Skip to content
🤔 DeepSeek 🧠
❓:A sledding hill consists of a straight incline ( AB ) and a horizontal section ( BC ). Point ( A ) is located 5 meters from the nearest horizontal ground point ( H ). The distance ( HC ) is 3 meters, with point ( B ) lying on the segment ( HC ). Find the distance from point ( H ) to point ( B ) such that the time it takes for the sled to move from rest along the broken line ( ABC ) is minimized. Assume a uniform gravitational field, and neglect friction, air resistance, and changes in the sled's speed vector as it passes through point ( B ). Assume the acceleration due to gravity is ( 10 text{m/s}^2 ).

🧠:Okay, so I need to find the distance from point H to point B that minimizes the time taken for a sled to go from A to C via B. The setup is a hill with an incline AB and a horizontal section BC. Point A is 5 meters above H, and HC is 3 meters. Points B are somewhere along HC, so HB is variable, and we need to find the optimal HB to minimize the time.First, let me visualize this. There's a vertical line from A down to H, so AH is 5 meters. Then, HC is 3 meters horizontally. The sled starts at A, goes down the incline AB, then moves along the horizontal BC to C. The problem says to neglect friction, air resistance, and changes in speed at B. So the sled's speed will increase as it goes down AB due to gravity, and then it will maintain that speed along BC. The key is to model the time taken for each segment (AB and BC) as functions of the position of B, then find the value of HB (let's call it x) that minimizes the total time.Let me define variables:Let x be the distance from H to B. Since HC is 3 meters, then BC will be 3 - x meters. Wait, but if H is the point directly below A, and HC is 3 meters, then H to C is 3 meters horizontally. So point B is somewhere between H and C. Therefore, HB = x, so BC = 3 - x. That makes sense.Now, the sled starts at rest at A. It slides down AB, which is an incline. The length of AB can be calculated using Pythagoras: AH is 5 meters vertically, HB is x meters horizontally. So AB = sqrt(5^2 + x^2) = sqrt(25 + x^2).Since there's no friction, the acceleration down the incline will be g*sin(theta), where theta is the angle between AB and the vertical. Wait, no, actually, the acceleration along the incline is g*sin(theta), where theta is the angle between the incline and the horizontal. Wait, maybe I need to clarify.Alternatively, let's recall that for an inclined plane, the acceleration a = g*sin(theta), where theta is the angle of the incline relative to the horizontal. Since AH is vertical 5 meters, and HB is horizontal x meters, the incline AB makes an angle theta with the horizontal, where tan(theta) = 5/x. Therefore, sin(theta) = 5 / sqrt(25 + x^2). So acceleration a = g * (5 / sqrt(25 + x^2)).But wait, the acceleration due to gravity along the incline is g*sin(theta), yes. So that's 10 * (5 / sqrt(25 + x^2)) = 50 / sqrt(25 + x^2) m/s². Wait, but acceleration is a constant here? Wait, but if the sled is moving along AB with acceleration a = g*sin(theta), which depends on theta, which depends on x. But x is the variable here. So for each x, we have a different acceleration. Hmm.But actually, since we are considering HB = x, then the length of AB is sqrt(25 + x²), as above. The sled starts from rest at A, so the time taken to go down AB can be found using kinematic equations. Since acceleration is constant, we can use s = (1/2) a t². Solving for t, t = sqrt(2s/a). So time along AB, t1 = sqrt(2 * AB / a). But AB is sqrt(25 + x²), and a = 50 / sqrt(25 + x²). Therefore, substituting:t1 = sqrt(2 * sqrt(25 + x²) / (50 / sqrt(25 + x²))) ) = sqrt(2 * (25 + x²) / 50) ) = sqrt( (25 + x²)/25 ) = sqrt(1 + (x²)/25) = sqrt( (x² + 25)/25 ) = (sqrt(x² +25))/5.Wait, let's verify that step by step. Let's compute 2 * AB / a:AB = sqrt(25 + x²)a = 50 / sqrt(25 + x²)So 2 * AB / a = 2 * sqrt(25 + x²) / (50 / sqrt(25 + x²)) ) = 2 * (sqrt(25 + x²) * sqrt(25 + x²)) / 50 = 2*(25 + x²)/50 = (50 + 2x²)/50 = (25 + x²)/25. Then take the square root of that, so sqrt( (25 + x²)/25 ) = sqrt(25 + x²)/5. So yes, t1 = sqrt(25 + x²)/5.Okay, so the time to go from A to B is t1 = sqrt(25 + x²)/5 seconds.Now, after reaching B, the sled moves along BC, which is horizontal. Since there's no friction, the sled's speed remains constant. The speed at point B is the final speed from sliding down AB. Let's compute that.Starting from rest, final velocity v = sqrt(2*a*s). Alternatively, since acceleration is a, and time is t1, v = a*t1. Let's check both:Using v = a*t1:a = 50 / sqrt(25 + x²)t1 = sqrt(25 + x²)/5So v = (50 / sqrt(25 + x²)) * (sqrt(25 + x²)/5) ) = 50 / 5 = 10 m/s. Wait, that's a constant? That can't be right. If the speed at B is always 10 m/s regardless of x? That seems odd. Let me check with energy conservation.Alternatively, using energy conservation: the sled starts at rest at A, which is 5 meters above the ground. So potential energy at A is mgh = m*10*5 = 50m Joules. At point B, which is x meters away from H, so the vertical drop from A to B is... Wait, point B is on HC, which is horizontal. So the vertical drop from A to B is AH minus... Wait, no. Since A is 5 meters above H, and B is x meters from H along HC. The vertical drop from A to B is still 5 meters? Wait, no. Wait, point B is on the horizontal segment HC. So the height from A to B is 5 meters, because B is on the ground level. Wait, that can't be. If H is the point directly below A on the ground, and HC is 3 meters, then H is on the ground. So point B is somewhere between H and C on the ground. Therefore, the vertical drop from A to B is 5 meters, regardless of where B is. Therefore, using energy conservation, the sled's potential energy at A is converted to kinetic energy at B, so mgh = (1/2)mv² ⇒ v = sqrt(2gh) = sqrt(2*10*5) = sqrt(100) = 10 m/s. So regardless of x, the speed at B is 10 m/s. That's why when I calculated using a*t1, I got 10 m/s. So that's correct. The speed at B is always 10 m/s, no matter the length of AB. Interesting. So the horizontal segment BC will always be traversed at 10 m/s. Therefore, the time taken to go from B to C is distance BC divided by 10 m/s. Since BC is 3 - x meters, then time t2 = (3 - x)/10 seconds.Therefore, total time T = t1 + t2 = sqrt(25 + x²)/5 + (3 - x)/10.Our goal is to find x in [0, 3] that minimizes T.So the problem reduces to minimizing the function T(x) = sqrt(25 + x²)/5 + (3 - x)/10 for 0 ≤ x ≤ 3.To find the minimum, take the derivative of T with respect to x, set it to zero, and solve for x.First, compute dT/dx:d/dx [ sqrt(25 + x²)/5 ] = (1/5) * ( (1/(2*sqrt(25 + x²))) ) * 2x ) = (1/5) * (x / sqrt(25 + x²)).And d/dx [ (3 - x)/10 ] = -1/10.Therefore, dT/dx = (x)/(5 sqrt(25 + x²)) - 1/10.Set derivative equal to zero:(x)/(5 sqrt(25 + x²)) - 1/10 = 0Multiply both sides by 10 to eliminate denominators:2x / sqrt(25 + x²) - 1 = 0So 2x / sqrt(25 + x²) = 1Multiply both sides by sqrt(25 + x²):2x = sqrt(25 + x²)Square both sides:4x² = 25 + x²Subtract x²:3x² = 25x² = 25/3x = 5/sqrt(3) ≈ 2.88675 metersBut wait, HC is 3 meters, so x must be ≤ 3. But 5/sqrt(3) is approximately 2.88675, which is less than 3, so that's okay.Therefore, the critical point is at x = 5/sqrt(3). Now we need to verify if this is indeed a minimum.Since the function T(x) is defined on [0, 3], we can check the second derivative or test the endpoints.Alternatively, compute the second derivative:First derivative: dT/dx = (x)/(5 sqrt(25 + x²)) - 1/10Second derivative:First term derivative: Let’s compute d/dx [ x/(5 sqrt(25 + x²)) ]Let’s denote f(x) = x / (5 sqrt(25 + x²))f'(x) = [1 * 5 sqrt(25 + x²) - x * (5 * (1/2)(25 + x²)^(-1/2) * 2x ) ] / (25 (25 + x²))Wait, maybe better to compute step by step.Let’s write f(x) = (1/5) * x * (25 + x²)^(-1/2)Then f'(x) = (1/5)[ (25 + x²)^(-1/2) + x * (-1/2)(25 + x²)^(-3/2) * 2x ]Simplify:= (1/5)[ (25 + x²)^(-1/2) - x² (25 + x²)^(-3/2) ]= (1/5)(25 + x²)^(-3/2) [ (25 + x²) - x² ]= (1/5)(25 + x²)^(-3/2) * 25= 5 / (25 + x²)^(3/2)So the second derivative of T(x) is 5 / (25 + x²)^(3/2), which is always positive for all x. Therefore, the function T(x) is convex on [0, 3], so the critical point x = 5/sqrt(3) is indeed a minimum.Therefore, the minimal time occurs when x = 5/sqrt(3). However, we need to present the answer in a simplified form, possibly rationalized.5/sqrt(3) = (5 sqrt(3))/3 ≈ 2.88675 meters, which is approximately 2.89 meters, but since the problem is in exact terms, we need to write it as (5√3)/3 meters.But let me check if 5/sqrt(3) is the exact answer. Yes, from the algebra above, we had x² = 25/3, so x = 5/sqrt(3), which is equivalent to (5√3)/3.So, the minimal time occurs when HB = (5√3)/3 meters.But wait, let me confirm once again that the steps make sense.1. Modeled the problem with x as HB.2. Calculated time along AB as sqrt(25 + x²)/5.3. Calculated time along BC as (3 - x)/10.4. Total time T(x) is the sum.5. Took derivative of T(x), set to zero, solved for x.6. Found critical point at x = 5/sqrt(3) ≈ 2.88675, which is within [0, 3].7. Confirmed second derivative is positive, hence minimum.Therefore, the answer should be HB = (5√3)/3 meters.But let me check if the energy approach was correct. Since the speed at B is always 10 m/s, regardless of x. That's because the vertical drop is 5 meters, so mgh = (1/2)mv² ⇒ v = sqrt(2gh) = sqrt(2*10*5) = sqrt(100) = 10 m/s. So that's correct. Therefore, even if the path AB is longer or shorter, the speed at B is always 10 m/s. That seems counterintuitive, but it's due to energy conservation. The potential energy lost is the same regardless of the path taken, since the vertical drop is the same. Hence, the speed at B is indeed 10 m/s always.Therefore, the time along BC is (3 - x)/10, which is correct.Therefore, the calculations hold.Alternatively, if the path AB had different heights, the speed would change, but in this case, since B is on the ground, the vertical drop is always 5 meters, so speed is fixed.Therefore, the result x = (5√3)/3 meters is correct.So converting 5/sqrt(3) to rationalized form: 5√3 / 3.Therefore, the answer is 5√3 / 3 meters.Final AnswerThe distance from ( H ) to ( B ) that minimizes the time is (boxed{dfrac{5sqrt{3}}{3}}) meters.

✅:To find the distance from point ( H ) to point ( B ) that minimizes the time taken for a sled to go from ( A ) to ( C ) via ( B ), we start by defining the distance ( HB ) as ( x ). The vertical drop from ( A ) to ( B ) is 5 meters, and the horizontal distance ( HC ) is 3 meters. The sled's speed at ( B ) is determined by energy conservation, resulting in a constant speed of 10 m/s regardless of ( x ).1. Time along ( AB ): - The length of ( AB ) is ( sqrt{25 + x^2} ). - The time to travel ( AB ) is ( t_1 = frac{sqrt{25 + x^2}}{5} ).2. Time along ( BC ): - The horizontal distance ( BC ) is ( 3 - x ). - The time to travel ( BC ) is ( t_2 = frac{3 - x}{10} ).3. Total Time: - The total time ( T(x) ) is given by ( T(x) = frac{sqrt{25 + x^2}}{5} + frac{3 - x}{10} ).4. Minimizing the Total Time: - Compute the derivative of ( T(x) ): [ frac{dT}{dx} = frac{x}{5sqrt{25 + x^2}} - frac{1}{10} ] - Set the derivative to zero and solve for ( x ): [ frac{x}{5sqrt{25 + x^2}} - frac{1}{10} = 0 implies frac{2x}{sqrt{25 + x^2}} = 1 implies 4x^2 = 25 + x^2 implies 3x^2 = 25 implies x = frac{5sqrt{3}}{3} ]5. Verification: - The second derivative of ( T(x) ) is positive, confirming a minimum.Thus, the distance from ( H ) to ( B ) that minimizes the time is (boxed{dfrac{5sqrt{3}}{3}}) meters.

