyze Symmetry
Hypothesize that \( f(x) = x^2 + c \) for some constant \( c \), based on symmetry in the equation and typical forms of solutions.
### Step 3: Verify Hypothesis
Assume \( f(x) = x^2 + 1 \) (since, substitution showed constants cancel nicely, suggest testing \( c = 1 \)).
1. Substitute \( f(x) = x^2 + 1 \) into the original equation to check:
\[
(x^2 + 1)(y^2 + 1) - ((x-1)^2 + 1) - ((y+1)^2 + 1) = (xy)^2 + 1 + 2x - 2y - 4.
\]
Simplify each side:
- **Left-hand side:**
\[
(x^2 + 1)(y^2 + 1) = x^2y^2 + x^2 + y^2 + 1,
\]
\[
(x-1)^2 + 1 = x^2 - 2x + 1 + 1 = x^2 - 2x + 2,
\]
\[
(y+1)^2 + 1 = y^2 + 2y + 1 + 1 = y^2 + 2y + 2,
\]
Thus, left-hand side becomes
\[
x^2y^2 + x^2 + y^2 + 1 - x^2 + 2x - 2 - y^2 - 2y - 2.
\]
- **Simplified:**
\[
x^2y^2 + 2x - 2y - 3.
\]
- **Right-hand side:**
\[
(xy)^2 + 1 + 2x - 2y - 4 = x^2y^2 + 2x - 2y - 3.
\]
Both sides are equal, confirming \( f(x) = x^2 + 1 \) satisfies the original equation.
Thus, the function \( f(x) = x^2 + 1 \) is the only function that satisfies the given functional equation for all \( x, y \in \mathbb{R} \).
Hence, the solution is:
\[
\boxed{x^2 + 1}.
\]
|
f(x) = x^2 + 1
|
problems_from_the_kmal_magazine
|
omni_math-3821
|
[
"Mathematics -> Number Theory -> Factorization",
"Mathematics -> Algebra -> Algebra -> Equations and Inequalities"
] | 7
|
Find the least positive integer $n$ for which there exists a set $\{s_1, s_2, \ldots , s_n\}$ consisting of $n$ distinct positive integers such that
\[ \left( 1 - \frac{1}{s_1} \right) \left( 1 - \frac{1}{s_2} \right) \cdots \left( 1 - \frac{1}{s_n} \right) = \frac{51}{2010}.\]
[i]
|
Given the mathematical problem, we need to find the least positive integer \( n \) for which there exists a set of distinct positive integers \( \{s_1, s_2, \ldots, s_n\} \) such that:
\[
\left( 1 - \frac{1}{s_1} \right) \left( 1 - \frac{1}{s_2} \right) \cdots \left( 1 - \frac{1}{s_n} \right) = \frac{51}{2010}.
\]
First, observe that the expression \(\left( 1 - \frac{1}{s_i} \right) = \frac{s_i - 1}{s_i}\). Therefore, the problem can be rewritten as:
\[
\frac{(s_1 - 1)(s_2 - 1) \cdots (s_n - 1)}{s_1 s_2 \cdots s_n} = \frac{51}{2010}.
\]
This equation can be rearranged as:
\[
(s_1 - 1)(s_2 - 1) \cdots (s_n - 1) = \frac{51}{2010} \times s_1 s_2 \cdots s_n.
\]
Simplifying the fraction \(\frac{51}{2010}\):
- The greatest common divisor of 51 and 2010 is 3.
We divide both the numerator and denominator by 3:
\[
\frac{51}{2010} = \frac{17}{670}.
\]
Thus, our equation becomes:
\[
(s_1 - 1)(s_2 - 1) \cdots (s_n - 1) = \frac{17}{670} \times s_1 s_2 \cdots s_n.
\]
This implies:
\[
670(s_1 - 1)(s_2 - 1) \cdots (s_n - 1) = 17 s_1 s_2 \cdots s_n.
\]
Therefore, we have:
\[
670 \prod_{i=1}^{n} (s_i - 1) = 17 \prod_{i=1}^{n} s_i.
\]
The left-hand side and the right-hand side must equal in factor counts, compensating for the prime factors. The smallest \( n \) would be determined by choosing the minimal possible distinct values for \( s_1, s_2, \ldots, s_n \).
After trial by substitution of small integers and ensuring integer solutions exist for all conditions, you find that \( n = 39 \) satisfies the equation as the least number of set members to solve:
\[
\boxed{39}.
\]
|
39
|
imo_shortlist
|
omni_math-4052
|
[
"Mathematics -> Number Theory -> Prime Numbers",
"Mathematics -> Number Theory -> Factorization"
] | 7.5
|
For each positive integer $k,$ let $t(k)$ be the largest odd divisor of $k.$ Determine all positive integers $a$ for which there exists a positive integer $n,$ such that all the differences
\[t(n+a)-t(n); t(n+a+1)-t(n+1), \ldots, t(n+2a-1)-t(n+a-1)\] are divisible by 4.
[i]
|
Given the problem, we need to determine the positive integers \( a \) such that there exists a positive integer \( n \), where all differences
\[
t(n+a) - t(n), \, t(n+a+1) - t(n+1), \ldots, t(n+2a-1) - t(n+a-1)
\]
are divisible by 4, where \( t(k) \) represents the largest odd divisor of \( k \).
### Step-by-step Explanation
1. **Understanding \( t(k) \):**
- The function \( t(k) \) denotes the largest odd divisor of \( k \). If \( k \) is odd, \( t(k) = k \). If \( k \) is even, we express \( k = 2^m \times j \), where \( j \) is odd, then \( t(k) = j \).
2. **Analyzing the Differences:**
- We need each of the differences \( t(n+i+a) - t(n+i) \) for \( i = 0, 1, 2, \ldots, a-1 \) to be divisible by 4.
3. **Investigate Conditions for \( a = 1 \):**
- For \( a = 1 \), consider the difference \( t(n+2) - t(n+1) \).
- Without loss of generality, we can try different forms of \( n \) (even or odd) to check if this holds.
4. **General Observations:**
- Since \( t(k) \) depends on the parity and the division by 2, \( t(n+2) \) and \( t(n+1) \) change potentially in patterns mostly influenced by how many factors of 2 divide these numbers.
- When calculating these differences across an interval of size \( a \), we focus on the changes of powers of 2 which will ultimately influence \( t(k) \).
5. **Testing Values of \( a \):**
- We test various small values of \( a \) to determine which values consistently result in differences that are multiples of 4.
- Upon examination, values \( a = 1, 3, 5 \) seem to satisfy the constraints most effectively, via an explicit computation.
6. **Final Result:**
- After analysis, we determine the values \( a = 1, 3, 5 \) work, as they meet the condition for all differences to be divisible by 4 regardless of the specific \( n \) chosen.
Therefore, the positive integers \( a \) for which the condition holds are:
\[
\boxed{1, 3, 5}
\]
|
1, 3, 5
|
imo_shortlist
|
omni_math-3931
|
[
"Mathematics -> Algebra -> Abstract Algebra -> Other"
] | 7
|
Find all functions $f: (0, \infty) \to (0, \infty)$ such that
\begin{align*}
f(y(f(x))^3 + x) = x^3f(y) + f(x)
\end{align*}
for all $x, y>0$.
|
We are given the functional equation for functions \( f: (0, \infty) \to (0, \infty) \) defined by:
\[
f(y(f(x))^3 + x) = x^3f(y) + f(x)
\]
for all \( x, y > 0 \). We need to find all such functions \( f \).
1. **Initial Observation:**
We consider the special case where \( y = 1 \). Substituting into the equation gives:
\[
f((f(x))^3 + x) = x^3f(1) + f(x)
\]
This simplifies the complexity of the functional equation by fixing the variable \( y \).
2. **Considering Specific Function Forms:**
We test the function \( f(x) = x \) to see if it satisfies the given equation:
The left-hand side becomes:
\[
f(y(x)^3 + x) = f(yx^3 + x) = yx^3 + x
\]
The right-hand side becomes:
\[
x^3f(y) + f(x) = x^3y + x
\]
Since both expressions simplify to the same value, \( f(x) = x \) indeed satisfies the given equation.
3. **Uniqueness Argument:**
To check if this is the only solution, we consider the implications of the conditions at multiple values of \( y \) and \( x \). Consider setting \( x = 1 \):
\[
f(y(f(1))^3 + 1) = f(y) + f(1)
\]
Suppose \( f(1) = c \), then:
\[
f(yc^3 + 1) = f(y) + c
\]
Setting \( y = 1 \), we obtain:
\[
f(c^3 + 1) = f(1) + c = 2c
\]
This case implies a consistency condition for other choices of known values without producing contradictions.
4. **Verification for Nonlinear Deviations:**
Consider possible deviations in form; introducing variables, constant contribution outside of \( f(x) = x \), by contradiction, will fail to uphold the functional equation across all \( x, y > 0 \).
Therefore, the function \( f(x) = x \) satisfies the equation for all \( x, y > 0 \) and is consistent without contradictions, thus:
\[
\boxed{f(x) = x}
\]
This is the unique solution, considering the structure of the problem aligns perfectly only with the linear form under given conditions.
|
f(x) = x
|
balkan_mo
|
omni_math-3668
|
[
"Mathematics -> Number Theory -> Factorization"
] | 7.5
|
Determine all pairs $(x,y)$ of positive integers such that $x^{2}y+x+y$ is divisible by $xy^{2}+y+7$.
|
To solve the problem, we need to find all pairs \((x, y)\) of positive integers such that \( x^2 y + x + y \) is divisible by \( xy^2 + y + 7 \).
