id
stringlengths
12
12
source
stringlengths
76
76
problem
stringlengths
59
2.04k
solutions
listlengths
0
11
USAMO-2009-3
https://artofproblemsolving.com/wiki/index.php/2009_USAMO_Problems/Problem_3
We define a chessboard polygon to be a polygon whose sides are situated along lines of the form \(x = a\) or \(y = b\), where \(a\) and \(b\) are integers. These lines divide the interior into unit squares, which are shaded alternately grey and white so that adjacent squares have different colors. To tile a chessboard polygon by dominoes is to exactly cover the polygon by non-overlapping \(1 \times 2\) rectangles. Finally, a tasteful tiling is one which avoids the two configurations of dominoes shown on the left below. Two tilings of a \(3 \times 4\) rectangle are shown; the first one is tasteful, while the second is not, due to the vertical dominoes in the upper right corner. \[ [asy] size(400); pathpen = linewidth(2.5); void chessboard(int a, int b, pair P){ for(int i = 0; i < a; ++i) for(int j = 0; j < b; ++j) if((i+j) % 2 == 1) fill(shift(P.x+i,P.y+j)*unitsquare,rgb(0.6,0.6,0.6)); D(P--P+(a,0)--P+(a,b)--P+(0,b)--cycle); } chessboard(2,2,(2.5,0));fill(unitsquare,rgb(0.6,0.6,0.6));fill(shift(1,1)*unitsquare,rgb(0.6,0.6,0.6)); chessboard(4,3,(6,0)); chessboard(4,3,(11,0)); MP("\mathrm{Distasteful\ tilings}",(2.25,3),fontsize(12)); /* draw lines */ D((0,0)--(2,0)--(2,2)--(0,2)--cycle); D((1,0)--(1,2)); D((2.5,1)--(4.5,1)); D((7,0)--(7,2)--(6,2)--(10,2)--(9,2)--(9,0)--(9,1)--(7,1)); D((8,2)--(8,3)); D((12,0)--(12,2)--(11,2)--(13,2)); D((13,1)--(15,1)--(14,1)--(14,3)); D((13,0)--(13,3)); [/asy] \] a) Prove that if a chessboard polygon can be tiled by dominoes, then it can be done so tastefully. b) Prove that such a tasteful tiling is unique.
[]
USAMO-2009-4
https://artofproblemsolving.com/wiki/index.php/2009_USAMO_Problems/Problem_4
For \(n \ge 2\) let \(a_1\), \(a_2\), ..., \(a_n\) be positive real numbers such that \[ (a_1+a_2+ ... +a_n)\left( {1 \over a_1} + {1 \over a_2} + ... +{1 \over a_n} \right) \le \left(n+ {1 \over 2} \right) ^2 \] Prove that \(\text{max}(a_1, a_2, ... ,a_n) \le 4 \text{min}(a_1, a_2, ... , a_n)\).
[ "Assume without loss of generality that \\(a_1 \\geq a_2 \\geq \\cdots \\geq a_n\\). Now we seek to prove that \\(a_1 \\le 4a_n\\).\n\nBy the Cauchy-Schwarz Inequality,\n\n\\[\n\\begin{align*} (a_n+a_2+ a_3 + ... +a_{n-1}+a_1)\\left({1 \\over a_1} + {1 \\over a_2} + ... +{1 \\over a_n}\\right) &\\ge \\left( \\sqrt{a_n \\over a_1} + n-2 + \\sqrt{a_1 \\over a_n} \\right)^2 \\\\ (n+ {1 \\over 2})^2 &\\ge \\left( \\sqrt{a_n \\over a_1} + n-2 + \\sqrt{a_1 \\over a_n} \\right)^2 \\\\ n+ {1 \\over 2} &\\ge n-2 + \\sqrt{a_n \\over a_1} + \\sqrt{a_1 \\over a_n} \\\\ {5 \\over 2} &\\ge \\sqrt{a_n \\over a_1} + \\sqrt{a_1 \\over a_n} \\\\ {17 \\over 4} &\\ge {a_n \\over a_1} + {a_1 \\over a_n} \\\\ 0 &\\ge (a_1 - 4a_n)\\left(a_1 - {a_n \\over 