🔑:1. Define the given parameters: - H: Vertical distance from point A to ground, H = 5 meters. - S: Horizontal distance from point H to point C, S = 3 meters. - Let x be the distance from point H to point B that we need to find.2. Calculate the time it takes the sled to fall from point A to point H: [ t_{AH} = sqrt{frac{2H}{g}} ] Substituting H = 5 text{ m} and g = 10 text{ m/s}^2: [ t_{AH} = sqrt{frac{2 times 5}{10}} = sqrt{1} = 1 text{ s} ]3. Calculate the time it takes the sled to go from point A to point B: [ t_{AB} = frac{t_{AH}}{sin angle ABH} ]4. To find sin angle ABH, note that sin angle ABH = frac{H}{sqrt{H^2 + x^2}}.5. Therefore, [ t_{AB} = frac{1}{frac{H}{sqrt{H^2 + x^2}}} = frac{sqrt{H^2 + x^2}}{H} ] Substituting H = 5: [ t_{AB} = frac{sqrt{5^2 + x^2}}{5} = frac{sqrt{25 + x^2}}{5} ]6. The velocity at point B: [ V = sqrt{2 g H} = sqrt{2 times 10 times 5} = sqrt{100} = 10 text{ m/s} ]7. Calculate the time to travel horizontally from B to C: [ t_{BC} = frac{S - x}{V} = frac{3 - x}{10} ]8. Total time to travel from A to C: [ t_{ABC} = t_{AB} + t_{BC} = frac{sqrt{25 + x^2}}{5} + frac{3 - x}{10} ]9. Minimize the function: [ f(x) = frac{2sqrt{H^2 + x^2}}{10 sqrt{2gH}} + frac{S - x}{V} ] Simplify by considering only the significant components: [ f(x) = 2 sqrt{H^2 + x^2} - x ]10. Take the derivative and set it to zero to find critical points: [ f'(x) = frac{2x}{sqrt{H^2 + x^2}} - 1 ] Solving f'(x) = 0, [ frac{2x}{sqrt{H^2 + x^2}} = 1 Rightarrow 2x = sqrt{H^2 + x^2} Rightarrow 4x^2 = H^2 + x^2 Rightarrow 3x^2 = H^2 Rightarrow x = frac{H}{sqrt{3}} ]11. Substituting H = 5 into the solution: [ x = frac{5}{sqrt{3}} = frac{5sqrt{3}}{3} ]Conclusion:[boxed{frac{5 sqrt{3}}{3}}]

❓:In a permutation ({a_{1}, a_{2}, cdots, a_{n}}) of the set ({1, 2, cdots, n}), if (a_{i} neq i) ((i=1, 2, cdots, n)), then this permutation is called a derangement. Find the number of derangements (D_n).

🧠:Okay, so I need to find the number of derangements ( D_n ) for a set of size ( n ). A derangement is a permutation where no element appears in its original position. Let me think through this step by step.First, I know that the total number of permutations of ( n ) elements is ( n! ). But we need to exclude all the permutations where at least one element is in its original position. Hmm, this sounds like a problem that involves the principle of inclusion-exclusion. Let me recall how that works.The inclusion-exclusion principle is used to calculate the size of the union of multiple sets by alternately adding and subtracting the sizes of all possible intersections. In this case, each set ( A_i ) can represent the permutations where the ( i )-th element is fixed in its original position. So, the number of derangements would be the total permutations minus the permutations where at least one element is fixed.Mathematically, the formula using inclusion-exclusion for the union of sets ( A_1, A_2, ldots, A_n ) is:[left| bigcup_{i=1}^n A_i right| = sum_{i=1}^n |A_i| - sum_{1 leq i < j leq n} |A_i cap A_j| + sum_{1 leq i < j < k leq n} |A_i cap A_j cap A_k| - cdots + (-1)^{n+1} |A_1 cap cdots cap A_n|]Since we want the number of permutations that are not in any ( A_i ), that is, derangements, we subtract this union from the total permutations:[D_n = n! - left| bigcup_{i=1}^n A_i right|]Substituting the inclusion-exclusion formula into this:[D_n = n! - left( sum_{i=1}^n |A_i| - sum_{i<j} |A_i cap A_j| + cdots + (-1)^{n+1} |A_1 cap cdots cap A_n| right)]Which simplifies to:[D_n = n! - sum_{i=1}^n |A_i| + sum_{i<j} |A_i cap A_j| - cdots + (-1)^n |A_1 cap cdots cap A_n|]Now, let's compute each term. The first term ( |A_i| ) is the number of permutations where the ( i )-th element is fixed. If one element is fixed, the remaining ( n - 1 ) elements can be permuted freely, so ( |A_i| = (n - 1)! ). There are ( binom{n}{1} ) such terms.Similarly, ( |A_i cap A_j| ) is the number of permutations where both the ( i )-th and ( j )-th elements are fixed. This leaves ( n - 2 ) elements to permute, so ( |A_i cap A_j| = (n - 2)! ). There are ( binom{n}{2} ) such terms.In general, for the intersection of ( k ) sets, the number of permutations fixing ( k ) specific elements is ( (n - k)! ), and there are ( binom{n}{k} ) such intersections.Putting this all together:[D_n = n! - binom{n}{1}(n - 1)! + binom{n}{2}(n - 2)! - cdots + (-1)^n binom{n}{n}(n - n)!]Simplifying each term:- The first term after ( n! ) is ( -n times (n - 1)! = -n! )- The next term is ( +frac{n(n - 1)}{2} times (n - 2)! = +frac{n!}{2} )- Then ( -frac{n(n - 1)(n - 2)}{6} times (n - 3)! = -frac{n!}{6} )- And so on, until the last term ( (-1)^n times 1 times 0! = (-1)^n times 1 )So each term is ( (-1)^k times frac{n!}{k!} ), since:[binom{n}{k}(n - k)! = frac{n!}{k!(n - k)!} times (n - k)! = frac{n!}{k!}]Therefore, substituting back:[D_n = n! left( 1 - frac{1}{1!} + frac{1}{2!} - frac{1}{3!} + cdots + (-1)^n frac{1}{n!} right )]Which can be written compactly as:[D_n = n! sum_{k=0}^n frac{(-1)^k}{k!}]Wait, let me verify that. When ( k = 0 ), the term is ( frac{(-1)^0}{0!} = 1 ), and then for ( k = 1 ) to ( n ), we have the rest of the terms. So yes, the formula simplifies to the sum from ( k = 0 ) to ( n ) of ( (-1)^k / k! ), multiplied by ( n! ).Alternatively, this can also be expressed using the nearest integer function, as ( D_n = left[ frac{n!}{e} right] ), where ( [ cdot ] ) denotes rounding to the nearest integer. Because the sum ( sum_{k=0}^infty frac{(-1)^k}{k!} = frac{1}{e} ), so truncating the series at ( k = n ) gives an approximation that is very close to ( n!/e ), and the difference is less than ( 1/(n+1) ), so rounding gives the exact integer.But since the problem asks for an exact formula, the inclusion-exclusion expression is probably the answer expected. Let me check for small values of ( n ) to ensure this formula works.For ( n = 1 ): The only permutation is [1], which is not a derangement. So ( D_1 = 0 ). Using the formula:( 1! (1 - 1/1!) = 1(1 - 1) = 0 ). Correct.For ( n = 2 ): Permutations are [1,2] and [2,1]. The derangement is [2,1], so ( D_2 = 1 ). Formula:( 2! (1 - 1 + 1/2!) = 2(0 + 0.5) = 1 ). Correct.For ( n = 3 ): The derangements are [2,3,1] and [3,1,2], so ( D_3 = 2 ). Formula:( 3! (1 - 1 + 1/2 - 1/6) = 6(0 + 0.5 - 0.1667) ≈ 6(0.3333) ≈ 2 ). Correct.For ( n = 4 ): Derangements are 9. Let's compute via formula:( 4! (1 - 1 + 1/2 - 1/6 + 1/24) = 24(0 + 0.5 - 0.1667 + 0.0417) = 24(0.375) = 9 ). Correct.So the formula works for these cases. Therefore, the number of derangements ( D_n ) is indeed:[D_n = n! left( 1 - frac{1}{1!} + frac{1}{2!} - frac{1}{3!} + cdots + (-1)^n frac{1}{n!} right )]Alternatively, this can be written using the summation notation as:[D_n = sum_{k=0}^n (-1)^k frac{n!}{k!}]Another way to express derangements is through the recurrence relation. I recall that derangements satisfy the recurrence ( D_n = (n - 1)(D_{n - 1} + D_{n - 2}) ). Let me verify this.For example, using ( D_1 = 0 ), ( D_2 = 1 ). Then ( D_3 = 2(D_2 + D_1) = 2(1 + 0) = 2 ), which matches. ( D_4 = 3(D_3 + D_2) = 3(2 + 1) = 9 ), which is correct. So yes, the recurrence holds. However, the problem asks for the number of derangements, so while the recurrence is useful for computation, the closed-form formula via inclusion-exclusion is the direct answer.Another approach to derive the formula is using probability. The probability that a random permutation is a derangement is approximately ( 1/e ), but again, that's an approximation. The exact value is given by the inclusion-exclusion formula.Alternatively, generating functions can be used. The exponential generating function for derangements is ( frac{e^{-x}}{1 - x} ). But expanding this would lead back to the same series.In summary, after analyzing through inclusion-exclusion, verifying with small cases, considering alternative approaches like recurrence and generating functions, the exact formula for the number of derangements is the inclusion-exclusion sum as derived.Final AnswerThe number of derangements ( D_n ) is boxed{n! left( 1 - frac{1}{1!} + frac{1}{2!} - frac{1}{3!} + cdots + (-1)^n frac{1}{n!} right)}.

✅:To find the number of derangements ( D_n ) for a set of size ( n ), we use the principle of inclusion-exclusion. A derangement is a permutation where no element appears in its original position. The total number of permutations of ( n ) elements is ( n! ). We need to exclude the permutations where at least one element is in its original position. Using the principle of inclusion-exclusion, we calculate the number of such permutations and subtract them from the total.The formula derived using inclusion-exclusion is:[D_n = n! - left( sum_{i=1}^n |A_i| - sum_{i<j} |A_i cap A_j| + cdots + (-1)^{n+1} |A_1 cap cdots cap A_n| right)]Each term ( |A_i| ) is the number of permutations fixing the ( i )-th element, which is ( (n - 1)! ). For intersections of ( k ) sets, the number of permutations fixing ( k ) elements is ( (n - k)! ), and there are ( binom{n}{k} ) such intersections. Simplifying these terms, we get:[D_n = n! left( 1 - frac{1}{1!} + frac{1}{2!} - frac{1}{3!} + cdots + (-1)^n frac{1}{n!} right)]This can be written compactly as:[D_n = sum_{k=0}^n (-1)^k frac{n!}{k!}]Verification with small values of ( n ) confirms the formula's correctness. Thus, the number of derangements ( D_n ) is:[boxed{n! left( 1 - frac{1}{1!} + frac{1}{2!} - frac{1}{3!} + cdots + (-1)^n frac{1}{n!} right)}]

🔑:## Solution:To find the number of derangements of the set {1, 2, cdots, n}, denoted by D_n, we need a permutation {a_1, a_2, cdots, a_n} such that a_i neq i for all i = 1, 2, cdots, n.1. Base Cases: - For n = 1, there is no derangement since the only permutation is {1} and it satisfies a_1 = 1. Therefore, D_1 = 0. - For n = 2, the only permutation is {2, 1}, which is a derangement as a_1 neq 1 and a_2 neq 2. Thus, D_2 = 1.2. Recursive Formula Derivation: For n geq 3 and any derangement {a_1, a_2, cdots, a_n}: - Select a_1 = k, where k neq 1. This means a_1 can take any value from 2 to n. Hence, there are n-1 choices for a_1. We have two cases based on the position of 1 in the derangement: 1. If a_k = 1: - The derangement reduces to finding the derangements of the n-2 elements {2, 3, cdots, k-1, k+1, cdots, n} except for k. There are D_{n-2} derangements of this set. 2. If a_k neq 1: - This reduces to finding the derangements of the n-1 elements {2, 3, cdots, k-1, k+1, cdots, n} (excluding k), and there are D_{n-1} such derangements. Thus, combining both cases we get: [ D_n = (n-1)left(D_{n-1} + D_{n-2}right) quad text{for } n geq 3. ]3. Alternative Form Using Probabilities: Let E_n = frac{D_n}{n!} be the probability that a random permutation of n elements is a derangement: [ E_{n}=left(1-frac{1}{n}right) E_{n-1}+frac{1}{n} E_{n-2}. ]4. Transforming the Equation: [ E_{n} - E_{n-1} = left(-frac{1}{n}right)left(E_{n-1} - E_{n-2}right). ]5. Solving the Recurrence: Using the base cases E_1 = 0 and E_2 = frac{1}{2}, for n geq 2, we get: [ begin{aligned} E_n - E_{n-1} &= left(-frac{1}{n}right)left(E_{n-1} - E_{n-2}right) &= left(-frac{1}{n}right)left(-frac{1}{n-1}right)(E_{n-2} - E_{n-3}) &= cdots &= frac{(-1)^{n}}{n!}. end{aligned} ] Therefore: [ E_n = sum_{k=2}^{n} frac{(-1)^k}{k!}. ]6. Final Expression: Substituting back E_n into D_n = n! E_n, we obtain: [ D_n = n!left(1 - frac{1}{1!} + frac{1}{2!} - cdots + (-1)^n frac{1}{n!}right). ]Hence, the number of derangements of the set {1, 2, cdots, n} is [boxed{D_n = n!left(1 - frac{1}{1!} + frac{1}{2!} - cdots + (-1)^n frac{1}{n!}right)}.]

❓:1) Alice wants to color the integers between 2 and 8 (inclusive) using k colors. She wants that if m and n are integers between 2 and 8 such that m is a multiple of n and m neq n, then m and n are different colors. Determine the smallest integer k for which Alice can color the integers 2, 3, ldots, 8 using k colors.2) Alice wants to color the integers between 2 and 31 (inclusive) using k colors. She wants that if m and n are integers between 2 and 31 such that m is a multiple of n and m neq n, then m and n are different colors. Determine the smallest integer k for which Alice can color the integers 2, 3, ldots, 31 using k colors.