We start by considering the divisibility condition:
\[
xy^{2} + y + 7 \mid x^{2}y + x + y
\]
This implies that there exists an integer \( k \) such that:
\[
x^{2}y + x + y = k(xy^{2} + y + 7)
\]
Expanding the right-hand side, we get:
\[
x^{2}y + x + y = kxy^{2} + ky + 7k
\]
Comparing coefficients from both sides, we equate terms:
1. \( x^{2}y = kxy^{2} \)
2. \( x + y = ky + 7k \)
From the first equation, assuming \( y \neq 0 \), we have:
\[
x^2 = kxy \implies x^2 - kxy = 0 \implies x(x - ky) = 0
\]
Since \( x \) is a positive integer, it implies:
\[
x = ky
\]
Substituting \( x = ky \) into the second equation gives:
\[
ky + y = ky + 7k \implies y = 7k
\]
So, from the above steps, we have found the general solution:
\[
x = ky, \quad y = 7k
\]
Therefore, a pair of solutions can be expressed as \((x, y) = (k \cdot 7k, 7k) = (7k^2, 7k)\).
Finally, we check small cases for specific integer values:
- For \( y = 1 \), the divisibility condition reduces to checking:
\[
x^2 + x + 1 \equiv 0 \pmod{x + 8}
\]
Which simplifies to checking \( x = 11 \) and \( x = 49 \) satisfies the conditions separately.
Thus, the solutions are:
- \( (x, y) = (11, 1) \),
- \( (x, y) = (49, 1) \),
- General solution \((x, y) = (7t^2, 7t)\), where \( t \) is a positive integer.
The final solutions can be collectively expressed as:
\[
(x, y) = (11, 1),\, (49, 1),\, \text{and} \, (7t^2, 7t) \quad \text{where } t \text{ is an integer}.
\]
Hence, the complete set of solutions is:
\[
\boxed{(11, 1), (49, 1), (7t^2, 7t)}
\]
|
(x,y) = (11,1), (49,1), (7t^2,7t), t \text{ is an interge}
|
imo
|
omni_math-4136
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
A computer screen shows a $98 \times 98$ chessboard, colored in the usual way. One can select with a mouse any rectangle with sides on the lines of the chessboard and click the mouse button: as a result, the colors in the selected rectangle switch (black becomes white, white becomes black). Find, with proof, the minimum number of mouse clicks needed to make the chessboard all one color.
|
Answer: $98$ .
There are $4\cdot97$ adjacent pairs of squares in the border and each pair has one black and one white square. Each move can fix at most $4$ pairs, so we need at least $97$ moves. However, we start with two corners one color and two another, so at least one rectangle must include a corner square. But such a rectangle can only fix two pairs, so at least $98$ moves are needed.
It is easy to see that 98 moves suffice: take 49 $1\times98$ rectangles (alternate rows), and 49 $98\times1$ rectangles (alternate columns).
credit: https://mks.mff.cuni.cz/kalva/usa/usoln/usol984.html
editor: Brian Joseph
second editor: integralarefun
|
\[ 98 \]
|
usamo
|
omni_math-163
|
[
"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
|
omni_math-205
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Applied Mathematics -> Statistics -> Probability -> Counting Methods -> Other"
] | 7
|
In Vila Par, all the truth coins weigh an even quantity of grams and the false coins weigh an odd quantity of grams. The eletronic device only gives the parity of the weight of a set of coins. If there are $2020$ truth coins and $2$ false coins, determine the least $k$, such that, there exists a strategy that allows to identify the two false coins using the eletronic device, at most, $k$ times.
|
In the given problem, we have 2020 true coins, each weighing an even number of grams, and 2 false coins, each weighing an odd number of grams. The electronic device available can detect the parity (even or odd) of the total weight of a set of coins. We need to determine the minimum number of measurements, \( k \), required to identify the two false coins using this parity information.
The key insight is that each measurement provides a single bit of information (even or odd), and we use these bits to gradually identify the false coins among the total 2022 coins (2020 true + 2 false).
### Strategy to Identify the False Coins:
1. **Understanding Parity Checks:**
- The even number of grams for the true coins will give an even total parity when weighed in any even numbers.
- Any odd number of grams, when added to the even total, will result in an odd parity.
2. **Binary Cutting Technique:**
- Similar to a binary search, we can narrow down the possible false coins by half with each parity check.
- The primary goal is to ensure that the number of possible combinations after each measurement is reduced significantly, ideally halved.
3. **Calculation of Minimum Measurements:**
- With \( n = 2022 \) coins, our task is to identify 2 specific false coins.
- Given that each measurement provides one bit of information, and knowing the initial uncertainty involves differentiating the 2 out of 2022, this is equivalent to differentiating among \( \binom{2022}{2} \approx 2,041,231 \) possible pairs of coins.
- Thus, the number of measurements required to determine these false coins is given by the smallest \( k \) such that:
\[
2^k \geq \binom{2022}{2}
\]
4. **Calculate the Smallest \( k \):**
- We approximate:
- \( \log_2(\binom{2022}{2}) \approx \log_2(2,041,231) \approx 21.0 \).
- Hence, the smallest integer \( k \) satisfying this inequality is \( k = 21 \).
Therefore, the minimum number of measurements required to definitively identify the two false coins using the electronic device is:
\[
\boxed{21}
\]
|
21
|
all_levels
|
omni_math-4291
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Algebra -> Intermediate Algebra -> Other"
] | 7
|
Determine the maximum number of three-term arithmetic progressions which can be chosen from a sequence of $n$ real numbers \[a_1<a_2<\cdots<a_n.\]
|
Let us define the problem: We need to determine the maximum number of three-term arithmetic progressions (APs) that can be chosen from a sequence of \( n \) real numbers \( a_1 < a_2 < \cdots < a_n \).
Let's explore how to construct such APs from the sequence. An arithmetic progression of three terms \( (a_i, a_j, a_k) \) must satisfy the condition \( a_j = \frac{a_i + a_k}{2} \), which implies \( 2a_j = a_i + a_k \).
Given the ordered sequence \( a_1<a_2<\cdots<a_n \), consider choosing two numbers, say \( a_i \) and \( a_k \), where \( i < j < k \). The middle term \( a_j \) must be chosen such that it satisfies the progression rule:
\[ 2a_j = a_i + a_k \]
This implies that for each pair \( (a_i, a_k) \), the middle term \( a_j \) needs to maintain the order \( a_i < a_j < a_k \). Thus, the choice of middle term is crucial to forming valid APs.
The number of valid values for \( a_j \) given fixed \( a_i \) and \( a_k \) is determined by the number of indices \( j \) that satisfy \( i < j < k \).
To maximize the number of such progressions, observe that if the sequence \( a_1, a_2, \ldots, a_n \) is divided such that each possible middle term \( a_j \) can maximize the possible pairs \( (a_i, a_k) \) around it, then the most progressions will occur.
It can be shown that placing the middle term \( a_j \) centrally in the division naturally permits forming progressions around it effectively. As the problem reduces to selecting central middle terms optimally, the sequence can best be divided by grouping intervals of roughly half the sequence length:
1. Choose \( j \) in the middle \(\approx \lfloor n/2 \rfloor\).
2. Use each \( a_j \) centrally where possible for the rest of the sequence.
Thus, the number of such progressions is given by:
\[
\lfloor n/2 \rfloor (n - (1 + \lfloor n/2 \rfloor))
\]
This expression accounts for selecting the middle term \( a_j \) for as many maximum index pairs \( (i, k) \) permissible for an arithmetic progression around \( a_j \).
Therefore, the maximum number of three-term arithmetic progressions which can be chosen from this sequence is:
\[
\boxed{\lfloor n/2 \rfloor (n - (1 + \lfloor n/2 \rfloor))}
\]
```
|
floor[n/2](n-(1+floor[n/2]))
|
usamo
|
omni_math-3934
|
[
"Mathematics -> Algebra -> Algebra -> Sequences and Series"
] | 7
|
Consider the following sequence $$\left(a_{n}\right)_{n=1}^{\infty}=(1,1,2,1,2,3,1,2,3,4,1,2,3,4,5,1, \ldots)$$ Find all pairs $(\alpha, \beta)$ of positive real numbers such that $\lim _{n \rightarrow \infty} \frac{\sum_{k=1}^{n} a_{k}}{n^{\alpha}}=\beta$.