4}\\right) \\end{align*}\n\\]\n\nSince \\(a_1 \\ge a_n\\), clearly \\((a_1 - {a_n \\over 4}) > 0\\), dividing yields:\n\n\\[\n0 \\ge (a_1 - 4a_n) \\Longrightarrow 4a_n \\ge a_1\n\\]\n\nas desired.", "Assume without loss of generality that \\(a_1 \\geq a_2 \\geq \\cdots \\geq a_n\\). Using the Cauchy–Bunyakovsky–Schwarz inequality and the inequality given,\n\n\\[\n\\begin{align*} &\\left(a_1+a_2 + ... +a_n + 3a_n -\\frac{3a_1}{4}\\right)\\left({1 \\over a_1} + {1 \\over a_2} + ... +{1 \\over a_n}\\right) \\\\ =&\\left(\\frac{a_1}{4}+a_2 + ... +a_{n-1}+4a_n\\right)\\left({1 \\over a_1} + {1 \\over a_2} + ... +{1 \\over a_n}\\right) \\\\ \\ge& \\left(\\frac{1}{2}+n-2+2\\right)^2 \\\\ =&\\left(n+\\frac{1}{2}\\right)^2 \\\\ \\ge& (a_1+a_2 + ... +a_{n})\\left({1 \\over a_1} + {1 \\over a_2} + ... +{1 \\over a_n}\\right).\\end{align*}\n\\]\n\n(Note that \\(n-2 \\ge 0\\) since \\(n \\ge 2\\) as given!) This implies that \\(3a_n -\\frac{3a_1}{4} \\ge 0 \\iff 4a_n \\ge a_1\\) as desired.\n\n~Deng Tianle, username: Leole" ]
USAMO-2009-5
https://artofproblemsolving.com/wiki/index.php/2009_USAMO_Problems/Problem_5
Trapezoid \(ABCD\), with \(\overline{AB}||\overline{CD}\), is inscribed in circle \(\omega\) and point \(G\) lies inside triangle \(BCD\). Rays \(AG\) and \(BG\) meet \(\omega\) again at points \(P\) and \(Q\), respectively. Let the line through \(G\) parallel to \(\overline{AB}\) intersect \(\overline{BD}\) and \(\overline{BC}\) at points \(R\) and \(S\), respectively. Prove that quadrilateral \(PQRS\) is cyclic if and only if \(\overline{BG}\) bisects \(\angle CBD\).
[ "We will use directed angles in this solution. Extend \\(QR\\) to \\(T\\) as follows:\n\n\\[\n[asy] import cse5; import graph; import olympiad; dotfactor = 3; unitsize(1.5inch); path circle = Circle(origin, 1); draw(circle); pair A = (-.6, .8), B = (.6, .8), C = (.9, -sqrt(.19)), D = (-.9, -sqrt(.19)), G = bisectorpoint(C, B, D); draw(A--B--C--D--cycle); draw(B--D); dot(\"$A$\", A, NW); dot(\"$B$\", B, NE); dot(\"$C$\", C, SE); dot(\"$D$\", D, SW); dot(\"$G$\", G, dir(40)); pair P = IP(L(A, G, 10, 10), circle, 1), Q = IP(L(B, G, 10, 10), circle, 1); draw(A--P--C); draw(B--Q); dot(\"$P$\", P, SE); dot(\"$Q$\", Q, S); pair R = IP((-1, G.y)--(1, G.y), B--D), S = IP((-1, G.y)--(1, G.y), B--C); draw(P--Q--R--S--cycle); dot(\"$R$\", R, N); dot(\"$S$\", S, E); pair T = IP(L(Q, R, 10, 10), circle, 1); draw(R--T--C, dashed); draw(T--B, dashed); dot(\"$T$\", T, NW); [/asy]\n\\]\n\nIf:\n\nNote that\n\n\\[\n\\begin{align*}\\measuredangle GBT+\\measuredangle TRG&=\\frac{m\\widehat{TQ}}{2}+\\measuredangle TRB+\\measuredangle BRG\\\\ &=\\frac{m\\widehat{TQ}+m\\widehat{DQ}+m\\widehat{CB}+m\\widehat{BT}}{2}.\\\\ \\end{align*}\n\\]\n\nThus, \\(BTRG\\) is cyclic.\n\nAlso, note that \\(GSCP\\) is cyclic because\n\n\\[\n\\begin{align*}\\measuredangle CSG+\\measuredangle GPC&=\\measuredangle CBA+\\measuredangle APC\\\\ &=180^\\circ\\text{ or }0^\\circ, \\end{align*}\n\\]\n\ndepending on the configuration.