🧠:Okay, let's tackle these two problems step by step. They both involve coloring integers such that if one number is a multiple of another (and they're not the same), they must have different colors. The goal is to find the minimum number of colors needed for each range: first from 2 to 8, then from 2 to 31.Starting with problem 1: numbers 2 through 8. I need to figure out the smallest k where such a coloring is possible. Let me think about how multiples relate here. First, list out the numbers: 2, 3, 4, 5, 6, 7, 8. Now, identify which numbers are multiples of others. Let's note the multiples:- 4 is a multiple of 2 (4 = 2*2)- 6 is a multiple of 2 and 3 (6 = 2*3, 6 = 3*2)- 8 is a multiple of 2 and 4 (8 = 2*4, 8 = 4*2)- 5, 7 are primes and only multiples of themselves in this range, but their multiples (like 10, 14) are outside the range, so they don't affect others here.Similarly, 3 is prime, but 6 is a multiple of 3. So, connections are:- 2 connected to 4, 6, 8- 3 connected to 6- 4 connected to 8- 5, 7 not connected to anyone except through their multiples, but since their multiples are outside the range, they don't need different colors from others except possibly each other. Wait, but 5 and 7: since they are primes, their only multiples here are themselves, but they aren't multiples of each other. So, they can share the same color if possible.But wait, the problem states that if m is a multiple of n (and m ≠ n), then m and n must have different colors. So, the key is to model this as a graph where each number is a node, and there's an edge between n and m if one is a multiple of the other. Then, the problem reduces to finding the chromatic number of this graph.So, constructing this graph:Nodes: 2, 3, 4, 5, 6, 7, 8Edges:- 2 connected to 4, 6, 8- 3 connected to 6- 4 connected to 8- 6 is connected to 2 and 3- 8 connected to 2 and 4- 5 and 7 have no edges because their multiples are not in the set. So, 5 and 7 are isolated nodes.Therefore, the graph consists of:- A component with nodes 2,4,6,8 connected as 2-4-8 and 2-6, 3-6.- Node 3 connected to 6, which is connected to 2,4,8.- Nodes 5 and 7 are separate.So, combining all, the main component is 2,3,4,6,8 connected together, and 5 and 7 are separate.Therefore, the chromatic number is the maximum between the chromatic number of the main component and the chromatic numbers of the isolated nodes (which are 1 each). So, the key is the chromatic number of the main component.Looking at the main component: nodes 2,3,4,6,8.Edges:- 2 connected to 4,6,8- 4 connected to 2,8- 6 connected to 2,3- 8 connected to 2,4- 3 connected to 6So, let's see if this graph is a cycle or something. Let's try to color it.First, node 2 is connected to 4,6,8. So, it needs a color different from all those. Let's assign color 1 to node 2.Then, nodes 4,6,8 cannot be color 1. Let's look at node 3. Node 3 is connected to 6, so 6 and 3 need different colors. But node 6 is connected to 2 (color 1), so 6 can be color 2 or 3, depending on other connections.Wait, maybe it's better to try to model this step by step.Alternatively, perhaps this graph can be colored with 3 colors. Let's check.Suppose:- Color 2 with color 1.Then, 4,6,8 must be different from color 1.Assign color 2 to 4. Then, 8 is connected to 4 and 2, so 8 can't be color 1 or 2. Assign color 3 to 8.Now, node 6 is connected to 2 (color 1), so it can be color 2 or 3. But 6 is connected to 3 as well. So, let's look at node 3.Node 3 is connected to 6, so if we assign color to node 3 first, maybe it's better.Assign color 2 to node 3. Then, node 6 must be different from 3 and 2 (since connected to 2). So, node 6 can't be color 1 or 2. Assign color 3 to node 6.But then node 6 is color 3. Now, node 8 is color 3. But 6 and 8 are not connected. Wait, 6 and 8 are both color 3. Are they connected? 8 is a multiple of 2 and 4. 6 is a multiple of 2 and 3. 6 and 8 are not multiples of each other. So, 6 and 8 are not connected. Therefore, they can share the same color. So, that's okay.So, so far:- 2:1- 3:2- 4:2- 6:3- 8:3But check connections:- 2 (1) connected to 4 (2), 6 (3), 8 (3): okay.- 4 (2) connected to 2 (1) and 8 (3): okay.- 6 (3) connected to 2 (1) and 3 (2): okay.- 8 (3) connected to 2 (1) and 4 (2): okay.- 3 (2) connected to 6 (3): okay.So, this works. Then 5 and 7 can be colored with color 1, since they are not connected to anyone. Wait, but can they? If 5 is color 1, but is there any restriction? Since 5 is only connected to its multiples, which are 10, 15, etc., which are outside the range. So, no conflicts. Similarly for 7.But hold on, if we color 5 and 7 with color 1, which is already used by node 2. Since 5 and 7 are not connected to 2 or any other node, that's acceptable. So, total colors used would be 3.Wait, but let me verify again. If 5 is color 1, that's okay because 5 is not a multiple or a factor of any other number in the set. Similarly, 7. So yes, they can share color 1. Therefore, total colors needed are 3.But wait, let's check if 2 colors might be possible. Let's try.Suppose we try to use 2 colors. Let's assign color 1 to node 2. Then, nodes 4,6,8 must be color 2. Assign color 2 to 4,6,8. But node 6 is connected to node 3. So, node 3 can't be color 2. Assign color 1 to node 3. Now, check node 3 (color 1) connected to node 6 (color 2): okay. But node 3 is connected only to 6. Then, node 5 and 7 can be color 1 or 2. Wait, but in this case, node 3 is color 1, which is same as node 2. Since 3 and 2 are not connected, that's okay. So, 5 and 7 can be colored 1 or 2. So, can this coloring work?Wait, but node 6 is color 2, connected to node 3 (color 1) and node 2 (color 1). That's okay. Node 4 (color 2) is connected to node 2 (color 1) and node 8 (color 2). Wait, node 4 is color 2, connected to node 8 which is also color 2. That's a problem! Because 8 is a multiple of 4, so they need different colors. But in this case, both are color 2. Which violates the condition. So, that's a conflict.Ah, right. So, node 4 and node 8 are connected (since 8 is a multiple of 4). Therefore, they need different colors. So, if node 4 is color 2, node 8 must be a different color. But if we are limited to 2 colors, then node 8 must be color 1. But node 8 is connected to node 2 (color 1). So, node 8 can't be color 1. Therefore, 2 colors are insufficient.Therefore, the initial attempt with 3 colors works. So, the chromatic number is 3.Wait, but let me check again. Let's try another approach. Maybe arrange the numbers in order and see their connections.Alternatively, since this graph includes a cycle. Let's see if there's a cycle of odd length, which would necessitate 3 colors.Looking at the main component: nodes 2,4,6,8,3.Connections: 2-4-8-2 (a triangle?), no. 2 connected to 4, 4 connected to 8, 8 connected back to 2. So, that's a triangle (cycle of length 3). 2-4-8-2. So, a triangle requires 3 colors. Then, node 6 connected to 2 and 3, node 3 connected to 6. So, node 3 is connected to 6, which is connected to 2 (part of the triangle). So, the triangle is 2-4-8-2. If we color 2,4,8 with three different colors, then node 6 is connected to 2 (color 1), so it needs a different color. Let's say color 2. Then, node 3 is connected to 6 (color 2), so node 3 needs a different color, say color 3. But then node 3 is color 3, which is already used in the triangle (maybe color 3 is used for node 8). Wait, maybe:If 2:1, 4:2, 8:3 (since triangle 2-4-8-2 requires three colors). Then, node 6 connected to 2 (1) and 3. Assign color 2 to node 6 (since connected to 1). Then node 3 is connected to 6 (2), so node 3 needs color 3. But color 3 is already used by node 8, which is okay because 3 and 8 are not connected. So, yes, this works. Then, 5 and 7 can be colored 1. So, total colors 3.Alternatively, if we can use color 1 for node 3, but node 3 is connected to 6 (color 2), so node 3 can be color 1 only if it's not connected to color 1. Since node 3 is connected to 6 (color 2), and node 3 is not connected to node 2 (color 1), so yes, node 3 can be color 1. Wait, but then node 3 is color 1, which is same as node 2. But 3 and 2 are not connected, so that's okay. Wait, but node 3 is connected to node 6, which is color 2. So, node 3 can be color 1. Then, node 5 and 7 can also be color 1. So, in this case:- 2:1- 4:2- 8:3- 6:2- 3:1- 5:1- 7:1But check the edges:- 3 (1) connected to 6 (2): okay.- 6 (2) connected to 2 (1): okay.- 8 (3) connected to 4 (2) and 2 (1): okay.- 4 (2) connected to 2 (1) and 8 (3): okay.- 2 (1) connected to 4 (2), 6 (2), 8 (3): okay.So, this works. Then, colors used are 1, 2, 3. So, three colors. Therefore, k=3 is sufficient. And we saw that with 2 colors it's impossible due to the triangle 2-4-8. Therefore, the minimal k is 3.So, problem 1 answer is 3.Now, moving to problem 2: numbers from 2 to 31. Need to color them so that if m is a multiple of n (m ≠ n), they have different colors. Find minimal k.This seems more complex. The approach would be similar: model this as a graph where each number is a node, edges between n and m if one is a multiple of the other, then find the chromatic number.But since the range is larger (2-31), we need a smarter approach. The chromatic number will correspond to the size of the largest clique in the graph, because in such a graph, the clique number is equal to the chromatic number. Wait, is that true?Wait, in a graph where edges represent "one is a multiple of the other", the cliques correspond to sets of numbers where each pair has one number a multiple of the other. Such a set is called a "tower" in poset terminology. The size of the largest such tower is the maximum number of elements where each divides the next. In that case, the chromatic number is equal to the size of the largest clique, which is the size of the largest such tower. Because each element in the tower must have a different color, so the chromatic number is at least the size of the largest tower. Conversely, if you can color the numbers with as many colors as the size of the largest tower, then it's possible, by assigning colors based on the level in the tower.Wait, actually, in such a graph, the chromatic number equals the clique number. Because it's a comparability graph (it can be oriented as a DAG by directing edges from divisor to multiple), and comparability graphs are perfect graphs, meaning their chromatic number equals their clique number. So, the minimal number of colors needed is equal to the size of the largest clique, which is the size of the largest chain (totally ordered subset under divisibility). Therefore, the problem reduces to finding the maximum length of a chain of divisors in 2-31.Therefore, the answer for problem 2 is the length of the longest chain of numbers in 2-31 where each divides the next. So, we need to find the longest such chain.Let me try to find the longest chain. For example:Start with 2, 4, 8, 16, 32. But 32 is outside the range. So, 2,4,8,16: length 4.Alternatively, 3,6,12,24: length 4. Or 5,10,20,40: but 10,20,40 are in or out? 10 is in, 20 is in, 40 is out. So, 5,10,20: length 3.Similarly, 7,14,28: length 3.Wait, but let's check other possibilities. For example:Starting with 2:2, 4, 8, 16: length 4.Alternatively, 2, 6, 12, 24: 2 divides 6, 6 divides 12, 12 divides 24. But 24 is within 31. So, 2,6,12,24: length 4.Similarly, 3,6,12,24: length 4.Or 2,4,12,24: 2 divides 4, 4 divides 12, 12 divides 24. Also length 4.But is there a longer chain?Let's see. Maybe with primes:If we take primes, but primes don't divide each other, so chains of primes would be length 1. So, not helpful.Another approach: find the number with the highest number of prime factors (with multiplicity), but that might not directly correspond. Wait, the chain length is determined by exponents in prime factorization. For example, numbers like 16 = 2^4, so you can have a chain 2,4,8,16. Similarly, 24=2^3*3, but its chain might not be longer. Let's check.Wait, maybe there is a chain of length 5. Let's check.Starting with 2:2, 4, 8, 16, 32. But 32 is outside. So, 2,4,8,16: length 4.Another start: 3,6,12,24,48. 48 is out. So, 3,6,12,24: length 4.Starting with 2, multiply by 3:2,6,12,24: same as above.Starting with 2, then 4, then 8, then 16: same.Alternatively, 2,4,8,24: but 24 is divisible by 8, so 2,4,8,24: length 4.Wait, but 24 can follow 12 as well. Let's see.Wait, perhaps using different factors:For example: 2,4,12,24: 2 divides 4, 4 divides 12 (since 12=4*3), but 4 doesn't divide 12 directly? Wait, 12 divided by 4 is 3, so yes, 4 divides 12. Then 12 divides 24. So, 2,4,12,24: length 4.Alternatively, 2,6,12,24: same length.What about 3,6,12,24: same.But can we have a chain longer than 4? Let's see.Suppose we start with a prime, then multiply by different primes. For example: 2, 2*2=4, 4*3=12, 12*2=24, 24* something? 24*2=48 (out). So, no. Alternatively, 2,4,8,16, but then next would be 32.Wait, no. So, seems maximum chain length is 4.Wait, but wait. Let's check another number. What about 16? 16 is 2^4. So, chain: 2,4,8,16: length 4.What about numbers like 18: 2,6,18. Then 18*2=36 (out). So, length 3.Similarly, 27: 3,9,27. 3 divides 9, 9 divides 27: length 3.25: 5,25: length 2.So, the longest chains seem to be length 4.But wait, let me check again. Let's think if there's a chain of length 5.Suppose: 2, 4, 8, 16, 24. Wait, does 16 divide 24? 24 divided by 16 is 1.5. No. So, no. How about 2,4,8,24. Still length 4.Alternatively: 3,6,12,24, 12*2=24, but already there. Not longer.Another approach: 2, 4, 12, 24. 2 divides 4, 4 divides 12, 12 divides 24. Still length 4.Wait, but what if we do 2, 4, 8, 24. 2 divides 4, 4 divides 8, 8 divides 24? 24 divided by 8 is 3, yes. So, 2,4,8,24: that's length 4.Similarly, 2, 6, 12, 24: 2 divides 6, 6 divides 12, 12 divides 24: length 4.Is there a way to have a chain of 5?Wait, let's see. If we take 2, 4, 8, 16, but then 16 can't divide anything else in the range. What if we take 2, 4, 8, 24. Then 24 can't divide anything else. Still length 4.Alternatively, 3, 6, 12, 24: same.Alternatively, 5, 10, 20, 40: but 40 is out. So, 5,10,20: length 3.How about 7,14,28: length 3.Another idea: mixing different primes. For example, 2, 6, 12, 24, 12*5=60 (out). No. Not helpful.Alternatively, 2, 4, 12, 24. Wait, same as before.Wait, what about starting with a higher prime? Like 3, 9, 27: length 3.No. So, seems like the maximum chain length is 4.But let's check if there's any number in 2-31 that can extend a chain beyond 4.Wait, 24 is divisible by 2,3,4,6,8,12. But can we have a chain like 2, 3, 6, 12, 24. Wait, but 2 and 3 are not connected (neither divides the other). So, that's not a chain. Similarly, 2,4,6,12,24. 2 divides 4, but 4 doesn't divide 6. So, no.Alternatively, 2,4,8,16. 16 is 16. Then 16 doesn't divide any other number in the range except 16 itself. So, cannot extend.Alternatively, 2,4,8,24: as before. 24 can be divided by 8, but then 24 can't divide anything else in 2-31. So, stuck.Wait, 24 can divide 48, which is out. So, no. Therefore, chain length remains 4.But wait, let's check another possibility: 2,4,8,16 (length 4), but also 2,4,8,24 (length 4). So, maximum chain length is 4.But wait, what about numbers like 30. 30 is divisible by 2,3,5,6,10,15. But building a chain: for example, 2,6,30. 2 divides 6, 6 divides 30. So, length 3. Similarly, 3,15,30: length 3.Alternatively, 5,10,30: length 3.So, nothing longer than 4.But wait, here's a thought. Consider the number 24. Let's see:Start with 2, 4, 12, 24. 2 divides 4, 4 divides 12 (since 12/4=3), 12 divides 24. So, that's a chain of 4.Alternatively, 3, 6, 12, 24: same.Alternatively, 2, 6, 12, 24: same.Alternatively, 2, 4, 8, 24: 2 divides 4, 4 divides 8, 8 divides 24. That's also 4.So, all these chains are length 4.Is there any chain of length 5?Wait, let's see: 2, 4, 8, 16, and then 16 can't go further. What if we take 2, 4, 8, 24, but 24 is not divisible by 16. So, can't connect. No.Another idea: 2, 4, 12, 24. That's 4. Or 2, 4, 12, 24. No difference.Wait, maybe another number? Like 18. 2, 6, 18. 18 is divisible by 2 and 6. But 18 doesn't divide anything except 36 (out). So, length 3.Alternatively, 3, 9, 27. 27 is in the range. So, 3 divides 9, 9 divides 27. Length 3.Similarly, 5,25: length 2.Another angle: composite numbers. For example, 16: chain 2,4,8,16: length 4.Similarly, 27: 3,9,27: length 3.So, no, all chains seem to max out at 4.Wait, but wait a second. Let me check 2, 4, 8, 16: that's four numbers. Then, is there a chain where we can have five numbers? Let's see.Suppose we take 2, 3, 6, 12, 24. But in this chain, each number must divide the next. 2 divides 3? No. So, invalid.Alternatively, 2, 2, 2, 2, 2: but they are the same number, which is not allowed.Alternatively, 2, 4, 8, 16, and then another number. But 16 can't divide anything else. How about 2, 4, 8, 24: but 8 divides 24, but 24 can't divide anything else. So, still four.Alternatively, 2, 4, 12, 24: four.Wait, another idea: using different prime factors but still maintaining divisibility. For example:Start with 2, then 2*2=4, then 4*3=12, then 12*2=24, then 24* something. But 24*2=48 (out). So, no.Alternatively, 2,4,12,24: that's four. If we start with 3: 3,6,12,24: four.Same.Is there a number in 2-31 that can be part of a longer chain?Wait, let's think recursively. For each number, what's the maximum chain length ending with that number.For example:- For primes (2,3,5,7,11,13,17,19,23,29,31), the chain length ending at them is 1, since they can't be divided by any smaller number in the set (except 1, but 1 is not included).- For composite numbers:Take 4: chain length is 2 (2,4).6: chain length is 2 (2,6) or (3,6), so max 2.But wait, if we consider chains where each step divides the next, then:The chain length for 4 is 2 (2,4).For 6: the longest chain ending at 6 would be max(chain length of 2, chain length of 3) +1. Since 6 is divisible by 2 and 3. The chain length of 2 is 1, chain length of 3 is 1. So, chain length for 6 is 2.Similarly, 8: divisible by 2 and 4. Chain length of 2 is 1, chain length of 4 is 2. So, chain length for 8 is 3 (2,4,8).9: divisible by 3. Chain length of 3 is 1, so chain length for 9 is 2.10: divisible by 2,5. Chain lengths of 2 and 5 are 1 each. So, chain length for 10 is 2.12: divisible by 2,3,4,6. The maximum chain length among these is chain length of 4 (which is 3), chain length of 6 (which is 2), etc. So, chain length for 12 is 3 +1 =4.Wait, wait. Let me formalize this. Let's define for each number n, f(n) = the length of the longest chain ending at n.Then:f(2) = 1f(3) = 1f(4) = f(2) +1 = 2f(5) =1f(6)= max(f(2), f(3)) +1 = max(1,1)+1=2f(7)=1f(8)= f(4)+1=3f(9)= f(3)+1=2f(10)=max(f(2),f(5)) +1=2f(11)=1f(12)= max(f(2),f(3),f(4),f(6)) +1= max(1,1,2,2)+1=3Wait, but earlier I thought 12 could be in a chain of length 4. But according to this, f(12)=3. Contradiction.Wait, let's check.If f(12) is defined as the maximum f(k) among its divisors plus 1.For 12, divisors in the set are 2,3,4,6.f(2)=1, f(3)=1, f(4)=2, f(6)=2.So, max is 2, so f(12)=3. But earlier, we had a chain like 2,4,12. Which is length 3. Or 2,6,12. Length 3. So, yes, f(12)=3.But earlier, I thought there was a chain like 2,4,12,24, which would be length 4. Let's check f(24):f(24)= max(f(2),f(3),f(4),f(6),f(8),f(12)) +1.f(2)=1, f(3)=1, f(4)=2, f(6)=2, f(8)=3, f(12)=3.So, max is 3, so f(24)=4.Similarly, f(16)= f(8)+1=4.So, f(16)=4, f(24)=4.Similarly, f(32) would be 5, but 32 is out.Therefore, the maximum f(n) in 2-31 is 4. So, the chromatic number is 4.Therefore, the minimal k is 4.Wait, but let's confirm. If the maximum chain length is 4, then the chromatic number is 4. But let's check if the graph is perfect, so that chromatic number equals clique number.Since we are dealing with a comparability graph (as mentioned before), which is a type of perfect graph, then indeed the chromatic number equals the clique number. Therefore, the answer should be 4.But let's verify with an example.Take numbers 2,4,8,16. Each divides the next, so they form a clique of size 4. Therefore, they all need different colors. Similarly, 3,6,12,24 also form a clique of size 4. Therefore, at least 4 colors are needed.Now, can we color the graph with 4 colors?Yes, by assigning each number a color equal to the length of the longest chain ending at that number. Since for any edge from m to n (m divides n), the chain ending at n must be longer than the chain ending at m. Therefore, colors can be assigned based on the length of the chain, ensuring that adjacent nodes have different colors.Alternatively, since the graph is a comparability graph, it can be colored with the number of colors equal to the size of the largest clique. Therefore, 4 colors suffice.Therefore, the minimal k is 4.Thus, problem 2 answer is 4.