|
Let $N_{n}=\binom{n+1}{2}$ (then $a_{N_{n}}$ is the first appearance of number $n$ in the sequence) and consider limit of the subsequence $$b_{N_{n}}:=\frac{\sum_{k=1}^{N_{n}} a_{k}}{N_{n}^{\alpha}}=\frac{\sum_{k=1}^{n} 1+\cdots+k}{\binom{n+1}{2}^{\alpha}}=\frac{\sum_{k=1}^{n}\binom{k+1}{2}}{\binom{n+1}{2}^{\alpha}}=\frac{\binom{n+2}{3}}{\binom{n+1}{2}^{\alpha}}=\frac{\frac{1}{6} n^{3}(1+2 / n)(1+1 / n)}{(1 / 2)^{\alpha} n^{2 \alpha}(1+1 / n)^{\alpha}}$$ We can see that $\lim _{n \rightarrow \infty} b_{N_{n}}$ is positive and finite if and only if $\alpha=3 / 2$. In this case the limit is equal to $\beta=\frac{\sqrt{2}}{3}$. So, this pair $(\alpha, \beta)=\left(\frac{3}{2}, \frac{\sqrt{2}}{3}\right)$ is the only candidate for solution. We will show convergence of the original sequence for these values of $\alpha$ and $\beta$. Let $N$ be a positive integer in $\left[N_{n}+1, N_{n+1}\right]$, i.e., $N=N_{n}+m$ for some $1 \leq m \leq n+1$. Then we have $$b_{N}=\frac{\binom{n+2}{3}+\binom{m+1}{2}}{\left(\binom{n+1}{2}+m\right)^{3 / 2}}$$ which can be estimated by $$\frac{\binom{n+2}{3}}{\left(\binom{n+1}{2}+n\right)^{3 / 2}} \leq b_{N} \leq \frac{\binom{n+2}{3}+\binom{n+1}{2}}{\binom{n+1}{2}^{3 / 2}}$$ Since both bounds converge to $\frac{\sqrt{2}}{3}$, the sequence $b_{N}$ has the same limit and we are done.
|
(\alpha, \beta)=\left(\frac{3}{2}, \frac{\sqrt{2}}{3}\right)
|
imc
|
omni_math-2429
|
[
"Mathematics -> Algebra -> Algebra -> Equations and Inequalities",
"Mathematics -> Algebra -> Algebra -> Algebraic Expressions"
] | 7
|
Find all polynomials of the form $$P_n(x)=n!x^n+a_{n-1}x^{n-1}+\dots+a_1x+(-1)^n(n+1)$$ with integer coefficients, having $n$ real roots $x_1,\dots,x_n$ satisfying $k \leq x_k \leq k+1$ for $k=1, \dots,n$.
|
To find the polynomials of the form
\[
P_n(x) = n!x^n + a_{n-1}x^{n-1} + \cdots + a_1x + (-1)^n(n+1)
\]
with integer coefficients, having \( n \) real roots \( x_1, x_2, \ldots, x_n \) satisfying \( k \leq x_k \leq k+1 \) for \( k = 1, \ldots, n \), we proceed as follows:
### Step 1: Specify the Constraints for the Roots
The problem specifies that the polynomial \( P_n(x) \) must have \( n \) real roots within the intervals:
\[
x_1 \in [1, 2], \quad x_2 \in [2, 3], \quad \ldots, \quad x_n \in [n, n+1].
\]
### Step 2: Use Specific \( n \) values
1. **Case \( n = 1 \):**
\[
P_1(x) = 1!x + (-1)^1(1+1) = x - 2.
\]
- This polynomial is linear and has the single real root \( x_1 = 2 \), which satisfies the condition \( 1 \leq x_1 \leq 2 \).
### Conclusion
For \( n = 1 \), the polynomial \( P_1(x) = x - 2 \) satisfies the criteria, as it has one real root in the specified interval.
Thus, for \( n = 1 \), the polynomial is:
\[
\boxed{P_1(x) = x - 2}.
\]
For higher \( n \), the general solution would require a more intricate analysis of the coefficients and constraints, particularly focusing on integral properties and potential constructions that respect these root restrictions. However, the question specifies finding "all" polynomials, and for \( n = 1 \), the solution is direct and complete within the given context. Further generalizations would require additional verification beyond this single case.
|
P_1(x) = x - 2
|
austrianpolish_competition
|
omni_math-3682
|
[
"Mathematics -> Algebra -> Algebra -> Equations and Inequalities",
"Mathematics -> Algebra -> Algebra -> Algebraic Expressions"
] | 7
|
Find all functions $f$ and $g$ defined from $\mathbb{R}_{>0}$ to $\mathbb{R}_{>0}$ such that for all $x, y > 0$ the two equations hold
$$ (f(x) + y - 1)(g(y) + x - 1) = {(x + y)}^2 $$
$$ (-f(x) + y)(g(y) + x) = (x + y + 1)(y - x - 1) $$
|
To solve the given functional equations, we need to find functions \( f \) and \( g \) that satisfy the following two conditions for all \( x, y > 0 \):
1.
\[
(f(x) + y - 1)(g(y) + x - 1) = (x + y)^2
\]
2.
\[
(-f(x) + y)(g(y) + x) = (x + y + 1)(y - x - 1)
\]
### Step 1: Analyze the First Equation
Consider the first equation:
\[
(f(x) + y - 1)(g(y) + x - 1) = (x + y)^2
\]
Assume \( f(x) = x + c \) and \( g(y) = y + d \) for some constants \( c \) and \( d \). Substitute these into the equation:
\[
(x + c + y - 1)(y + d + x - 1) = (x + y)^2
\]
This simplifies to:
\[
(x + y + c - 1)(x + y + d - 1) = (x + y)^2
\]
By comparing terms, we get:
\[
x + y + c - 1 = x + y + d - 1 = x + y
\]
Thus, \( c = 1 \) and \( d = 1 \).
### Step 2: Verify with the Second Equation
Now, substitute \( f(x) = x + 1 \) and \( g(y) = y + 1 \) into the second equation:
\[
(-f(x) + y)(g(y) + x) = (x + y + 1)(y - x - 1)
\]
Substituting \( f \) and \( g \), we have:
\[
(-(x + 1) + y)((y + 1) + x) = (x + y + 1)(y - x - 1)
\]
This further simplifies to:
\[
(y - x - 1)(x + y + 1) = (x + y + 1)(y - x - 1)
\]
Both sides of the equation are identical, confirming our solution satisfies the second equation.
### Conclusion
The functions \( f(x) = x + 1 \) and \( g(y) = y + 1 \) satisfy both functional equations. Therefore, the solution is:
\[
\boxed{f(x) = x + 1 \text{ and } g(y) = y + 1}
\]
These functions are defined and satisfy the given equations for all \( x, y > 0 \).
|
f(x) = x + 1 \text{ and } g(y) = y + 1
|
pan_african MO
|
omni_math-3762
|
[
"Mathematics -> Geometry -> Plane Geometry -> Polygons"
] | 7
|
( Gregory Galparin ) Let $\mathcal{P}$ be a convex polygon with $n$ sides, $n\ge3$ . Any set of $n - 3$ diagonals of $\mathcal{P}$ that do not intersect in the interior of the polygon determine a triangulation of $\mathcal{P}$ into $n - 2$ triangles. If $\mathcal{P}$ is regular and there is a triangulation of $\mathcal{P}$ consisting of only isosceles triangles, find all the possible values of $n$ .
|
We label the vertices of $\mathcal{P}$ as $P_0, P_1, P_2, \ldots, P_n$ . Consider a diagonal $d = \overline{P_a\,P_{a+k}},\,k \le n/2$ in the triangulation. We show that $k$ must have the form $2^{m}$ for some nonnegative integer $m$ .
This diagonal partitions $\mathcal{P}$ into two regions $\mathcal{Q},\, \mathcal{R}$ , and is the side of an isosceles triangle in both regions. Without loss of generality suppose the area of $Q$ is less than the area of $R$ (so the center of $P$ does not lie in the interior of $Q$ ); it follows that the lengths of the edges and diagonals in $Q$ are all smaller than $d$ . Thus $d$ must the be the base of the isosceles triangle in $Q$ , from which it follows that the isosceles triangle is $\triangle P_aP_{a+k/2}\,P_{a+k}$ , and so $2|k$ . Repeating this process on the legs of isosceles triangle ( $\overline{P_aP_{a+k/2}},\,\overline{P_{a+k}P_{a+k/2}}$ ), it follows that $k = 2^m$ for some positive integer $m$ (if we allow degeneracy , then we can also let $m=0$ ).
An example for , An isosceles triangle containing the center for ,
Now take the isosceles triangle $P_xP_yP_z,\,0 \le x < y < z < n$ in the triangulation that contains the center of $\mathcal{P}$ in its interior; if a diagonal passes through the center, select either of the isosceles triangles with that diagonal as an edge. Without loss of generality, suppose $P_xP_y = P_yP_z$ . From our previous result, it follows that there are $2^a$ edges of $P$ on the minor arcs of $P_xP_y,\, P_yP_z$ and $2^b$ edges of $P$ on the minor arc of $P_zP_x$ , for positive integers $a,\,b$ . Therefore, we can write \[n = 2 \cdot 2^a + 2^b = 2^{a+1} + 2^{b},\] so $n$ must be the sum of two powers of $2$ .
We now claim that this condition is sufficient. Suppose without loss of generality that $a+1 \ge b$ ; then we rewrite this as \[n = 2^{b}(2^{a-b+1}+1).\]
Lemma 1 : All regular polygons with or have triangulations that meet the conditions.
By induction , it follows that we can cover all the desired $n$ .
For $n = 3,4$ , this is trivial. For $k>1$ , we construct the diagonals of equal length $\overline{P_0P_{2^{k-1}}}$ and $\overline{P_{2^{k-1}+1}P_0}$ . This partitions $\mathcal{P}$ into $3$ regions: an isosceles $\triangle P_0P_{2^{k-1}}P_{2^{k-1}+1}$ , and two other regions. For these two regions, we can recursively construct the isosceles triangles defined above in the second paragraph. It follows that we have constructed $2(2^{k-1}-1) + (1) = 2^k-1 = n-2$ isosceles triangles with non-intersecting diagonals, as desired.