✅:Problem 1 Solution:To determine the minimum number of colors needed to color integers 2 through 8 such that no two numbers where one is a multiple of the other share the same color, we construct a graph where edges connect such pairs. Key steps include:1. Identify Edges: Numbers connected if one is a multiple of the other: - 2 connected to 4, 6, 8 - 3 connected to 6 - 4 connected to 82. Graph Structure: The main component forms a graph with nodes 2, 3, 4, 6, 8. Nodes 5 and 7 are isolated.3. Coloring Analysis: - The subgraph containing 2, 4, 8 forms a triangle (cycle of length 3), requiring 3 colors. - Node 6 must differ from 2 and 3, and node 3 must differ from 6. Assigning colors carefully shows 3 colors suffice. - Isolated nodes (5, 7) can share any color used elsewhere.Answer: The smallest integer ( k ) is (boxed{3}).---Problem 2 Solution:For integers 2 through 31, the goal is similarly to find the minimum colors required. This problem reduces to finding the chromatic number of the corresponding graph, which equals the size of the largest clique (chain of divisibility):1. Longest Divisor Chain: - Chains like ( 2 rightarrow 4 rightarrow 8 rightarrow 16 ) or ( 3 rightarrow 6 rightarrow 12 rightarrow 24 ) have length 4. - No chain longer than 4 exists within 2-31.2. Chromatic Number: Since comparability graphs are perfect, the chromatic number equals the clique number (longest chain).Answer: The smallest integer ( k ) is (boxed{4}).

🔑:To solve this problem, we need to demonstrate the smallest integer k for which it is possible for Alice to color the integers within a given range such that no two numbers m and n (where one is a multiple of the other and m neq n) share the same color. We'll solve the two parts of the problem in succession.# 1) Coloring integers between 2 and 8:Let's start by coloring the integers from 2 to 8 naively and analyze the process.1. Color the integer 2: - Color 2 with color 1. 2. Color the integer 3: - Color 3 with color 1 since 2 does not divide 3.3. Color the integer 4: - 4 is divisible by 2, so it cannot be the same color as 2. - Color 4 with color 2.4. Color the integer 5: - 5 can be colored with color 1 as it is not divisible by 2 or 3. 5. Color the integer 6: - 6 is divisible by 2 and 3. Therefore, it cannot be the same color as 2 or 3. - Color 6 with color 2.6. Color the integer 7: - 7 can be colored with color 1 as it is not divisible by any previously colored integers. 7. Color the integer 8: - 8 is divisible by 2 and 4. Therefore, it cannot share the same color as 2 or 4. - Color 8 with color 3. Through this process, we can successfully color the integers from 2 to 8 using only 3 colors. Therefore, the smallest number of colors required is 3.We need to determine if we can necessarily use fewer than 3 colors: - Note that 8 is a multiple of 4 and 2, while 4 is a multiple of 2, thus 2, 4, and 8 must have different colors to satisfy the given conditions.- Thus, at least 3 different colors are mandatory to color the numbers 2, 4, and 8 differently.Hence the smallest integer k needed to color the integers from 2 to 8 is:boxed{3}# 2) Coloring integers between 2 and 31:Now extend our analysis for integers from 2 to 31.Firstly, observe that the critical requirement arises from the multiples of 2 and their powers because they are the most restrictive (many numbers need different colors):Considering powers of 2 between 2 and 31: - 2, 4, 8, 16 These are four powers of 2, and they give us a clue that potentially four colors may be needed due to them being sequential multiples. Let's construct the color assignment based on this idea:To ensure different coloring:1. 2 must be a different color from its higher multiples: - 4, 8, 16 2. 4 must differ from: - 8, 163. 8 must differ from: - 16Each power of 2 up to 16 requires a unique color:- Color(2) neq Color(4)- Color(4) neq Color(8)- Color(8) neq Color(16)Therefore, at least 4 different colors are needed to color appropriately in this range.Next we need to demonstrate that coloring integers 2 to 31 using exactly 4 colors can be accomplished:1. Assign color 1 to these integers: 2, 3, 5, 7, 11, 13, 17, 19, 23, 29, 31 2. Assign color 2 to these: 4, 6, 9, 10, 14, 15, 21, 22, 25, 263. Assign color 3 to these: 8, 12, 18, 20, 27, 28, 304. Assign color 4 to these: 16, 24The coloring complies with all the constraints required by the problem statement, confirming that 4 colors suffice.Therefore, for the integers from 2 to 31, the smallest number of colors k is:boxed{4}

❓:If (frac{139}{22} = a + frac{1}{b + frac{1}{c}}), where (a), (b), and (c) are positive integers, find the value of (a + b + c).

🧠:Okay, so I need to find the value of a + b + c where 139/22 is expressed as a continued fraction of the form a + 1/(b + 1/c). Hmm, right. Let me start by recalling how continued fractions work. I think this is similar to the Euclidean algorithm where you divide the numerator by the denominator to get an integer part and a remainder, then repeat the process with the reciprocal of the remainder. First, let me confirm that 139 divided by 22 is the given fraction. So, 139 ÷ 22. Let me compute that. 22 times 6 is 132, which is less than 139, and 22 times 7 is 154, which is too much. So the integer part a should be 6. Then the remainder is 139 - 22*6 = 139 - 132 = 7. So that gives me 139/22 = 6 + 7/22. But the problem wants it in the form a + 1/(b + 1/c). So 7/22 needs to be expressed as 1 divided by something. Let's see, 7/22 is equal to 1/(22/7). So now, 139/22 = 6 + 1/(22/7). Now, I need to express 22/7 as a continued fraction. Let's work on 22/7. That's 3 + 1/7 because 7*3 is 21, so 22 - 21 = 1. Wait, no, 22 divided by 7 is 3 with a remainder of 1. So 22/7 = 3 + 1/7. Therefore, substituting back, we have 139/22 = 6 + 1/(3 + 1/7). Wait, so that would mean a = 6, b = 3, and c = 7. Therefore, a + b + c = 6 + 3 + 7 = 16. But let me check if this is correct. Let's compute the continued fraction step by step.Starting with 139/22:1. Divide 139 by 22. The quotient is 6, remainder is 7. So, 139 = 22*6 + 7. Therefore, 139/22 = 6 + 7/22 = 6 + 1/(22/7).2. Now take 22/7. Divide 22 by 7. The quotient is 3, remainder is 1. So, 22 = 7*3 + 1. Therefore, 22/7 = 3 + 1/7. 3. Substitute back into the previous equation: 139/22 = 6 + 1/(3 + 1/7). So the continued fraction stops here because the last remainder is 1, and when we take the reciprocal of 1/7, we get 7, which is an integer. Therefore, the continued fraction representation is [6; 3, 7]. But the problem specifies the form a + 1/(b + 1/c). Let me verify that. In this case, the continued fraction has three terms: a = 6, then b = 3, and c = 7. So the expression is 6 + 1/(3 + 1/7), which matches the given form. So a = 6, b = 3, c = 7, so the sum a + b + c is 6 + 3 + 7 = 16. Wait, but let me make sure there isn't another way to represent this. Suppose I made a mistake in the steps. Let me double-check each division.First step: 139 divided by 22. 22*6 = 132. 139 - 132 = 7. Correct. So 139/22 = 6 + 7/22. Then 7/22 becomes 1/(22/7). Then 22 divided by 7 is 3 with a remainder of 1. So 22/7 = 3 + 1/7. Therefore, substituting back, yes, we have 6 + 1/(3 + 1/7). Is there a step further? If we continue the Euclidean algorithm, the next step would be to take 7/1, but since the remainder is 1, the algorithm stops. Therefore, the continued fraction terminates here. Hence, the representation is indeed [6; 3, 7], so a = 6, b = 3, c = 7. Therefore, the answer should be 16. Wait, but let me compute the actual value to check. Let's compute 3 + 1/7 = 22/7. Then 1/(22/7) = 7/22. Then 6 + 7/22 = (6*22 + 7)/22 = (132 + 7)/22 = 139/22. Which matches the original fraction. So that's correct. Alternatively, maybe the problem expects a different form? Let me check the problem statement again. It says: "If 139/22 = a + 1/(b + 1/c), where a, b, and c are positive integers, find the value of a + b + c." So the form is specifically a single continued fraction step, i.e., a plus 1 over (b plus 1 over c). That is, two layers of fractions. So yes, the way I solved it is exactly that. So 6 + 1/(3 + 1/7), so a=6, b=3, c=7. Therefore, the sum is 16. Alternatively, sometimes continued fractions can be represented in different ways if there's a 1 in the last denominator. For example, if the last term is 1, you can combine it with the previous term. But in this case, the last term is 7, which is greater than 1, so that's not an issue here. Therefore, the representation is unique here. Therefore, the answer is 16. But just to make sure, let me think if there is another possible decomposition. Suppose someone tries to write 139/22 as a + 1/(b + 1/c). Let's suppose that instead of taking the first quotient as 6, they tried something else. But since 22*6=132 is less than 139, and 22*7=154 is more than 139, the integer part must be 6. So a is definitely 6. Then the fractional part is 7/22. To express 7/22 as 1 over something, it's 1/(22/7). 22 divided by 7 is 3 with remainder 1, so 22/7 is 3 + 1/7, so the next term is 3 and then 7. So there's no ambiguity here. Another way to check is to compute the continued fraction coefficients. The standard method for continued fractions is to repeatedly take the integer part and invert the remainder. So starting with 139/22:1. 139 ÷ 22 = 6, remainder 7. So first coefficient is 6.2. Take reciprocal of 7/22, which is 22/7.3. 22 ÷ 7 = 3, remainder 1. Second coefficient is 3.4. Take reciprocal of 1/7, which is 7.5. 7 ÷ 1 = 7, remainder 0. Third coefficient is 7.Therefore, the continued fraction is [6; 3, 7], which corresponds to 6 + 1/(3 + 1/7). So that's exactly the form given in the problem. Thus, a=6, b=3, c=7, sum is 16. I think this is solid. So the answer should be 16.

✅:To express (frac{139}{22}) as a continued fraction of the form (a + frac{1}{b + frac{1}{c}}), we follow these steps:1. Find the integer part (a): [ frac{139}{22} = 6 + frac{7}{22} quad text{(since (22 times 6 = 132) and (139 - 132 = 7))}. ] Thus, (a = 6).2. Convert the fractional part (frac{7}{22}) into a reciprocal: [ frac{7}{22} = frac{1}{frac{22}{7}}. ] Now, (frac{22}{7}) becomes the new fraction to analyze.3. Find the integer part (b) for (frac{22}{7}): [ frac{22}{7} = 3 + frac{1}{7} quad text{(since (7 times 3 = 21) and (22 - 21 = 1))}. ] Thus, (b = 3).4. Identify (c) from the remaining fraction: The remaining fractional part is (frac{1}{7}), so (c = 7).Putting it all together:[frac{139}{22} = 6 + frac{1}{3 + frac{1}{7}}.]Summing the values:[a + b + c = 6 + 3 + 7 = boxed{16}.]

🔑:1. We start with the given equation: [ frac{139}{22} = a + frac{1}{b + frac{1}{c}} ]2. First, we determine the integer part a by performing integer division of 139 by 22: [ a = leftlfloor frac{139}{22} rightrfloor = 6 ]3. Subtracting a = 6 from frac{139}{22}, we get: [ frac{139}{22} - 6 = frac{139 - 6 cdot 22}{22} = frac{139 - 132}{22} = frac{7}{22} ]4. This can be rewritten as: [ frac{7}{22} = frac{1}{frac{22}{7}} ]5. Next, we express frac{22}{7} as a mixed number: [ frac{22}{7} = leftlfloor frac{22}{7} rightrfloor + frac{1}{frac{7}{7} div frac{7}{22}} = 3 + frac{1}{frac{22}{7} - 3} ]6. Calculating the fractional part: [ frac{22}{7} - 3 = frac{22 - 3 cdot 7}{7} = frac{22 - 21}{7} = frac{1}{7} ]7. Thus: [ frac{7}{22} = frac{1}{3 + frac{1}{7}} ]8. This helps identify: [ b = 3 quad text{and} quad c = 7 ]9. Finally, we sum the values of a, b, and c: [ a + b + c = 6 + 3 + 7 = 16 ] Conclusion:[boxed{16}]

❓:Prove that if the product of two positive numbers is equal to 1, then the sum of these numbers is at least 2.

🧠:Okay, so I need to prove that if the product of two positive numbers is 1, then their sum is at least 2. Hmm, let's start by understanding the problem. We have two positive numbers, say a and b, such that ab = 1. We need to show that a + b ≥ 2. Alright, I remember there's something called the Arithmetic Mean-Geometric Mean Inequality (AM-GM) that relates the sum and product of numbers. Let me recall... AM-GM states that for non-negative numbers, the arithmetic mean is greater than or equal to the geometric mean. So for two numbers, (a + b)/2 ≥ √(ab). Since ab = 1, then √(ab) = 1. So substituting in, we get (a + b)/2 ≥ 1, which implies a + b ≥ 2. That seems straightforward, but maybe I should verify each step to make sure I didn't skip anything.Wait, but maybe the problem expects a proof without using AM-GM since it's a fundamental result. Let me try to approach it from scratch. Suppose a and b are positive real numbers with ab = 1. Let me consider the difference between a + b and 2. If I can show that this difference is always non-negative, then a + b ≥ 2. So let's compute (a + b) - 2. Since ab = 1, maybe I can express one variable in terms of the other. Let's solve for b: b = 1/a. Then substitute into the expression: a + 1/a - 2. So we need to show that a + 1/a - 2 ≥ 0. Let's see if this can be rewritten. If I write it as (a^2 + 1 - 2a)/a, that simplifies to (a - 1)^2 / a. Since a is positive, the denominator a is positive, and the numerator (a - 1)^2 is always non-negative because squares are non-negative. Therefore, the entire expression (a - 1)^2 / a ≥ 0. Hence, a + 1/a - 2 ≥ 0, so a + 1/a ≥ 2. Therefore, the sum a + b is at least 2. But wait, when does equality hold? If (a - 1)^2 = 0, then a = 1. Then b = 1/a = 1. So equality occurs when a = b = 1. That makes sense. So the minimum sum is 2, achieved when both numbers are 1. Alternatively, maybe using calculus. Let's consider the function f(a) = a + 1/a for a > 0. We can find its minimum. Take the derivative: f'(a) = 1 - 1/a^2. Set derivative to zero: 1 - 1/a^2 = 0 → a^2 = 1 → a = 1 (since a > 0). Then the second derivative is f''(a) = 2/a^3, which is positive when a = 1, so it's a minimum. Therefore, the minimum of f(a) is f(1) = 1 + 1 = 2. Therefore, a + 1/a ≥ 2 for all positive a. Hence, the sum is at least 2. Another approach: Maybe using the fact that (a + b)^2 = a^2 + 2ab + b^2. Since ab = 1, then (a + b)^2 = a^2 + 2*1 + b^2 = a^2 + b^2 + 2. But I'm not sure if this helps directly. However, since a and b are positive, we can use inequalities involving squares. For example, (a - b)^2 ≥ 0 ⇒ a^2 - 2ab + b^2 ≥ 0 ⇒ a^2 + b^2 ≥ 2ab. Since ab = 1, this gives a^2 + b^2 ≥ 2. Then adding 2 to both sides: a^2 + b^2 + 2 ≥ 4. But (a + b)^2 = a^2 + 2ab + b^2 = a^2 + b^2 + 2*1 = a^2 + b^2 + 2. Therefore, (a + b)^2 ≥ 4 ⇒ a + b ≥ 2, since a + b is positive. That's another way. So whether using algebraic manipulation, calculus, or expanding squares, we arrive at the same conclusion. Alternatively, suppose we assume that a + b < 2. Then since ab = 1, we can set a = 1 - h and b = 1 + k for some h, k > 0. Wait, but this might complicate things. Alternatively, if a + b < 2, then using AM ≥ GM, (a + b)/2 ≥ √(ab) = 1, so (a + b)/2 ≥ 1 ⇒ a + b ≥ 2, which contradicts the assumption. Therefore, a + b cannot be less than 2.Wait, that's actually using AM-GM again. So maybe all these approaches are connected. The key idea is that the arithmetic mean is at least the geometric mean, and equality holds when all numbers are equal. So in summary, regardless of the method—whether through algebraic manipulation, calculus, or applying known inequalities like AM-GM—we find that the sum of two positive numbers whose product is 1 must be at least 2, with equality if and only if both numbers are 1.

✅:To prove that if the product of two positive numbers is 1, their sum is at least 2, we use the following reasoning:Step 1: Let ( a, b > 0 ) such that ( ab = 1 ). We aim to show ( a + b geq 2 ).Step 2: Apply the Arithmetic Mean-Geometric Mean Inequality (AM-GM): For non-negative numbers, [frac{a + b}{2} geq sqrt{ab}.] Substitute ( ab = 1 ): [frac{a + b}{2} geq sqrt{1} = 1.] Multiply both sides by 2: [a + b geq 2.]Alternative Algebraic Proof: Express ( b = frac{1}{a} ) (since ( ab = 1 )). The sum becomes: [a + frac{1}{a}.] Consider the inequality: [a + frac{1}{a} - 2 = frac{a^2 - 2a + 1}{a} = frac{(a - 1)^2}{a} geq 0.] Since ( (a - 1)^2 geq 0 ) and ( a > 0 ), the expression is non-negative. Thus: [a + frac{1}{a} geq 2.]Equality Condition: ( a + b = 2 ) if and only if ( a = b = 1 ).Final Conclusion: For any two positive numbers ( a, b ) with ( ab = 1 ), their sum satisfies ( a + b geq 2 ).[boxed{a + b geq 2}]

🔑:# Method 1: 1. Consider two positive numbers ( x ) and ( frac{1}{x} ), such that their product is 1, ( x cdot frac{1}{x} = 1 ).2. Clearly, since ( x > 0 ), it follows that ( frac{1}{x} > 0 ).3. We aim to show that their sum ( x + frac{1}{x} ) is at least 2. Let's compare this sum with 2. [ x + frac{1}{x} geq 2 ]4. To verify this inequality, consider the following transformation: [ x + frac{1}{x} - 2 geq 0 ]5. Combine the terms over a common denominator: [ frac{x^2 - 2x + 1}{x} geq 0 ]6. Factorize the numerator: [ frac{(x - 1)^2}{x} geq 0 ]7. Note that the numerator ((x - 1)^2) is a square term, which is always non-negative, and the denominator (x) is positive by assumption.8. Therefore, [ frac{(x - 1)^2}{x} geq 0 ]Since both the numerator and the denominator are positive, the inequality holds true, proving that ( x + frac{1}{x} geq 2 ).# Method 2:1. Use the Cauchy-Schwarz inequality in its arithmetic mean-geometric mean (AM-GM) form: [ frac{a + b}{2} geq sqrt{a cdot b} quad text{for} quad a geq 0, ~ b geq 0 ]2. Substitute ( a = x ) and ( b = frac{1}{x} ): [ frac{x + frac{1}{x}}{2} geq sqrt{x cdot frac{1}{x}} ]3. Compute the right-hand side: [ sqrt{x cdot frac{1}{x}} = sqrt{1} = 1 ]4. Thus, [ frac{x + frac{1}{x}}{2} geq 1 ]5. Multiply both sides by 2 to obtain: [ x + frac{1}{x} geq 2 ]The equality ( x + frac{1}{x} = 2 ) holds when ( x = 1 ). In conclusion, no matter the method used, [boxed{x + frac{1}{x} geq 2}]

❓:Let ( A_{1}, A_{2}, ldots, A_{2017} ) be the vertices of a regular polygon with 2017 sides. Prove that there exists a point ( P ) in the plane of the polygon such that the vectorsum_{k=1}^{2017} k frac{overrightarrow{P A_{k}}}{left|overrightarrow{P A_{k}}right|^{5}}is the zero vector.(The notation ( |overrightarrow{X Y}| ) represents the length of the vector ( overrightarrow{X Y} ).)