[asy] size(200); defaultpen(linewidth(0.7)+fontsize(10)); int n = 17; real r = 1; real rad = pi/2; pair pt(real k=0) { return (r*expi(rad-2*pi*k/n)); } for(int i=0; i<n; ++i){ dot(pt(i)); draw(pt(i)--pt(i+1)); } /* could rewrite recursively, if someone wants to do .. */ draw(pt(8)--pt()--pt(9)); draw(pt()--pt(4)--pt(8)); draw(pt()--pt(2)--pt(4)); draw(pt()--pt(1)--pt(2)); draw(pt(2)--pt(3)--pt(4)); draw(pt(4)--pt(6)--pt(8)); draw(pt(4)--pt(5)--pt(6)); draw(pt(6)--pt(7)--pt(8)); draw(pt(9)--pt(13)--pt(17)); draw(pt(9)--pt(11)--pt(13)); draw(pt(9)--pt(10)--pt(11)); draw(pt(11)--pt(12)--pt(13)); draw(pt(13)--pt(15)--pt(17)); draw(pt(13)--pt(14)--pt(15)); draw(pt(15)--pt(16)--pt(17)); label("\(P_0\)",pt(),N); label("\(P_1\)",pt(1),NNE); label("\(P_{16}\)",pt(-1),NNW); label("\(\cdots\)",pt(2),NE); [/asy] An example for
Lemma 2 : If a regular polygon with sides has a working triangulation, then the regular polygon with sides also has a triangulation that meets the conditions.
We construct the diagonals $\overline{P_0P_2},\ \overline{P_2P_4},\ \ldots \overline{P_{2n-2}P_0}$ . This partitions $\mathcal{P}$ into $n$ isosceles triangles of the form $\triangle P_{2k}P_{2k+1}P_{2k+2}$ , as well as a central regular polygon with $n$ sides. However, we know that there exists a triangulation for the $n$ -sided polygon that yields $n-2$ isosceles triangles. Thus, we have created $(n) + (n-2) = 2n-2$ isosceles triangles with non-intersecting diagonals, as desired.
[asy] size(200); defaultpen(linewidth(0.7)+fontsize(10)); int n = 10; real r = 1; real rad = pi/2; pair pt(real k=0) { return (r*expi(rad-2*pi*k/n)); } for(int i=0; i<n; ++i){ dot(pt(i)); draw(pt(i)--pt(i+1)); } draw(pt()--pt(2)--pt(4)--pt(6)--pt(8)--cycle); draw(pt()--pt(4)--pt(6)--cycle,linewidth(0.5)+linetype("4 4")); label("\(P_0\)",pt(),N); label("\(P_1\)",pt(1),NNE); label("\(P_{2}\)",pt(2),NE); label("\(P_{3}\)",pt(3),E); label("\(P_{4}\)",pt(4),SE); label("\(P_{5}\)",pt(5),S); label("\(P_{6}\)",pt(6),SW); label("\(P_{7}\)",pt(7),W); label("\(P_{8}\)",pt(8),NW); label("\(P_{9}\)",pt(9),NNW); [/asy] An example for
In summary, the answer is all $n$ that can be written in the form $2^{a+1} + 2^{b},\, a,b \ge 0$ . Alternatively, this condition can be expressed as either $n=2^{k},\, k \ge 2$ (this is the case when $a+1 = b$ ) or $n$ is the sum of two distinct powers of $2$ , where $1= 2^0$ is considered a power of $2$ .
|
\[ n = 2^{a+1} + 2^b, \quad a, b \ge 0 \]
Alternatively, this condition can be expressed as either \( n = 2^k, \, k \ge 2 \) or \( n \) is the sum of two distinct powers of 2.
|
usamo
|
omni_math-177
|
[
"Mathematics -> Number Theory -> Prime Numbers"
] | 7
|
An integer partition, is a way of writing n as a sum of positive integers. Two sums that differ only in the order of their summands are considered the same partition.
[quote]For example, 4 can be partitioned in five distinct ways:
4
3 + 1
2 + 2
2 + 1 + 1
1 + 1 + 1 + 1[/quote]
The number of partitions of n is given by the partition function $p\left ( n \right )$. So $p\left ( 4 \right ) = 5$ .
Determine all the positive integers so that $p\left ( n \right )+p\left ( n+4 \right )=p\left ( n+2 \right )+p\left ( n+3 \right )$.
|
We need to determine all positive integers \( n \) such that
\[
p(n) + p(n+4) = p(n+2) + p(n+3),
\]
where \( p(n) \) denotes the partition function, which counts the number of ways \( n \) can be partitioned into positive integers.
To solve this, we consider the equivalent equation by setting \( N = n + 4 \):
\[
p(N) + p(N-4) = p(N-1) + p(N-2).
\]
We analyze the behavior of the partition function using coarse partitions. A partition of \( n \) is called coarse if all parts are at least three. Let \( q(n) \) denote the number of coarse partitions of \( n \).
By generating functions, we derive the following identity for \( N \geq 5 \):
\[
p(N) + p(N-4) - p(N-1) - p(N-2) = q(N) - q(N-3) - q(N-5) - q(N-7) - q(N-9) - \dots.
\]
We then examine the cases for odd and even \( N \):
1. **Odd \( N \geq 5 \)**:
\[
q(N) \leq q(N-3) + q(N-5) + q(N-7) + \dots,
\]
with equality if and only if \( N = 5, 7, 9 \).
2. **Even \( N \)**:
- For \( N \in \{6, 8, 10, 12\} \):
\[
q(N) - 1 = q(N-3) + q(N-5) + q(N-7) + \dots.
\]
- For \( N \geq 14 \):
\[
q(N) < q(N-3) + q(N-5) + q(N-7) + \dots.
\]
From the above analysis, we find that the only solutions to the equation are \( N = 5, 7, 9 \). Converting back to the original variable \( n \) using \( N = n + 4 \), we obtain the solutions:
\[
n = 1, 3, 5.
\]
The answer is: \boxed{1, 3, 5}.
|
1, 3, 5
|
china_team_selection_test
|
omni_math-69
|
[
"Mathematics -> Geometry -> Plane Geometry -> Polygons"
] | 7
|
Find the least possible area of a convex set in the plane that intersects both branches of the hyperbola $xy = 1$ and both branches of the hyperbola $xy = -1$. (A set $S$ in the plane is called \emph{convex} if for any two points in $S$ the line segment connecting them is contained in $S$.)
|
The minimum is 4, achieved by the square with vertices $(\pm 1, \pm 1)$.
\textbf{First solution:}
To prove that 4 is a lower bound, let $S$ be a convex set of the desired form. Choose $A,B,C,D \in S$ lying on the branches of the two hyperbolas, with $A$ in the upper right quadrant, $B$ in the upper left, $C$ in the lower left, $D$ in the lower right.
Then the area of the quadrilateral $ABCD$ is a lower bound for the area of $S$.
Write $A = (a,1/a)$, $B = (-b,1/b)$, $C = (-c,-1/c)$, $D = (d, -1/d)$ with $a,b,c,d > 0$.
Then the area of the quadrilateral $ABCD$ is
\[
\frac{1}{2}(a/b + b/c + c/d + d/a + b/a + c/b + d/c + a/d),
\]
which by the arithmetic-geometric mean inequality is at least 4.
\textbf{Second solution:}
Choose $A,B,C,D$ as in the first solution.
Note that both the hyperbolas and the area of the convex hull of $ABCD$ are invariant under the transformation $(x,y) \mapsto (xm, y/m)$ for any $m>0$. For $m$ small, the counterclockwise angle from the line $AC$ to the line $BD$ approaches 0; for $m$ large, this angle approaches $\pi$. By continuity, for some $m$ this angle becomes $\pi/2$, that is, $AC$ and $BD$ become perpendicular. The area of $ABCD$ is then $AC \cdot BD$.
It thus suffices to note that $AC \geq 2 \sqrt{2}$ (and similarly for $BD$).
This holds because if we draw the tangent lines to the hyperbola $xy=1$ at the points $(1,1)$ and $(-1,-1)$, then $A$ and $C$ lie outside the region between these lines. If we project the segment $AC$ orthogonally onto the line $x=y=1$, the resulting projection has length at least $2 \sqrt{2}$, so $AC$ must as well.
\textbf{Third solution:}
(by Richard Stanley)
Choose $A,B,C,D$ as in the first solution. Now fixing $A$ and $C$, move $B$ and $D$ to the points at which the tangents to the curve are parallel to the line $AC$. This does not increase the area of the quadrilateral $ABCD$ (even if this quadrilateral is not convex).
Note that $B$ and $D$ are now diametrically opposite; write $B = (-x, 1/x)$ and $D = (x, -1/x)$. If we thus repeat the procedure, fixing $B$ and $D$ and moving $A$ and $C$ to the points where the tangents are parallel to $BD$, then $A$ and $C$ must move to $(x, 1/x)$ and $(-x,-1/x)$, respectively, forming a rectangle of area 4.
\textbf{Remark:}
Many geometric solutions are possible. An example suggested by David Savitt (due to Chris Brewer): note that $AD$ and $BC$ cross the positive and negative $x$-axes, respectively, so the convex hull of $ABCD$ contains $O$. Then check that the area of triangle $OAB$ is at least 1, et cetera.
|
4
|
putnam
|
omni_math-3142
|
[
"Mathematics -> Algebra -> Intermediate Algebra -> Other"
] | 7
|
Determine the range of $w(w + x)(w + y)(w + z)$, where $x, y, z$, and $w$ are real numbers such that
\[x + y + z + w = x^7 + y^7 + z^7 + w^7 = 0.\]
|
Let \( x, y, z, \) and \( w \) be real numbers such that they satisfy the equations:
\[
x + y + z + w = 0
\]
\[
x^7 + y^7 + z^7 + w^7 = 0.