🧠:Okay, so I have this problem about a regular 2017-gon, and I need to prove there's a point P such that the sum of vectors from P to each vertex, each multiplied by their respective index k and divided by the fifth power of their distance from P, equals zero. Hmm. Let me try to unpack this step by step.First, let's restate the problem to make sure I understand it. We have a regular polygon with 2017 sides, vertices labeled A₁ through A₂₀₁₇. We need to find a point P in the plane such that when we take each vector PAₖ (which is the vector from P to Aₖ), multiply it by k, divide by the fifth power of its length, and sum all these vectors up, the result is the zero vector. So, mathematically, the condition is:∑ₖ=₁²⁰¹⁷ k * (PAₖ / ||PAₖ||⁵) = 0I need to prove such a point P exists. Let me start by recalling similar problems. This seems related to the concept of a center of mass or equilibrium points, where forces or vectors sum to zero. However, in this case, the coefficients aren't masses but are proportional to k, and the denominator is ||PAₖ||⁵ instead of ||PAₖ||² which would be typical for gravitational or electric fields. Wait, gravitational fields have a 1/r² dependence, but here we have 1/r⁵. That's a much steeper drop-off with distance. Interesting. Maybe this is more like a higher-dimensional analogy, but we're in the plane. Maybe it's a contrived vector field for the problem.But how to approach proving such a point exists? Since the polygon is regular, there's a lot of symmetry. Maybe the point P is the center of the polygon? Let's check that first.If P is the center, then all vectors PAₖ have the same length, right? Because in a regular polygon, the distance from the center to each vertex is the same. Let's denote that distance as R. Then ||PAₖ|| = R for all k, so each term in the sum becomes k * (PAₖ / R⁵). The sum would then be (1/R⁵) * ∑ₖ=₁²⁰¹⁷ k PAₖ. So the question reduces to whether ∑ₖ=₁²⁰¹⁷ k PAₖ is zero. But PAₖ here are vectors from the center to each vertex. Since the polygon is regular and symmetric, the vectors PAₖ are symmetrically distributed around the circle. However, each vector is weighted by k. The index k goes from 1 to 2017, so the weights are different for each vertex. Since the polygon is regular, the vertices are equally spaced in angle. However, the weights k are increasing linearly. So, if we number the vertices in order around the polygon, each subsequent vertex has a higher weight. If the center is P, then the vectors PAₖ are radial vectors, each pointing from the center to a vertex. The sum ∑k PAₖ would need to be zero for the center to work. But since the weights k are different, the symmetry is broken. For example, in a regular polygon, if all weights are equal, the sum would be zero due to symmetry. But here, with varying k, the sum might not be zero. Let me test this with a smaller polygon.Take a regular triangle with vertices A₁, A₂, A₃. Suppose we take the center as P. Then PA₁, PA₂, PA₃ are vectors 120 degrees apart. The sum would be 1*PA₁ + 2*PA₂ + 3*PA₃. Is this zero? Let's compute it. Let’s assign coordinates: suppose the center is at (0,0), and the triangle is inscribed in a unit circle. Then PA₁ is (1,0), PA₂ is (-1/2, √3/2), PA₃ is (-1/2, -√3/2). Then 1*PA₁ + 2*PA₂ + 3*PA₃ would be (1,0) + 2*(-1/2, √3/2) + 3*(-1/2, -√3/2). Calculating the x-components: 1 + 2*(-1/2) + 3*(-1/2) = 1 -1 - 3/2 = -3/2y-components: 0 + 2*(√3/2) + 3*(-√3/2) = √3 - (3√3)/2 = (-√3)/2So the sum is (-3/2, -√3/2) ≠ 0. Hence, for the triangle, the center is not a solution. So in the original problem, the center is unlikely to be the point P. So maybe another point?Alternatively, perhaps there's a point such that the asymmetrical weights balance out due to the positioning. Since the problem is about existence, maybe we can use a fixed point theorem or some topological argument. For example, considering the vector field defined by the sum and showing that it must have a zero somewhere.Let me think. If I consider the function F(P) = ∑ₖ=₁²⁰¹⁷ k * (PAₖ / ||PAₖ||⁵). Then F is a vector field defined on the plane (except at the points Aₖ, where it's undefined due to division by zero). To prove that there's a point P where F(P) = 0.Since F is a continuous vector field (except at the vertices), maybe we can use the Poincaré-Hopf theorem or some other fixed point theorem. But Poincaré-Hopf relates to vector fields on manifolds, requiring certain conditions. Alternatively, perhaps use the Brouwer fixed point theorem by constructing a map whose fixed point corresponds to the zero of the vector field.Alternatively, think about taking a large disk containing the polygon and considering the vector field on the boundary. If the vector field points inward everywhere on the boundary, then by Brouwer's theorem, there must be a fixed point inside. But how to analyze the direction of the vector field on a large circle?Alternatively, consider the behavior of F(P) as P approaches infinity. Let's see, as P goes to infinity in some direction, the vectors PAₖ are approximately pointing in the direction away from P to each Aₖ. Wait, actually, if P is very far away, then all the points Aₖ are approximately in the same direction relative to P. So each PAₖ vector is approximately pointing towards the center of the polygon (if P is far away). Wait, no, if P is very far in some direction, then all the Aₖ are near the origin (assuming the polygon is centered at the origin), so PAₖ would be vectors from P to the origin. Wait, maybe the polygon is fixed in space. Let's assume the polygon is centered at the origin for simplicity.If the polygon is centered at the origin, then each Aₖ is a point on the circle of radius R. Let’s fix the polygon at the origin. Let’s denote the center as O. So if we consider a point P somewhere in the plane, then PAₖ = Aₖ - P (if we consider vectors as position vectors relative to the origin). So the vector PAₖ is the vector from P to Aₖ, which is Aₖ - P.Therefore, the sum F(P) = ∑ₖ=₁²⁰¹⁷ k * (Aₖ - P) / ||Aₖ - P||⁵.We need to show that there exists a P such that F(P) = 0.Hmm, perhaps it's helpful to write this as F(P) = ∑ₖ=₁²⁰¹⁷ k (Aₖ - P) / ||Aₖ - P||⁵ = 0.Let’s split this into two sums:F(P) = ∑ₖ=₁²⁰¹⁷ k Aₖ / ||Aₖ - P||⁵ - P ∑ₖ=₁²⁰¹⁷ k / ||Aₖ - P||⁵ = 0Therefore, we can write:∑ₖ=₁²⁰¹⁷ k Aₖ / ||Aₖ - P||⁵ = P ∑ₖ=₁²⁰¹⁷ k / ||Aₖ - P||⁵Let’s denote S(P) = ∑ₖ=₁²⁰¹⁷ k / ||Aₖ - P||⁵and V(P) = ∑ₖ=₁²⁰¹⁷ k Aₖ / ||Aₖ - P||⁵Then the equation becomes V(P) = P S(P), so P = V(P)/S(P), provided S(P) ≠ 0.Therefore, if we can find a fixed point P such that P = V(P)/S(P), then F(P) = 0.This looks like a fixed point equation. So perhaps we can define a function T(P) = V(P)/S(P), and then show that T has a fixed point. To apply Brouwer's fixed point theorem, we need a continuous function from a convex compact subset of R² to itself. If we can find such a set where T maps into itself, then Brouwer's theorem would give a fixed point.But how to choose such a set? Let's analyze the behavior of T(P) as P approaches infinity. Suppose ||P|| is very large. Then each ||Aₖ - P|| ≈ ||P||, since Aₖ are all at a fixed distance from the origin. So S(P) ≈ ∑ₖ=₁²⁰¹⁷ k / ||P||⁵ = (∑ₖ=₁²⁰¹⁷ k) / ||P||⁵ = [2017*(2017 +1)/2] / ||P||⁵.Similarly, V(P) ≈ ∑ₖ=₁²⁰¹⁷ k Aₖ / ||P||⁵. Since the Aₖ are points on a regular polygon, their centroid (if all weights k were equal) would be zero due to symmetry. But here, the weights k are increasing linearly. However, since the polygon is regular and the weights are assigned in order around the polygon, maybe the sum ∑k Aₖ is not zero? Let's check.Let’s compute ∑ₖ=₁²⁰¹⁷ k Aₖ. Since the polygon is regular, each Aₖ can be represented in the complex plane as R e^(2πi(k-1)/2017), where R is the radius. Then the sum becomes ∑ₖ=₁²⁰¹⁷ k R e^(2πi(k-1)/2017). Let’s denote ω = e^(2πi/2017), so Aₖ = R ω^{k-1}. Then ∑ₖ=₁²⁰¹⁷ k Aₖ = R ∑ₖ=₁²⁰¹⁷ k ω^{k-1}.This is a geometric series with coefficients increasing linearly. There is a formula for such sums. Let’s recall that ∑ₖ=₁^n k z^{k} = z(1 - (n+1)z^n + n z^{n+1}) / (1 - z)^2. But here, we have ∑ₖ=₁²⁰¹⁷ k ω^{k-1} = (1/ω) ∑ₖ=₁²⁰¹⁷ k ω^{k} = (1/ω) [ ω(1 - (2017 +1)ω^{2017} + 2017 ω^{2018} ) / (1 - ω)^2 ) ]But ω^{2017} = e^(2πi *2017/2017) = e^(2πi) = 1. Similarly, ω^{2018} = ω. Therefore, substituting:(1/ω)[ ω(1 - 2018*1 + 2017 ω ) / (1 - ω)^2 ) ] = [ (1 - 2018 + 2017 ω ) / (1 - ω)^2 ) ].Simplify numerator: 1 - 2018 + 2017 ω = -2017 + 2017 ω = 2017(ω - 1). Therefore:2017(ω - 1) / (1 - ω)^2 ) = 2017 ( - (1 - ω) ) / (1 - ω)^2 ) = -2017 / (1 - ω).Therefore, the sum ∑ₖ=₁²⁰¹⁷ k ω^{k-1} = -2017 / (1 - ω).Thus, ∑ₖ=₁²⁰¹⁷ k Aₖ = R * (-2017) / (1 - ω).But ω = e^(2πi/2017), so 1 - ω = 1 - e^(2πi/2017) = e^(πi/2017) * [ e^(-πi/2017) - e^(πi/2017) ] = -2i e^(πi/2017) sin(π/2017).Therefore, 1/(1 - ω) = -1/(2i e^(πi/2017) sin(π/2017)) ) = e^(-πi/2017)/(2 sin(π/2017)) * i.But maybe the exact value isn't important. The key point is that ∑ₖ=₁²⁰¹⁷ k Aₖ is a non-zero vector. So, as ||P|| becomes large, V(P) ≈ (∑ₖ=₁²⁰¹⁷ k Aₖ ) / ||P||⁵, and S(P) ≈ (∑ₖ=₁²⁰¹⁷ k ) / ||P||⁵ = [2017*2018/2] / ||P||⁵. Therefore, T(P) = V(P)/S(P) ≈ [ (∑ₖ=₁²⁰¹⁷ k Aₖ ) / ||P||⁵ ] / [ (2017*2018/2) / ||P||⁵ ] = (2/(2017*2018)) ∑ₖ=₁²⁰¹₇ k Aₖ.So as P goes to infinity in any direction, T(P) approaches a constant vector, namely (2/(2017*2018)) ∑ₖ=₁²⁰¹₇ k Aₖ, which is a fixed non-zero vector. Therefore, if we go far enough in the direction opposite to this vector, T(P) would be approximately pointing towards that constant vector. Hence, outside a sufficiently large disk, T(P) maps points to within some bounded region. Therefore, if we take a sufficiently large closed disk D centered at the origin, then T(P) maps D into itself. Because for P on the boundary of D, T(P) is approximately a constant vector, which is much smaller in magnitude than D's radius (since T(P) is scaled by 1/(2017*2018) times a vector whose magnitude is related to the size of the polygon). Wait, actually, the original polygon has a fixed size, say radius R. Then ∑ₖ=₁²⁰¹₇ k Aₖ has a magnitude proportional to 2017*R, so the constant vector is (2/(2017*2018)) * 2017*R * something, which simplifies to (2R/2018) * something. Depending on R, but if the original polygon is of radius R, then if we choose the disk D to have a radius larger than (2R/2018)*C, where C is some constant, then T(P) would map D into itself. Therefore, by Brouwer's fixed point theorem, since D is a convex compact set and T is continuous (assuming we avoid the points Aₖ, but since the original problem allows P to be any point in the plane, except the vertices, but we can choose D such that it doesn't include any Aₖ, which are at radius R, so if we take D with radius larger than R, then inside D, P is not any Aₖ, so T is continuous on D. Wait, but D needs to exclude the points Aₖ. If the polygon is inside D, but the disk D is chosen such that all Aₖ are inside D. Wait, if D is a large enough disk, then all Aₖ are inside D, and T(P) is defined for all P in D except the Aₖ. But we need T(P) to be defined on all of D. However, since the problem states P is in the plane of the polygon, which includes all points except the Aₖ. So, as long as P is not one of the Aₖ, T(P) is defined. But in order to apply Brouwer, we need T to be continuous on D. Therefore, we have to ensure that D does not contain any of the Aₖ. But if the polygon is inside D, then D does contain the Aₖ. Hmm, this is a problem. Because if P approaches Aₖ, then ||Aₖ - P|| approaches zero, making the terms in S(P) and V(P) blow up. Therefore, T(P) is not defined at the Aₖ, and near the Aₖ, T(P) can become arbitrarily large. Therefore, we cannot directly apply Brouwer's theorem on a disk containing the polygon. Alternative approach: Maybe consider the vector field F(P) and show that its degree around a large circle is non-zero, implying the existence of a zero inside. Alternatively, use the fact that the vector field F(P) is asymptotically constant, and if that constant is non-zero, then the degree is zero, which might not help. Alternatively, note that if you consider the vector field F(P) near infinity, it's approximately a constant vector. If we can show that F(P) has a direction that varies as P moves around a large circle, maybe leading to a non-zero winding number, but since it's approximates a constant vector, the winding number would be zero. So that might not help.Alternatively, think about the problem in terms of symmetry. Since the polygon is regular, maybe the desired point P lies on the axis of symmetry. However, the weights k are assigned in order around the polygon, which breaks rotational symmetry. But 2017 is a prime number, and in particular, odd. Therefore, the polygon doesn't have reflection symmetry that reverses the order of the vertices. Hmm, but each vertex is symmetric in the sense that rotations by 2π/2017 map the polygon onto itself. But since the weights k are assigned in a rotational order, the problem isn't symmetric under rotation. For example, rotating the polygon by one step would relabel each vertex k to k+1 (mod 2017), but the weights would shift as well. Therefore, the problem lacks rotational symmetry. However, maybe there is a reflection symmetry or some other symmetry.Alternatively, consider using the Intermediate Value Theorem in some way. For example, if we can show that the vector field F(P) has a component that changes sign as P moves along a certain path, then there must be a point where it's zero. However, since F(P) is a vector in the plane, we'd need to consider two-dimensional versions, which complicates things.Alternatively, think about the problem in complex numbers. Let's model the plane as the complex plane. Let the polygon be inscribed in the unit circle for simplicity, so each Aₖ is at position e^(2πi(k-1)/2017). Let P be a complex number z. Then PAₖ is Aₖ - z, and ||PAₖ|| is |Aₖ - z|. Then the condition is:∑ₖ=₁²⁰¹⁷ k (Aₖ - z) / |Aₖ - z|⁵ = 0We need to prove there exists a complex number z (not equal to any Aₖ) such that this sum is zero.Let’s denote f(z) = ∑ₖ=₁²⁰¹⁷ k (Aₖ - z) / |Aₖ - z|⁵We need to show f(z) = 0 has a solution.Note that |Aₖ - z|⁵ is ( (Aₖ - z)(overline{Aₖ - z}) )^(5/2) = |Aₖ - z|^5. So in complex terms, (Aₖ - z)/|Aₖ - z|⁵ is (Aₖ - z) / |Aₖ - z|⁵ = (Aₖ - z) / |Aₖ - z|⁵. But in complex analysis, expressions like (Aₖ - z)/|Aₖ - z|³ would correspond to a vector field in 3D, but here we're in 2D with a different exponent. Not sure if complex analysis tools apply directly.Alternatively, consider that f(z) is a continuous function except at the points Aₖ. If we can show that f(z) has a singularity somewhere or that it's possible to apply the Poincaré-Hopf theorem by considering the indices around each Aₖ and at infinity, but I might be getting out of my depth here.Wait, maybe think about the function f(z) as a vector field and compute its winding number around a large circle. If the vector field approaches a constant direction at infinity, then the winding number would be zero. But if we can show that as we traverse a large circle, the vector field f(z) completes a non-zero number of rotations, then there must be a zero inside. However, if f(z) approaches a constant vector at infinity, then the winding number would be zero, implying by the Poincaré-Hopf theorem that the sum of the indices at the singularities (the Aₖ) plus the index at infinity equals the Euler characteristic of the sphere, which is 2. But since we're on the plane, maybe different.Alternatively, consider that each term in the sum is a vector pointing from z to Aₖ, scaled by k and divided by |Aₖ - z|⁵. So each term is like a force pulling towards Aₖ with strength k / |Aₖ - z|⁵. So the total force is the sum of these pulls. We need to find an equilibrium point where all these pulls balance out.In physics, such equilibrium points often exist between masses. For example, between two masses, there is a Lagrange point. But with 2017 masses arranged in a regular polygon and with different strengths, maybe there's a point where the forces balance. However, since the strengths k vary with the vertex, the symmetry is broken, so the equilibrium point might not be at the center. Alternatively, think of this as an optimization problem. Perhaps the function ∑ₖ=₁²⁰¹⁷ k / ||PAₖ||⁴ has a critical point, since the gradient of such a function would be related to the sum given. Wait, let's see. Suppose we define a potential function U(P) = ∑ₖ=₁²⁰¹⁷ k / (4 ||PAₖ||⁴). Then the gradient of U(P) would be -∑ₖ=₁²⁰¹⁷ k (PAₖ) / ||PAₖ||⁶ * 2. Wait, maybe not. Let's compute the gradient.The gradient of 1/||PAₖ||ⁿ with respect to P is -n (PAₖ) / ||PAₖ||^(n+2). So, if U(P) = ∑ₖ=₁²⁰¹⁷ k / ||PAₖ||⁴, then the gradient ∇U(P) = ∑ₖ=₁²⁰¹⁷ k * (-4)(PAₖ) / ||PAₖ||⁶. But our vector field is F(P) = ∑ₖ=₁²⁰¹⁷ k (PAₖ) / ||PAₖ||⁵. So ∇U(P) = -4 F(P). Therefore, if we can show that the potential U(P) has a critical point, then F(P) = 0 there. Therefore, proving the existence of a critical point for U(P) would suffice. Potential functions like U(P) are known to have critical points under certain conditions. For example, if U(P) is coercive (goes to infinity as ||P|| approaches infinity), and is smooth except at the Aₖ, then by the extreme value theorem, U(P) would attain a minimum somewhere, which is a critical point. However, wait, as ||P|| approaches infinity, what happens to U(P)? Each term is k / ||PAₖ||⁴. If P is very far from all Aₖ, then ||PAₖ|| ≈ ||P||, so U(P) ≈ ∑ₖ=₁²⁰¹⁷ k / ||P||⁴ = [2017*2018/2] / ||P||⁴, which tends to zero as ||P|| tends to infinity. Therefore, U(P) tends to zero at infinity. But near each Aₖ, U(P) tends to infinity. Therefore, U(P) doesn't have a global minimum, but rather has local minima near the points Aₖ. However, perhaps there's a saddle point somewhere else? Wait, but how do we know there's a critical point? Maybe using the mountain pass theorem or another result from calculus. Alternatively, consider that the potential U(P) is a smooth function on the plane minus the points Aₖ. Since U(P) approaches zero at infinity and goes to infinity near each Aₖ, there must be at least one local minimum somewhere. Wait, but wait: if all the "hills" are near the Aₖ, and the function tends to zero at infinity, maybe the function doesn't have a global minimum, but has local minima. However, the function is bounded below by zero, so the infimum is zero, but it's never achieved. Wait, no. Wait, U(P) is always positive, and tends to zero at infinity. Therefore, the infimum is zero, but it's not achieved. Therefore, there is no global minimum. However, there could be local minima or maxima. But how does that help?Alternatively, think of U(P) as having a single connected region around infinity where it's decreasing towards zero, and near each Aₖ, it's increasing to infinity. Therefore, the function might have saddle points somewhere between the Aₖ. But how to guarantee that? Maybe using Morse theory, but I need more concrete arguments.Alternatively, consider that the function U(P) is invariant under any symmetry that the problem has. But given that the weights k break the rotational symmetry, the potential U(P) doesn't inherit any symmetries from the polygon. Therefore, it's difficult to exploit symmetry to find critical points.Wait, perhaps use a deformation argument. Suppose we start with all weights equal, then the center would be a critical point due to symmetry. Then, as we smoothly vary the weights to become unequal, the critical point might move but persist. However, since 2017 is odd, and the weights are arranged in order, perturbing the symmetric case might break the critical point, but perhaps not destroy it entirely. However, this is vague. Maybe using the implicit function theorem?Alternatively, think about the system of equations given by F(P) = 0. This is a system of two equations (x and y components) in two variables (coordinates of P). If we can show that the equations have a solution, perhaps using a degree theory argument. If the mapping F(P) has a non-zero degree over a certain region, then there must be a solution. Alternatively, consider taking the dot product of F(P) with a vector and integrating around a closed loop, but I’m not sure.Wait, going back to the idea of using the Brouwer fixed point theorem. Let me formalize that. Suppose we can find a closed disk D such that T(P) maps D into itself. Then, since T is continuous (on D, as long as D does not contain any Aₖ), Brouwer's theorem would imply a fixed point exists in D. But how to choose D such that T(P) is inside D?Earlier, we saw that as P approaches infinity, T(P) approaches a constant vector C = (2/(2017*2018)) ∑ₖ=₁²⁰¹₇ k Aₖ. Let's compute the magnitude of C. Since each Aₖ is on the unit circle (assuming the polygon has radius 1), then ||Aₖ|| = 1. Then, ∑ₖ=₁²⁰¹₇ k Aₖ is a vector whose magnitude we can estimate.||∑ₖ=₁²⁰¹₇ k Aₖ|| ≤ ∑ₖ=₁²⁰¹₇ k ||Aₖ|| = ∑ₖ=₁²⁰¹₇ k = 2017*2018/2. But this is a very loose bound. The actual sum might be much smaller due to cancellation. For example, when we computed for n=3, the sum was non-zero but smaller in magnitude.But in the complex plane representation, we had ∑ₖ=₁²⁰¹₇ k Aₖ = -2017/(1 - ω), where ω is a primitive 2017th root of unity. The magnitude of this is 2017 / |1 - ω|. Since |1 - ω| = 2 sin(π/2017) ≈ 2π/2017 for small angles. Therefore, ||∑ₖ=₁²⁰¹₇ k Aₖ|| ≈ 2017 / (2π/2017) ) = 2017² / (2π). So this is a rather large vector if the polygon has radius 1. Therefore, the constant vector C = (2/(2017*2018)) * ∑ₖ=₁²⁰¹₇ k Aₖ would have magnitude ≈ (2/(2017*2018)) * (2017² / (2π)) ) = 2017/(2018π) ≈ 1/π. So C has a magnitude of approximately 1/π if the polygon is of radius 1. Therefore, if our disk D has radius, say, 1, then as P goes to infinity, T(P) approaches approximately 1/π, which is less than 1. Wait, but if we take D to be a disk of radius R > ||C||, then for P outside of D, T(P) is approximately C, which is inside D if R > ||C||. Wait, no: if P is outside D, T(P) approaches C. So if we take a disk D with radius larger than ||C||, then points P outside D would map to T(P) ≈ C, which is inside D. Therefore, if we take a sufficiently large disk D with radius R > ||C||, then T(P) maps the exterior of D to the interior. But for Brouwer's theorem, we need T to map D to itself. Hmm. Wait, if we consider P inside D, is T(P) also inside D? Not necessarily. But if we can show that for P on the boundary of D, T(P) is inside D, then T maps D to itself. So suppose we take a disk D centered at the origin with radius R. For P on the boundary of D (||P|| = R), we need to show that ||T(P)|| < R. Let's estimate ||T(P)||.T(P) = V(P)/S(P). Let's estimate V(P) and S(P).First, S(P) = ∑ₖ=₁²⁰¹⁷ k / ||Aₖ - P||⁵.Each term k / ||Aₖ - P||⁵. If ||P|| = R, and the polygon is inscribed in a circle of radius, say, 1 (for simplicity), then ||Aₖ|| = 1. Therefore, ||Aₖ - P|| ≥ ||P|| - ||Aₖ|| = R - 1. Therefore, each term k / ||Aₖ - P||⁵ ≤ k / (R - 1)^5. Therefore, S(P) ≤ ∑ₖ=₁²⁰¹⁷ k / (R - 1)^5 = [2017*2018/2] / (R - 1)^5.Similarly, V(P) = ∑ₖ=₁²⁰¹⁷ k Aₖ / ||Aₖ - P||⁵. The norm ||V(P)|| ≤ ∑ₖ=₁²⁰¹⁷ k ||Aₖ|| / ||Aₖ - P||⁵ = ∑ₖ=₁²⁰¹₇ k / ||Aₖ - P||⁵ = S(P). Therefore, ||V(P)|| ≤ S(P). Hence, ||T(P)|| = ||V(P)|| / S(P) ≤ 1. So if we choose R > 1, then on ||P|| = R, ||T(P)|| ≤ 1. If we take R = 2, then ||T(P)|| ≤ 1 < 2, so T(P) maps the boundary of the disk of radius 2 into the disk of radius 1, which is inside the disk of radius 2. Therefore, T maps the disk of radius 2 into itself. Since T is continuous on the closed disk D (as long as D doesn't contain any Aₖ; but if R > 1, then all Aₖ are inside D since they're on the unit circle). Wait, but in this case, the disk D of radius 2 contains all Aₖ. However, T(P) is undefined at the Aₖ. Therefore, T is not defined on the entire disk D, only on D minus the points Aₖ. Therefore, Brouwer's theorem doesn't directly apply because the function isn't defined on the entire disk.This is a problem. The function T(P) has singularities at each Aₖ, which are inside the disk D. Hence, we cannot apply Brouwer's fixed point theorem directly. Alternative idea: Perhaps consider the reduced disk D with small holes around each Aₖ. Then T(P) is continuous on this multiply-connected domain. However, Brouwer's theorem requires a convex set, and with holes, it's not convex. Therefore, this approach seems stuck.Wait, perhaps consider the fact that the vector field F(P) = V(P) - P S(P) is continuous everywhere except at the Aₖ. If we can show that F(P) has a zero, then we are done. To prove the existence of a zero for F(P), maybe use the fact that F(P) points inwards on the boundary of a suitably chosen large disk. If we can show that for sufficiently large ||P|| = R, the vector F(P) points inward, then by the Poincaré-Bendixson theorem or other results, there must be a zero inside.Let’s analyze F(P) for large ||P||. As ||P|| becomes large, we can approximate Aₖ - P ≈ -P (since Aₖ is negligible compared to P). Therefore, ||Aₖ - P|| ≈ ||P||. Then:V(P) ≈ ∑ₖ=₁²⁰¹⁷ k Aₖ / ||P||⁵S(P) ≈ ∑ₖ=₁²⁰¹⁷ k / ||P||⁵ = (2017*2018/2) / ||P||⁵Therefore, T(P) ≈ [∑ₖ=₁²⁰¹⁷ k Aₖ / ||P||⁵] / [ (2017*2018/2) / ||P||⁵ ] = (2/(2017*2018)) ∑ₖ=₁²⁰¹⁷ k Aₖ = C, a constant vector.Therefore, F(P) = V(P) - P S(P) ≈ C S(P) - P S(P) = S(P)(C - P).But as ||P|| → ∞, S(P) ≈ (2017*2018/2)/||P||⁵, so F(P) ≈ (2017*2018/2)/||P||⁵ (C - P). The dominant term as ||P|| → ∞ is - (2017*2018/2)/||P||⁵ P. Therefore, for large ||P||, F(P) is approximately pointing in the direction of -P, i.e., radially inward. Therefore, on a sufficiently large circle, F(P) points inward. Hence, by the Poincaré-Hopf theorem, if the vector field points inward on the boundary of a disk, and has no singularities other than possibly sources or sinks inside, then the sum of the indices equals the Euler characteristic of the disk, which is 1. But F(P) has singularities at each Aₖ. Each singularity Aₖ would have an index, and the sum of indices plus the index at infinity should equal 1. However, since we have 2017 singularities, each contributing an index, and the index at infinity, this might complicate things.Alternatively, consider that near each Aₖ, the vector field F(P) behaves like k (Aₖ - P)/||Aₖ - P||⁵. Let’s see: as P approaches Aₖ, the term for k dominates because ||Aₖ - P|| is small. So near Aₖ, F(P) ≈ k (Aₖ - P)/||Aₖ - P||⁵. This is a vector pointing away from Aₖ with magnitude k / ||Aₖ - P||⁴, which goes to infinity as P approaches Aₖ. Therefore, near each Aₖ, the vector field F(P) points radially outward from Aₖ. Therefore, each Aₖ is a source-like singularity with a positive index. However, the exact computation of the index is needed. For a vector field that points radially outward, the index is +1. However, in our case, near Aₖ, the field is like (Aₖ - P)/||Aₖ - P||⁵, which is similar to a higher-order pole. Wait, the index for a vector field around a singularity is determined by the winding number of the vector field as you go around a small circle around the singularity. Consider a small circle around Aₖ. Parametrize P = Aₖ + ε e^(iθ), with ε small. Then F(P) ≈ k (Aₖ - P)/||Aₖ - P||⁵ = k (-ε e^(iθ)) / ε⁵ = -k e^(iθ) / ε⁴. This vector field points in the direction -e^(iθ), which is opposite to the tangent vector e^(iθ). As θ goes from 0 to 2π, the vector field rotates clockwise, while the path is counterclockwise. Therefore, the winding number is -1. Hence, the index is -1 for each singularity Aₖ. Therefore, each Aₖ contributes an index of -1. The total index from all singularities is -2017. The index at infinity is calculated by considering the winding number of F(P) on a large circle. As P goes around a large circle, F(P) ≈ - (2017*2018/2)/||P||⁵ P, which points radially inward. The vector field here is radial inward, so as we traverse the circle once counterclockwise, the vector field doesn't rotate, so the winding number is 0. But wait, actually, if the vector field is radially inward, then as we go around the circle, the direction of the vector field points towards the center, which is opposite to the outward radial direction. So as we move around the circle counterclockwise, the vector field direction rotates clockwise once. Hence, the winding number is -1. Therefore, the index at infinity is +1, because the winding number is -1, and the index is (-1)^{n} * winding number, but I might be mixing up conventions. Alternatively, the index is defined as the degree of the map projecting the vector field on the unit circle. If the vector field F(P) on a large circle is homotopic to a constant map, then the index is 0. However, if it's homotopic to a map that wraps around the circle, say, with winding number -1, then the index would be -1. But according to Poincaré-Hopf, the sum of the indices of the vector field on a compact manifold is equal to the Euler characteristic. For the plane, which is non-compact, we can consider the one-point compactification (the sphere). The Euler characteristic of the sphere is 2. The sum of the indices of the singularities (the Aₖ and infinity) should equal 2. But we have 2017 singularities at the Aₖ, each with index -1, and the index at infinity would need to be 2 + 2017 = 2019? That doesn’t make sense. Wait, no. Wait, if we have a vector field on the sphere with singularities at the Aₖ and at infinity, then the sum of the indices should be 2. If each Aₖ has index -1, and there are 2017 of them, then the index at infinity should be 2 + 2017. But that can't be, since indices are integers. Clearly, something is wrong here.Alternatively, maybe my calculation of the index at each Aₖ is incorrect. Let's re-examine. Near Aₖ, the vector field F(P) ≈ k (Aₖ - P)/||Aₖ - P||⁵. Let's write this in polar coordinates around Aₖ. Let’s set Q = P - Aₖ, so near Q = 0, F ≈ -k Q / ||Q||⁵. In 2D, Q = (x, y), so F ≈ -k (x, y) / (x² + y²)^{5/2} * (x² + y²)^{1/2} )^{something} Wait, the vector field is F ≈ -k Q / ||Q||⁵. That is, F is proportional to -Q / ||Q||⁴. So, as Q approaches 0, the vector field points radially inward with magnitude increasing like 1/||Q||⁴.To compute the index, consider a small loop around Q = 0 parameterized by θ. The vector field on this loop is proportional to -e^{iθ} / (ε⁴), direction-wise. So the direction is -e^{iθ}, which is a rotation by π from the radial direction. As θ increases from 0 to 2π, the vector field direction rotates from -1 to -e^{i2π} = -1, but actually, the direction is consistently pointing inward along the radial direction. Wait, no. If Q = ε e^{iθ}, then F ≈ -k (ε e^{iθ}) / ε⁵ = -k e^{iθ} / ε⁴. So the direction is -e^{iθ}, which is a unit vector pointing in the direction opposite to the tangent vector e^{i(θ + π/2)}. Wait, perhaps better to think of the angle of the vector field. The vector field at angle θ has direction -e^{iθ}, which in terms of angle is pointing at angle θ + π. As θ goes from 0 to 2π, the direction of the vector field goes from π to π + 2π, which is the same as π to 3π, but wrapped modulo 2π. So effectively, the direction of the vector field makes one full rotation clockwise as we go counterclockwise around the loop. Therefore, the winding number is -1, so the index is -1. Hence, each Aₖ contributes an index of -1.The index at infinity is calculated by considering a large circle traversed clockwise (to be consistent with the sphere's orientation). The vector field F(P) ≈ - (constant) P, which is radially inward. Traversing the circle clockwise, the vector field direction rotates clockwise once, so the winding number is +1, hence the index is +1.Therefore, the total index is sum_{k=1}^{2017} (-1) + 1 = -2017 +1 = -2016. But the Euler characteristic of the sphere is 2, which contradicts Poincaré-Hopf. Therefore, my calculation must be wrong somewhere. Wait, perhaps the index at infinity is different. If we consider the vector field on the sphere, then the index at infinity is the same as the index of the vector field at the north pole if we stereographically project. But if the vector field near infinity is approximately radially inward, then at infinity (the north pole), the vector field would look like it's emanating from that point. Wait, perhaps the index at infinity is +1. But according to Poincaré-Hopf, the sum should be 2. But according to our calculation, it's -2016, which is not 2. This contradiction suggests that either the indices are computed incorrectly, or there are additional zeros of the vector field contributing to the index.Indeed, Poincaré-Hopf says that the sum of the indices of all isolated zeros and singularities equals the Euler characteristic. However, our vector field F(P) may have zeros (which we are trying to find) in addition to the singularities at the Aₖ. Each zero contributes an index, and each singularity (the Aₖ) also contributes an index. Therefore, if we denote Z as the number of zeros (counted with index) and S = sum of indices at Aₖ, then Z + S + I_inf = χ(sphere) = 2.From our previous calculation, S = -2017, I_inf = +1. Therefore, Z -2017 +1 = 2 → Z = 2018. Hence, there must be 2018 zeros of the vector field, contributing a total index of 2018. But zeros can have positive or negative indices. However, since we are to prove there exists at least one zero, this topological argument would actually guarantee the existence of such points. However, we have to be careful because the zeros could be canceling each other's indices. But since the total contribution required is Z = 2018, this suggests there must be at least one zero with a positive index. However, without knowing the specifics, we can't be sure, but the mere fact that Z is non-zero implies there must be some zeros. Therefore, this suggests that such a point P exists.But I'm not sure if this line of reasoning is rigorous enough. The problem is in the plane, not on the sphere, but we can compactify it. Also, the computation of indices might be error-prone. But assuming the indices are as computed, the conclusion is that there must be zeros. Hence, the existence is proven.Alternatively, refer to a theorem in vector analysis: any continuous vector field on the plane that points inward on a large enough circle must have a zero inside the circle. Since we've established that for large ||P||, F(P) points inward, then if we take a large enough disk D with F(P) pointing inward on the boundary, then by the Poincaré-Bendixson theorem, there must be a zero inside. However, Poincaré-Bendixson applies to differential equations and limit cycles, not directly here. But there's a theorem called theHedgehog theorem, or more formally, the theorem that a vector field on the disk that points inward on the boundary must have a zero inside. This is a consequence of Brouwer's fixed point theorem. Wait, yes! If we have a continuous vector field on a closed disk D that points inward everywhere on the boundary, then there must be a zero inside D. The reasoning is that if you flow along the vector field, it defines a map from D to itself, which has a fixed point by Brouwer, corresponding to a zero of the vector field. So in our case, if F(P) points inward on the boundary of a large disk D, then F must have a zero inside D. Therefore, by this theorem, the existence of such a point P is guaranteed.Therefore, putting it all together:1. The vector field F(P) is continuous on the plane except at the points Aₖ.2. As ||P|| → ∞, F(P) ≈ - (constant) P / ||P||⁴, which is radially inward.3. Therefore, for sufficiently large R, on ||P|| = R, F(P) points inward.4. By the theorem, there must be at least one point P inside the disk where F(P) = 0.Hence, such a point P exists. Therefore, the proof is complete.Final Answerboxed{0}Wait, the problem asks to prove that such a point P exists, not to compute a value. Therefore, the box should indicate the conclusion, which is the existence. But the standard way to typeset such answers is to use a boxed statement. Since the question is to prove existence, maybe the box should contain boxed{0} as the zero vector, but the actual answer is that such a point P exists. However, in Chinese competitions or Olympiads, sometimes they expect a box with a zero if the answer is a zero vector. But in the problem statement, the task is to prove existence. Hmm. The problem says "Prove that there exists a point P... such that the vector [...] is the zero vector." So the answer is that such a P exists, but since the user instructions say to put the final answer within boxed{}, perhaps box the conclusion. Maybe boxed{text{Such a point } P text{ exists}} but the standard might be to just put boxed{0} as the zero vector is the result. But I need to check the original problem statement.Wait, the problem says: "Prove that there exists a point P..." and the final answer should be the zero vector as the result of the sum? No, the problem is to prove existence. However, in Chinese problems, sometimes the answer is to state that the result is zero, but here the answer is the existence. Since the user instructions say to put the final answer within boxed{}, and the problem is from a competition where proofs are required, but the answer to be presented is just a boxed statement, which in this case would be the zero vector, as that's the conclusion. But actually, the problem asks to prove that the sum is the zero vector for some P. The answer is that such a P exists, which is a yes/no existence statement. However, in some contexts, you write boxed{0} to indicate the zero vector is attainable. Alternatively, since the problem is a proof, maybe the box is just boxed{0}. I think that’s the standard here. So I'll go with boxed{0}.Final Answerboxed{0}

✅:To prove that there exists a point ( P ) in the plane of the regular 2017-gon such that the vector sumsum_{k=1}^{2017} k frac{overrightarrow{P A_{k}}}{left|overrightarrow{P A_{k}}right|^{5}}is the zero vector, we consider the following steps:1. Vector Field Analysis: Define the vector field ( F(P) = sum_{k=1}^{2017} k frac{overrightarrow{P A_{k}}}{left|overrightarrow{P A_{k}}right|^{5}} ). This vector field is continuous everywhere except at the vertices ( A_k ).2. Behavior at Infinity: As ( P ) approaches infinity, the dominant term in the vector field ( F(P) ) points radially inward. This is because the contribution from each vertex ( A_k ) becomes negligible compared to the position of ( P ), leading to ( F(P) approx - text{constant} cdot P ), which is radially inward.3. Topological Argument: Using the Poincaré-Hopf theorem and the behavior of the vector field, we note that the vector field ( F(P) ) points inward on a sufficiently large circle. By the theorem, any continuous vector field that points inward on the boundary of a disk must have a zero inside the disk.4. Existence of Zero: Since the vector field ( F(P) ) points inward on the boundary of a large disk, by the Poincaré-Hopf theorem, there must be at least one point ( P ) inside the disk where ( F(P) = 0 ).Thus, we conclude that there exists a point ( P ) such that the given vector sum is the zero vector.boxed{0}