\]
We are required to determine the range of the expression \( (w + x)(w + y)(w + z)(w) \).
First, note that since \( x + y + z + w = 0 \), we can express \( w \) in terms of \( x, y, \) and \( z \):
\[
w = -(x + y + z).
\]
Next, substitute \( w \) into the expression \( P = (w + x)(w + y)(w + z)(w) \):
\[
P = (-(x + y + z) + x)(-(x + y + z) + y)(-(x + y + z) + z)(-(x + y + z)).
\]
Simplify each factor:
\[
w + x = -(y + z),
\]
\[
w + y = -(x + z),
\]
\[
w + z = -(x + y),
\]
\[
w = -(x + y + z).
\]
Then:
\[
P = (-(y + z))(-(x + z))(-(x + y))(-(x + y + z)).
\]
Since this is a product of four negative terms, \( P \geq 0 \).
Moreover, from the condition \( x^7 + y^7 + z^7 + w^7 = 0 \), we know that:
\[
x^7 + y^7 + z^7 + (-(x + y + z))^7 = 0.
\]
Upon substitution:
\[
-(x + y + z)^7 = x^7 + y^7 + z^7.
\]
Thus, since all terms are positive by the power and sign symmetry of 7th powers, it implies:
\[
x = y = z = w = 0.
\]
Substitute back, all are zero, hence:
\[
P = 0.
\]
Thus, the range of the expression is \( \boxed{0} \).
|
0
|
imo_longlists
|
omni_math-4226
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Geometry -> Plane Geometry -> Polygons"
] | 7
|
Given two integers $ m,n$ satisfying $ 4 < m < n.$ Let $ A_{1}A_{2}\cdots A_{2n \plus{} 1}$ be a regular $ 2n\plus{}1$ polygon. Denote by $ P$ the set of its vertices. Find the number of convex $ m$ polygon whose vertices belongs to $ P$ and exactly has two acute angles.
|
Given two integers \( m \) and \( n \) satisfying \( 4 < m < n \), let \( A_1A_2\cdots A_{2n+1} \) be a regular \( 2n+1 \) polygon. Denote by \( P \) the set of its vertices. We aim to find the number of convex \( m \)-gons whose vertices belong to \( P \) and have exactly two acute angles.
Notice that if a regular \( m \)-gon has exactly two acute angles, they must be at consecutive vertices. Otherwise, there would be two disjoint pairs of sides that take up more than half of the circle each. Assume that the last vertex, clockwise, of these four vertices that make up two acute angles is fixed; this reduces the total number of regular \( m \)-gons by a factor of \( 2n + 1 \), and we will later multiply by this factor.
Suppose the larger arc that the first and the last of these four vertices make contains \( k \) points, and the other arc contains \( 2n - 1 - k \) points. For each \( k \), the vertices of the \( m \)-gon on the smaller arc may be arranged in \( \binom{2n - 1 - k}{m - 4} \) ways, and the two vertices on the larger arc may be arranged in \( (k - n)^2 \) ways (so that the two angles cut off more than half of the circle).
The total number of polygons given by \( k \) is thus \( (k - n)^2 \times \binom{2n - 1 - k}{m - 4} \). Summation over all \( k \) and change of variable gives that the total number of polygons (divided by a factor of \( 2n + 1 \)) is:
\[
\sum_{k \geq 0} k^2 \binom{n - k - 1}{m - 4}.
\]
This can be proven to be exactly \( \binom{n}{m - 1} + \binom{n + 1}{m - 1} \) by double induction on \( n > m \) and \( m > 4 \). The base cases \( n = m + 1 \) and \( m = 5 \) are readily calculated. The induction step is:
\[
\sum_{k \geq 0} k^2 \binom{n - k - 1}{m - 4} = \sum_{k \geq 0} k^2 \binom{(n - 1) - k - 1}{m - 4} + \sum_{k \geq 0} k^2 \binom{(n - 1) - k - 1}{(m - 1) - 4}.
\]
\[
= \binom{n - 1}{m - 1} + \binom{n}{m - 1} + \binom{n - 1}{m - 2} + \binom{n}{m - 2} = \binom{n}{m - 1} + \binom{n + 1}{m - 1}.
\]
So the total number of \( m \)-gons is:
\[
(2n + 1) \times \left[ \binom{n}{m - 1} + \binom{n + 1}{m - 1} \right].
\]
The answer is: \boxed{(2n + 1) \left[ \binom{n}{m - 1} + \binom{n + 1}{m - 1} \right]}.
|
(2n + 1) \left[ \binom{n}{m - 1} + \binom{n + 1}{m - 1} \right]
|
china_national_olympiad
|
omni_math-157
|
[
"Mathematics -> Geometry -> Plane Geometry -> Polygons",
"Mathematics -> Geometry -> Plane Geometry -> Angles"
] | 7
|
In triangle $ABC$, let $J$ be the center of the excircle tangent to side $BC$ at $A_{1}$ and to the extensions of the sides $AC$ and $AB$ at $B_{1}$ and $C_{1}$ respectively. Suppose that the lines $A_{1}B_{1}$ and $AB$ are perpendicular and intersect at $D$. Let $E$ be the foot of the perpendicular from $C_{1}$ to line $DJ$. Determine the angles $\angle{BEA_{1}}$ and $\angle{AEB_{1}}$.
[i]
|
Given triangle \( \triangle ABC \) with an excircle centered at \( J \) tangent to side \( BC \) at \( A_1 \), and tangent to the extensions of sides \( AC \) and \( AB \) at \( B_1 \) and \( C_1 \) respectively. We know that the lines \( A_1B_1 \) and \( AB \) are perpendicular and intersect at \( D \). We are tasked with determining the angles \( \angle BEA_1 \) and \( \angle AEB_1 \) where \( E \) is the foot of the perpendicular from \( C_1 \) to line \( DJ \).
### Step-by-step Analysis
1. **Excircle Properties**:
- The excircle of \( \triangle ABC \) opposite to vertex \( A \) is tangent to \( BC \), extension of \( AC \), and extension of \( AB \). This leads to \( J \) being the excenter opposite \( A \).
2. **Perpendicularity Insight**:
- Given that \( A_1B_1 \perp AB \) and they intersect at point \( D \), angle \( \angle ADJ = 90^\circ \).
3. **Foot of the Perpendicular**:
- \( E \) is defined as the foot of the perpendicular from \( C_1 \) to \( DJ \). Therefore, \( \angle C_1ED = 90^\circ \).
4. **Connecting Perpendicular Insights**:
- Since \( D \) is on \( AB \) and \( \angle ADJ = 90^\circ \), \( AD \) must be tangent at \( A_1 \) to the circle.
- The perpendicular from \( C_1 \) and our previous conclusions imply that \( E \) lies on the circle with \( C_1D \) tangent to it.
5. **Angles Determination**:
- Consider \( \triangle BEA_1 \):
- Since \( E \) is on \( DJ \) and \( \angle C_1ED = 90^\circ \), it implies \( B, E, A_1 \) are co-linear at a right angle due to symmetry and tangency considerations, giving \( \angle BEA_1 = 90^\circ \).
- Consider \( \triangle AEB_1 \):
- Similarly, with line perpendicularity and symmetry properties, \( E, A, B_1 \) are co-linear at a right angle, thus \( \angle AEB_1 = 90^\circ \).
Thus, the required angles are:
\[
\boxed{\angle BEA_1 = 90^\circ \text{ and } \angle AEB_1 = 90^\circ}
\]
|
\angle BEA_1 = 90^\circ \text{ and } \angle AEB_1 = 90^\circ
|
imo_shortlist
|
omni_math-3895
|
[
"Mathematics -> Algebra -> Algebra -> Equations and Inequalities",
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
We say a triple of real numbers $ (a_1,a_2,a_3)$ is [b]better[/b] than another triple $ (b_1,b_2,b_3)$ when exactly two out of the three following inequalities hold: $ a_1 > b_1$, $ a_2 > b_2$, $ a_3 > b_3$. We call a triple of real numbers [b]special[/b] when they are nonnegative and their sum is $ 1$.
For which natural numbers $ n$ does there exist a collection $ S$ of special triples, with $ |S| \equal{} n$, such that any special triple is bettered by at least one element of $ S$?
|
To solve this problem, we need to determine for which natural numbers \( n \) there exists a set \( S \) of special triples, with \( |S| = n \), such that any special triple is bettered by at least one element of \( S \).
### Understanding the Definitions
A **special triple** \((a_1, a_2, a_3)\) is defined as a triple of nonnegative real numbers satisfying \( a_1 + a_2 + a_3 = 1 \).
A triple \((a_1, a_2, a_3)\) is **better** than a triple \((b_1, b_2, b_3)\) when exactly two of the following inequalities hold:
\[
a_1 > b_1, \quad a_2 > b_2, \quad a_3 > b_3.
\]
### Problem Analysis
To construct a set \( S \) such that any special triple is bettered by at least one element of \( S \), consider the properties of comparisons:
1. **Trivial Cases**:
- For \( n = 1 \), if \( S \) contains only one special triple, say \( (c_1, c_2, c_3) \), there exist other triples such as the same \((c_1, c_2, c_3)\) or others not bettered due to symmetry and equality conditions.
- For \( n = 2 \) or \( n = 3 \), it is not possible to construct a set \( S \) where every other special triple is worse than at least one in \( S \) due to the one-dimensional constraint and the rule of exactly two strict inequalities.
2. **General Case for \( n \geq 4 \)**:
Consider the corners of an equilateral triangle within the space defined by the sum of coordinates being 1. The ideas of domination are achievable by considering triples closer to pure corners, such as:
- \( (1, 0, 0) \)
- \( (0, 1, 0) \)
- \( (0, 0, 1) \)
- Additional points like \( (0.5, 0.5, 0) \)
These configurations cover enough of the special triples such that for any arbitrary special triple \((a_1, a_2, a_3)\) there is at least one element in \( S \) that will better it.
### Conclusion
Based on this construction, it's concluded that for \( n \geq 4 \), a collection \( S \) can be created such that any special triple is bettered by at least one element in \( S \).
Thus, the set of natural numbers \( n \) for which such a collection \( S \) can exist is:
\[
\boxed{n \geq 4}
\]
|
n\geq4
|
imc
|
omni_math-3990
|
[
"Mathematics -> Algebra -> Prealgebra -> Other"
] | 7
|
A \emph{base $10$ over-expansion} of a positive integer $N$ is an expression of the form \[ N = d_k 10^k + d_{k-1} 10^{k-1} + \cdots + d_0 10^0 \] with $d_k \neq 0$ and $d_i \in \{0,1,2,\dots,10\}$ for all $i$. Which positive integers have a unique base 10 over-expansion?
|
These are the integers with no $0$'s in their usual base $10$ expansion. If the usual base $10$ expansion of $N$ is $d_k 10^k + \cdots + d_0 10^0$ and one of the digits is $0$, then there exists an $i \leq k-1$ such that $d_i = 0$ and $d_{i+1} > 0$; then we can replace $d_{i+1} 10^{i+1} + (0) 10^i$ by $(d_{i+1}-1) 10^{i+1} + (10) 10^i$ to obtain a second base $10$ over-expansion. We claim conversely that if $N$ has no $0$'s in its usual base $10$ expansion, then this standard form is the unique base $10$ over-expansion for $N$. This holds by induction on the number of digits of $N$: if $1\leq N\leq 9$, then the result is clear. Otherwise, any base $10$ over-expansion $N = d_k 10^k + \cdots + d_1 10 + d_0 10^0$ must have $d_0 \equiv N \pmod{10}$, which uniquely determines $d_0$ since $N$ is not a multiple of $10$; then $(N-d_0)/10$ inherits the base $10$ over-expansion $d_k 10^{k-1} + \cdots + d_1 10^0$, which must be unique by the induction hypothesis.
|
Integers with no $0$'s in their base 10 expansion.
|
putnam
|
omni_math-3544
|
[
"Mathematics -> Applied Mathematics -> Probability -> Other"
] | 7.5
|
You have to organize a fair procedure to randomly select someone from $ n$ people so that every one of them would be chosen with the probability $ \frac{1}{n}$. You are allowed to choose two real numbers $ 0<p_1<1$ and $ 0<p_2<1$ and order two coins which satisfy the following requirement: the probability of tossing "heads" on the first coin $ p_1$ and the probability of tossing "heads" on the second coin is $ p_2$. Before starting the procedure, you are supposed to announce an upper bound on the total number of times that the two coins are going to be flipped altogether. Describe a procedure that achieves this goal under the given conditions.
|
To solve this problem, we must design a procedure that ensures each of the \( n \) people is selected with probability \( \frac{1}{n} \). We are given the flexibility to choose two real numbers \( 0 < p_1 < 1 \) and \( 0 < p_2 < 1 \), which are the probabilities of obtaining "heads" on the first and second coin, respectively. Additionally, we must establish an upper bound for the number of coin tosses required.
The procedure can be broken down into the following steps:
1. **Define an Appropriate Sequence of Coin Tosses:**
Since we want to utilize probabilities \( p_1 \) and \( p_2 \), our aim is to construct a sequence which will allow a fair selection with a stopping condition. A standard method is to use stochastic processes or finite-state automata to map sequences of "heads" and "tails" to each person.
2. **Selection Probabilities Using Two Coins:**
We can construct binary sequences using \( p_1 \) and \( p_2 \) that represent distinct decimal values. These values can be uniformly mapped to the probability interval \([0, 1)\). Using these probabilities, derive a method to ensure each segment \(\left[\frac{k}{n}, \frac{k+1}{n}\right)\) for \( k = 0, 1, \ldots, n-1\) corresponds to selecting each individual.
3. **Adapting the Method with Repeated Flips:**
Construct a method using repeated coin flips to mimic a random draw from a uniform distribution over \( n \) items, ensuring each person has an equal chance.
4. **Upper Bound on Number of Coin Tosses:**
The method's efficiency will require determining suitable \( m \) such that in the worst-case scenario, a decision is reached within \( m \) coin flips. This might involve setting \( m \) high enough while optimizing \( p_1 \) and \( p_2 \) to minimize the number of flips needed.
5. **Implementation and Termination:**
For example:
\begin{itemize}
\item Toss the first coin with probability \( p_1 \) repeatedly until a head appears.
\item Count the number of tosses needed to get this result.
\item Repeat with the second coin probability \( p_2 \).
\item Use the combination of these two counts to select one of \( n \) individuals.
\end{itemize}
6. **Probability Fairness Assurance:**
The methodology has to ensure that by fine-tuning \( p_1 \) and \( p_2 \) and using suitable encoding of the sequence lengths, the probability of selecting any specific individual from \( n \) approaches exactly \( \frac{1}{n} \).
Thus, by cautiously setting \( p_1 \), \( p_2 \), and the total process, we can achieve fair selection. Therefore, it is always possible to choose adequate \( p_1 \), \( p_2 \), and the sequence length \( m \) to guarantee each participant has the desired equal probability of selection.
The final conclusion is:
\[
\boxed{\text{It is always possible to choose an adequate } p \text{ and } m \text{ to achieve a fair selection.}}
\]
|
\text{It is always possible to choose an adequate } p \text{ and } m \text{ to achieve a fair selection.}
|
hungaryisrael_binational
|
omni_math-4414
|
[
"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
|
omni_math-250
|
[
"Mathematics -> Number Theory -> Factorization"
] | 7.5
|
Let $n\ge 2$ be a given integer. Find the greatest value of $N$, for which the following is true: there are infinitely many ways to find $N$ consecutive integers such that none of them has a divisor greater than $1$ that is a perfect $n^{\mathrm{th}}$ power.
|
Let \( n \geq 2 \) be a given integer. We are tasked with finding the greatest value of \( N \) such that there are infinitely many ways to select \( N \) consecutive integers where none of them has a divisor greater than 1 that is a perfect \( n^{\text{th}} \) power.
To solve this, consider the properties of divisors and the structure of \( n^{\text{th}} \) powers:
1. **Understanding \( n^{\text{th}} \) Powers:**
A perfect \( n^{\text{th}} \) power is any number of the form \( k^n \), where \( k \) is an integer. Our aim is to ensure that none of the integers in the sequence has such a divisor greater than 1.
2. **Generating a Sequence:**
The task is to find the largest \( N \) such that there exists a sequence of \( N \) consecutive integers \( a, a+1, \ldots, a+N-1 \) which satisfy the condition.
3. **Evaluating Small Values:**
Begin by checking small values of \( N \):
- If \( N = 1 \), choosing any integer \( a \) works since 1 has no non-unit divisors.
- If \( N = 2 \), choose integers such that neither of them is divisible by any prime raised to the power \( n \).
4. **Extending to Larger \( N \):**
To extend this logic, note that each prime number appears raised to a power at least \( n \) in a sequence of consecutive integers spanning a distance of more than \( n \).
5. **Conclusion Using the Chinese Remainder Theorem:**
The construction must ensure that each \( a+k \) where \( 0 \leq k < N \), avoids having a divisor that is a perfect \( n^{\text{th}} \) power. By the Chinese Remainder Theorem, such arrangements are possible, provided \( N \leq 2^n - 1 \) because beyond this, some number must necessarily be divisible by an \( n^{\text{th}} \) power.
Therefore, the greatest value of \( N \) for which such a sequence exists, based on extending the logic above, is:
\[
N = 2^n - 1
\]
Hence, the greatest value of \( N \) is:
\[
\boxed{2^n - 1}
\]
|
2^n - 1
|
problems_from_the_kmal_magazine
|
omni_math-3950
|
[
"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
|
omni_math-245
|
[
"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
|
omni_math-260
|
[
"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
|
omni_math-283
|
[
"Mathematics -> Geometry -> Plane Geometry -> Polygons"
] | 7
|
Let $m$ be a positive integer. A triangulation of a polygon is [i]$m$-balanced[/i] if its triangles can be colored with $m$ colors in such a way that the sum of the areas of all triangles of the same color is the same for each of the $m$ colors. Find all positive integers $n$ for which there exists an $m$-balanced triangulation of a regular $n$-gon.
|
To solve this problem, we need to understand the conditions under which an \(m\)-balanced triangulation of a regular \(n\)-gon is possible. The concept of \(m\)-balanced means that each color covers exactly the same total area across all triangles of that color. Here's a breakdown of the solution:
Consider a regular \(n\)-gon, and let's triangulate it. The total area of the \(n\)-gon is equally divided among the triangles formed. When coloring these triangles with \(m\) colors in an \(m\)-balanced manner, each color must cover \(\frac{1}{m}\) of the total polygon area.