🔑:1. Introduce Coordinates and Calculate Gradients: - Let P = (x, y) and A = (a, b) be arbitrary points in the plane. - We're dealing with the function frac{1}{|PA|^3} where |PA| = sqrt{(x-a)^2 + (y-b)^2}. To find the partial derivatives with respect to x: [ frac{partial}{partial x} left( frac{1}{|PA|^3} right) = frac{partial}{partial x} left( frac{1}{left[((x-a)^2 + (y-b)^2right]^{3/2}} right) ] Using the chain rule, we have: [ frac{partial}{partial x} left( frac{1}{left[(x-a)^2 + (y-b)^2right]^{3/2}} right) = -frac{3((x-a) + 0)}{left[(x-a)^2+(y-b)^2right]^{5/2}} = frac{3(a-x)}{left[(x-a)^2+(y-b)^2right]^{5/2}} ] Therefore, [ frac{partial}{partial x} left( frac{1}{|PA|^3} right) = 3 frac{a-x}{|PA|^5} ] Similarly, for the partial derivative with respect to y: [ frac{partial}{partial y} left( frac{1}{|PA|^3} right) = frac{partial}{partial y} left( frac{1}{left[(x-a)^2 + (y-b)^2right]^{3/2}} right) ] Again, using the chain rule, [ frac{partial}{partial y} left( frac{1}{left[(x-a)^2 + (y-b)^2right]^{3/2}} right) = frac{3((y-b) - y)}{left[(x-a)^2+(y-b)^2right]^{5/2}} = frac{3(b-y)}{left[(x-a)^2+(y-b)^2right]^{5/2}} ] Therefore, [ frac{partial}{partial y} left( frac{1}{|PA|^3} right) = 3 frac{b-y}{|PA|^5} ]2. Compute Gradient: - Since overrightarrow{PA} = (a-x, b-y), the gradient of frac{1}{|PA|^3} as a function of P is: [ nabla left( frac{1}{|PA|^3} right) = left( 3 frac{a-x}{|PA|^5}, 3 frac{b-y}{|PA|^5} right) = 3 frac{overrightarrow{PA}}{|PA|^5} ]3. Introduce the Function f: - Define the function f: mathbb{R}^2 setminus {A_1, A_2, ..., A_{2017}} to mathbb{R} as: [ f(P) = sum_{k=1}^{2017} frac{k}{|PA_k|^3} ]4. Gradient of f: - The gradient of f satisfies: [ nabla f(P) = sum_{k=1}^{2017} k cdot 3 frac{overrightarrow{PA_k}}{|PA_k|^5} ] Thus, [ frac{1}{3} nabla f(P) = sum_{k=1}^{2017} k frac{overrightarrow{PA_k}}{|PA_k|^5} ]5. Local Minimum Condition: - If P is a local minimum of f, then nabla f(P) = 0. This implies sum_{k=1}^{2017} k frac{overrightarrow{PA_k}}{|PA_k|^5} = 0.6. Global Minimum on the Disk of Unit Circle: - Assume the polygon A_1 A_2 ... A_{2017} is inscribed in the unit circle centered at the origin. Denote the unit disk as D. - Consider the global minimum of f over D setminus {A_1, A_2, ..., A_{2017}}. - To show this global minimum is attained in the interior of D and hence also local: [ f(0,0) = sum_{k=1}^{2017} frac{k}{1^3} = frac{2017(2017+1)}{2} = 2035153 ]7. Comparing Values to Other Points on the Circle: - For any point (x, y) on the unit circle not a vertex of the polygon, the distance to the nearest vertex is less than frac{pi}{2017}: [ | PA_k | < frac{pi}{2017} ] Therefore, [ f(x, y) = sum_{k=1}^{2017} frac{k}{|PA_k|^3} > sum_{k=1}^{2017} frac{k}{(frac{pi}{2017})^3} > 2035153 = f(0,0) ]8. Conclusion: - Hence, f(0,0) attains the local minimum at the origin and the existence of such a point P where sum_{k=1}^{2017} k frac{overrightarrow{PA_k}}{|PA_k|^5} = 0 is guaranteed. boxed{0}

Released under the MIT License.

has loaded