**Key Conditions:**
1. Each triangle in the triangulation has an equal area because the \(n\)-gon is regular.
2. The total number of triangles that can be formed in a regular \(n\)-gon triangulation is \(n - 2\).
3. For the triangulation to be \(m\)-balanced, the total number of triangles \(n - 2\) has to be divisible by \(m\), i.e., \(m \mid (n-2)\).
Additionally, because a regular polygon of \(n\) sides can only be triangulated if \(n \geq 3\), we have \(n \geq 3\). Furthermore, to be able to partition these triangles into \(m\) groups of equal area, clearly \(n\) needs to be larger than \(m\), leading to the second condition \(n > m\).
**Final Condition:**
Combining these conditions, we conclude:
- \( m \mid (n-2) \)
- \( n > m \)
- \( n \geq 3 \)
Thus, the set of all positive integers \(n\) for which there exists an \(m\)-balanced triangulation of a regular \(n\)-gon is characterized by:
\[
\boxed{m \mid n \text{ with } n > m \text{ and } n \geq 3}
\]
This conclusion provides a comprehensive characterization of all such \(n\) where an \(m\)-balanced triangulation exists for a given \(m\)-gon.
|
m \mid n \text{ with } n > m \text{ and } n \geq 3.
|
usamo
|
omni_math-3913
|
[
"Mathematics -> Algebra -> Intermediate Algebra -> Other",
"Mathematics -> Number Theory -> Greatest Common Divisors (GCD)"
] | 7
|
Two rational numbers \(\tfrac{m}{n}\) and \(\tfrac{n}{m}\) are written on a blackboard, where \(m\) and \(n\) are relatively prime positive integers. At any point, Evan may pick two of the numbers \(x\) and \(y\) written on the board and write either their arithmetic mean \(\tfrac{x+y}{2}\) or their harmonic mean \(\tfrac{2xy}{x+y}\) on the board as well. Find all pairs \((m,n)\) such that Evan can write 1 on the board in finitely many steps.
|
Given two rational numbers \(\tfrac{m}{n}\) and \(\tfrac{n}{m}\) on a blackboard, where \(m\) and \(n\) are relatively prime positive integers, we want to determine all pairs \((m,n)\) such that it is possible for Evan to write 1 on the board after finitely many steps using the following operations:
- Write the arithmetic mean \(\tfrac{x+y}{2}\) of any two numbers \(x\) and \(y\) on the board.
- Write the harmonic mean \(\tfrac{2xy}{x+y}\) of any two numbers \(x\) and \(y\) on the board.
### Analysis
To solve this problem, we utilize the ideas from number theory and properties of rational numbers. We essentially want Evan to be able to reach the number 1, which is equivalent to having:
\[
\frac{m}{n} = 1 \quad \text{or} \quad \frac{n}{m} = 1
\]
Using the arithmetic and harmonic means, a key observation is that both operations preserve the sum and product of the two numbers. Thus, to obtain 1 on the board, it is essential that we could eventually reach \((m,n)=(1,1)\) after several operations.
### Approach
1. **Initial Setup:** Begin with two rational numbers:
\[
\frac{m}{n} \quad \text{and} \quad \frac{n}{m}
\]
2. **Arithmetic and Harmonic Means:** For any rational numbers \( \frac{a}{b} \) and \(\frac{b}{a}\), the arithmetic mean is:
\[
\frac{\frac{a}{b} + \frac{b}{a}}{2} = \frac{a^2 + b^2}{2ab}
\]
The harmonic mean is:
\[
\frac{2 \cdot \frac{a}{b} \cdot \frac{b}{a}}{\frac{a}{b} + \frac{b}{a}} = \frac{2}{\frac{a^2 + b^2}{ab}} = \frac{2ab}{a^2 + b^2}
\]
3. **Finding \(m\) and \(n\):**
- Assume that the board should evolve toward 1 through these operations.
- Without loss of generality, we consider the transformation of a more complex expression to a simplified rational number structure.
4. **Conclusion for Pair \((m,n)\):** Through the sequence of applying these arithmetic and harmonic means, we aim to show that:
- The ability to express 1 would imply transformations involving powers of 2.
- Specifically, the pairs \((a, 2^k - a)\) such that \(a\) is odd and \(k\) is positive are favorable.
- This general form arises because via strategic applications of the means, Evan can produce numbers that transform the board entries to a scenario where the operations systematically reduce to or participate in increment chains characterized by odd numbers balancing with powers of two.
Thus, the complete set of solutions where Evan can write 1 on the board in finitely many steps is given by:
\[
\boxed{(a, 2^k - a) \text{ for odd } a \text{ and positive } k}
\]
The transformation structure of powers of two ensures the feasibility of achieving the unit value through successive means operation.
|
(a, 2^k - a) \text{ for odd } a \text{ and positive } k
|
usamo
|
omni_math-3953
|
[
"Mathematics -> Geometry -> Plane Geometry -> Angles"
] | 7
|
Let $ABC$ be a triangle with $\angle BAC = 60^{\circ}$. Let $AP$ bisect $\angle BAC$ and let $BQ$ bisect $\angle ABC$, with $P$ on $BC$ and $Q$ on $AC$. If $AB + BP = AQ + QB$, what are the angles of the triangle?
|
Given a triangle \( ABC \) with the angle \( \angle BAC = 60^\circ \), we need to determine the other angles \(\angle B\) and \(\angle C\) given that \( AP \) bisects \( \angle BAC \) and \( BQ \) bisects \( \angle ABC \), where \( P \) is on \( BC \) and \( Q \) is on \( AC \), and the condition \( AB + BP = AQ + QB \) holds.
### Step-by-step Solution
1. **Apply Angle Bisector Theorems**:
- Since \( AP \) is the angle bisector of \( \angle BAC \), by the Angle Bisector Theorem:
\[
\frac{BP}{PC} = \frac{AB}{AC}
\]
- Similarly, since \( BQ \) is the angle bisector of \( \angle ABC \), by the Angle Bisector Theorem:
\[
\frac{AQ}{QC} = \frac{AB}{BC}
\]
2. **Use the condition \( AB + BP = AQ + QB \)**:
- Rearrange this equation: \( AB + BP = AQ + QB \) implies:
\[
BP - QB = AQ - AB
\]
- Let's express everything in terms of the sides of the triangle, and use information from the angle bisectors:
- We consider substitutions where \( BP \) and \( QB \) utilize the Angle Bisector Theorem relations.
3. **Express the angle conditions**:
- Note:
\[
\angle A + \angle B + \angle C = 180^\circ
\]
- Since \( \angle A = 60^\circ \):
\[
\angle B + \angle C = 120^\circ
\]
- Since \( AB + BP = AQ + QB \) and considering symmetry and equalities from the problem statement, let:
\[
\angle B = x, \quad \angle C = 120^\circ - x
\]
4. **Set equal angles based on given angle measures**:
- Considering the problem's internal symmetrical structure and condition:
\[
\angle B = 2 \times \angle C
\]
- Solving these gives:
\[
x = 80^\circ
\]
- Therefore:
\[
\angle C = 120^\circ - x = 40^\circ
\]
5. **Conclude the results**:
- Thus, the angles of the triangle \( ABC \) are:
\[
\angle B = 80^\circ, \quad \angle C = 40^\circ
\]
Hence, the angles of triangle \( ABC \) are:
\[
\boxed{\angle B = 80^\circ, \angle C = 40^\circ}
\]
|
\angle B=80^{\circ},\angle C=40^{\circ}
|
imo
|
omni_math-4092
|
[
"Mathematics -> Algebra -> Algebra -> Equations and Inequalities",
"Mathematics -> Number Theory -> Factorization"
] | 7
|
Find all triples $(x,y,z)$ of positive integers such that $x \leq y \leq z$ and
\[x^3(y^3+z^3)=2012(xyz+2).\]
|
To find all triples \((x, y, z)\) of positive integers such that \(x \leq y \leq z\) and
\[x^3(y^3 + z^3) = 2012(xyz + 2),\]
we proceed as follows:
First, note that \(2012 \cdot 2 = 2^3 \cdot 503\). Taking the equation modulo \(x\), we see that \(x \mid 2012\). Therefore, \(x\) can be \(1, 2, 4, 503, 1006, 2012\). We will consider the feasible values of \(x\) under the constraint \(x \leq y \leq z\).
### Case 1: \(x = 1\)
The equation becomes:
\[y^3 + z^3 = 2012(yz + 2).\]
Factoring the left-hand side, we get:
\[(y + z)(y^2 - yz + z^2) = 2012(yz + 2).\]
For \(503 \mid y + z\) or \(503 \mid y^2 - yz + z^2\), we apply the lemma that \(503 \nmid y^2 - yz + z^2\) for positive integers \(y\) and \(z\) not both divisible by 503. Therefore, \(503 \mid y + z\). Let \(y + z = 503n\).
Rewriting the equation:
\[n(y^2 - yz + z^2) = 4(yz + 2).\]
If \(n \geq 4\), then \((z - y)^2 \leq 2\), implying \(z = y\) or \(z = y + 1\). Both cases lead to contradictions. Thus, \(n < 4\).
Since \(y\) and \(z\) have the same parity, \(n = 2\) is required, leading to \(y + z = 1006\). This results in:
\[(y + z)^2 = 1006^2 = 5yz + 4,\]
which is a contradiction modulo 5. Hence, no solutions exist for \(x = 1\).
### Case 2: \(x = 2\)
The equation becomes:
\[y^3 + z^3 = 503(yz + 1).\]
Letting \(y + z = 503n\), we rewrite this as:
\[n(y^2 - yz + z^2) = yz + 1.\]
If \(n \geq 3\), then \(n(z - y)^2 = 1 + (1 - n)yz < 0\), a contradiction.
For \(n = 1\), we have:
\[y^2 - yz + z^2 = yz + 1,\]
or \((z - y)^2 = 1\), leading to \(z = y + 1\). Given \(y + z = 503\), we find \(y = 251\) and \(z = 252\). Verifying, \((x, y, z) = (2, 251, 252)\) satisfies the original equation.
For \(n = 2\), we get:
\[2(y^2 - yz + z^2) = yz + 1,\]
or \(2(z - y)^2 = 1 - yz\), forcing \(y = z = 1\), which contradicts \(y + z = 1006\).
Thus, the only solution is \((x, y, z) = (2, 251, 252)\).
The answer is: \(\boxed{(2, 251, 252)}\).
|
(2, 251, 252)
|
usa_team_selection_test
|
omni_math-91
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics",
"Mathematics -> Algebra -> Prealgebra -> Integers"
] | 7
|
Determine all positive integers $n$, $n\ge2$, such that the following statement is true:
If $(a_1,a_2,...,a_n)$ is a sequence of positive integers with $a_1+a_2+\cdots+a_n=2n-1$, then there is block of (at least two) consecutive terms in the sequence with their (arithmetic) mean being an integer.
|
To determine all positive integers \( n \), \( n \ge 2 \), such that the following statement is true:
If \((a_1, a_2, \ldots, a_n)\) is a sequence of positive integers with \( a_1 + a_2 + \cdots + a_n = 2n - 1 \), then there is a block of (at least two) consecutive terms in the sequence with their (arithmetic) mean being an integer.
We start by examining small values of \( n \):
- For \( n = 2 \), consider the sequence \((1, 2)\). The sum is \(1 + 2 = 3 = 2 \cdot 2 - 1\). The arithmetic mean of the block \((1, 2)\) is \(\frac{1 + 2}{2} = 1.5\), which is not an integer. However, if we consider the sequence \((2, 1)\), the arithmetic mean of \((2, 1)\) is \(\frac{2 + 1}{2} = 1.5\), which is not an integer either. Therefore, \( n = 2 \) satisfies the condition.
- For \( n = 3 \), consider the sequence \((2, 1, 2)\). The sum is \(2 + 1 + 2 = 5 = 2 \cdot 3 - 1\). The arithmetic mean of the block \((2, 1)\) is \(\frac{2 + 1}{2} = 1.5\), and the arithmetic mean of the block \((1, 2)\) is \(\frac{1 + 2}{2} = 1.5\), neither of which are integers. Therefore, \( n = 3 \) satisfies the condition.
Next, we use induction and casework to show that for \( n \ge 4 \), there will always be a block of consecutive terms whose arithmetic mean is an integer.
### Case 1: \( n = 4k \)
If \( n = 4k \), then we have \( 2k \) odd and \( 2k \) even integers. Their sum is even, which contradicts the requirement that the sum is \( 2n - 1 \), an odd number.
### Case 2: \( n = 4k + 1 \)
If \( n = 4k + 1 \), the sum \( S(1, n) \) is odd and \( 6k + 3 \le S(1, n) \le 8k + 1 \). Using strong induction and the properties of sums of sequences, we can show that there will always be a block of consecutive terms whose arithmetic mean is an integer.
### Case 3: \( n = 4k + 2 \)
If \( n = 4k + 2 \), the sum \( S(1, n) \) is odd and \( 6k + 5 \le S(1, n) \le 8k + 3 \). Similar to Case 2, using strong induction and the properties of sums of sequences, we can show that there will always be a block of consecutive terms whose arithmetic mean is an integer.
### Case 4: \( n = 4k + 3 \)
If \( n = 4k + 3 \), the sum \( S(1, n) \) is odd and \( 6k + 7 \le S(1, n) \le 8k + 5 \). Again, using strong induction and the properties of sums of sequences, we can show that there will always be a block of consecutive terms whose arithmetic mean is an integer.
Therefore, the only positive integers \( n \) that satisfy the given condition are \( n = 2 \) and \( n = 3 \).
The answer is: \boxed{2, 3}.
|
2, 3
|
usa_team_selection_test
|
omni_math-30
|
[
"Mathematics -> Discrete Mathematics -> Combinatorics"
] | 7
|
In a party with $1982$ people, among any group of four there is at least one person who knows each of the other three. What is the minimum number of people in the party who know everyone else?
|
We induct on $n$ to prove that in a party with $n$ people, there must be at least $(n-3)$ people who know everyone else. (Clearly this is achievable by having everyone know everyone else except three people $A, B, C$ , who do not know each other.)
Base case: $n = 4$ is obvious.
Inductive step: Suppose in a party with $k$ people (with $k \ge 4$ ), at least $(k-3)$ people know everyone else. Consider a party with $(k+1)$ people. Take $k$ of the people (leaving another person, $A$ , out) and apply the inductive step to conclude that at least $(k-3)$ people know everyone else in the $k$ -person group, $G$ .
Now suppose that everyone in the group $G$ knows each other. Then take $3$ of these people and $A$ to deduce that $A$ knows a person $B \in G$ , which means $B$ knows everyone else. Then apply the inductive step on the remaining $k$ people (excluding $B$ ) to find $(k-3)$ people out of them that know everyone else (including $B$ , of course). Then these $(k-3)$ people and $B$ , which enumerate $(k-2)$ people, know everyone else.
Suppose that there exist two people $B, C \in G$ who do not know each other. Because $k-3 \ge 1$ , there exist at least one person in $G$ , person $D$ , who knows everyone else in $G$ . Now, take $A, B, C, D$ and observe that because $B, C$ do not know each other, either $A$ or $D$ knows everyone else of $A, B, C, D$ (by the problem condition), so in particular $A$ and $D$ know each other. Then apply the inductive step on the remaining $k$ people (excluding $D$ ) to find $(k-3)$ people out of them that know everyone else (including $D$ , of course). Then these $(k-3)$ people and $D$ , which enumerate $(k-2)$ people, know everyone else.
This completes the inductive step and thus the proof of this stronger result, which easily implies that at least $1982 - 3 = \boxed{1979}$ people know everyone else.
|
\[ 1979 \]
|
usamo
|
omni_math-327
|
[
"Mathematics -> Algebra -> Abstract Algebra -> Field Theory"
] | 7
|
Determine all functions $f: \mathbb{R} \to \mathbb{R}$ such that
$$ f(x^3) + f(y)^3 + f(z)^3 = 3xyz $$
for all real numbers $x$, $y$ and $z$ with $x+y+z=0$.
|
We are tasked with finding all functions \( f: \mathbb{R} \to \mathbb{R} \) that satisfy the equation
\[
f(x^3) + f(y)^3 + f(z)^3 = 3xyz
\]
for all real numbers \( x \), \( y \), and \( z \) such that \( x + y + z = 0 \).
First, we consider substituting specific values to simplify and gain insights into the function \( f \).
### Step 1: Substituting Special Values
1. **Substitute \( y = z = 0 \) and \( x = 0 \):**
\[
x + y + z = 0 \implies x = 0
\]
The equation becomes:
\[
f(0^3) + f(0)^3 + f(0)^3 = 0
\]
Simplifying, we have:
\[
f(0) + 2f(0)^3 = 0
\]
Let \( f(0) = a \), then:
\[
a + 2a^3 = 0 \implies a(1 + 2a^2) = 0
\]
Hence, \( a = 0 \).
Therefore, \( f(0) = 0 \).
### Step 2: Testing Linear Functions
To satisfy the equation without restricting the linearity initially, assume \( f(t) = ct \) for some constant \( c \).
Substitute into the equation:
\[
f(x^3) = c(x^3), \quad f(y)^3 = (cy)^3 = c^3y^3, \quad f(z)^3 = (cz)^3 = c^3z^3
\]
Then the original equation becomes:
\[
c(x^3) + c^3y^3 + c^3z^3 = 3xyz
\]
which simplifies to:
\[
cx^3 + c^3(y^3 + z^3) = 3xyz
\]
Given that \( x + y + z = 0 \), \( z = -(x+y) \) riduces the expression for cubics to:
\[
cx^3 + c^3(y^3 + (-(x+y))^3) = 3xyz
\]
Solving for these confirms linear characteristics compatible with \( f(x) = x \):
\[
c(x^3) + c^3(y^3 + x^3 + 3xy(x+y)) = 3xyz
\]
Confirming, the simplification yields plausible results when \( c = 1 \).
Thus, the candidate solution is:
\[
f(x) = x
\]
### Step 3: Validating
To confirm, check \( f(x) = x \) satisfies:
Substitute back:
\[
f(x^3) = x^3, \quad f(y)^3 = y^3, \quad f(z)^3 = z^3
\]
This yields:
\[
x^3 + y^3 + z^3 = 3xyz
\]
For \( x+y+z=0 \), the identity holds, confirming:
Therefore, the only such function is
\[
\boxed{f(x) = x}
\]
|
f(x) = x
|
european_mathematical_cup
|
omni_math-3679
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.