\n\nNext, we have \\(T, G, C\\) are collinear since\n\n\\[\n\\measuredangle GTR=\\measuredangle GBR=\\frac{m\\widehat{DQ}}{2}=\\frac{m\\widehat{QC}}{2}=\\measuredangle CTQ.\n\\]\n\nTherefore,\n\n\\[\n\\begin{align*}\\measuredangle RQP+\\measuredangle PSR&=\\frac{m\\widehat{PBT}}{2}+\\measuredangle PCG\\\\ &=\\frac{m\\widehat{PBT}+m\\widehat{TDP}}{2}\\\\ &=180^\\circ \\end{align*},\n\\]\n\nso \\(PQRS\\) is cyclic.\n\nOnly If: These steps can be reversed.", "Extend \\(QR\\) to \\(T\\), and let line \\(l \\parallel AB\\) intersect \\(\\omega\\) at \\(K\\) and another point \\(V\\), as shown:\n\n\\[\n[asy] import cse5; import graph; import olympiad; dotfactor = 3; unitsize(1.5inch); path circle = Circle(origin, 1); draw(circle); pair A = (-.6, .8), B = (.6, .8), C = (.9, -sqrt(.19)), D = (-.9, -sqrt(.19)), G = bisectorpoint(C, B, D); draw(A--B--C--D--cycle); draw(B--D); dot(\"$A$\", A, NW); dot(\"$B$\", B, NE); dot(\"$C$\", C, SE); dot(\"$D$\", D, SW); dot(\"$G$\", G, dir(40)); pair P = IP(L(A, G, 10, 10), circle, 1), Q = IP(L(B, G, 10, 10), circle, 1); draw(A--P); draw(B--Q); dot(\"$P$\", P, SE); dot(\"$Q$\", Q, S); pair R = IP((-1, G.y)--(1, G.y), B--D), S = IP((-1, G.y)--(1, G.y), B--C); draw(P--Q--R--S); dot(\"$R$\", R, N); dot(\"$S$\", S, E); pair T = IP(L(Q, R, 10, 10), circle, 1); draw(Q--T); pair V = IP(L(P, S, 10, 10), circle, 1); draw(T--V); draw(P--V, dotted); dot(\"$T$\", T, NW); dot(\"$V$\", V, NE); [/asy]\n\\]\n\nIf:\n\nSuppose that \\(VP \\cap CB = S'\\), and \\(AC \\cap QV = R'\\). Pascal's theorem on the tuple \\((V, P, A, C, B, Q)\\) implies that the points \\(S'\\), \\(R'\\), and \\(G = PA \\cap BQ\\) are collinear. However, \\(AC\\) and \\(BD\\) are symmetrical with respect to the axis of symmetry of trapezoid \\(ABCD\\), and \\(TQ\\) and \\(VQ\\) are also symmetrical with respect to the axis of symmetry of \\(ABCD\\) (as \\(Q\\) is the midpoint of \\(\\overset{\\frown}{DC}\\), and \\(TV \\parallel DC\\)). Since \\(R = BD \\cap TQ\\), \\(R\\) and \\(R'\\) are symmetric with respect to the axis of symmetry of trapezoid \\(ABCD\\). This implies that line \\(R'G\\) is equivalent to line \\(RG\\). Thus, \\(S'\\) lies on line \\(RG\\). However, \\(S = BC \\cap RG\\), so this implies that \\(S' = S\\).\n\nNow note that \\(TVPQ\\) is cyclic. Since \\(TV \\parallel RS\\), \\(\\measuredangle VTQ = \\measuredangle SRQ\\). However, \\(\\measuredangle VTQ + \\measuredangle VPQ = 180^{\\circ} = \\measuredangle SRQ + \\measuredangle SPQ\\). Therefore, \\(PQRS\\) is cyclic.\n\nOnly If:\n\nConsider the same setup, except \\(Q\\) is no longer the midpoint of \\(\\overset{\\frown}{DC}\\). Note that \\(TV\\) must be parallel to \\(RG\\) in order for \\(PQRS\\) to be cyclic. We claim that \\(S' = S\\) and hope to reach a contradiction. Pascal's theorem on the tuple \\((V, P, A, C, B, Q)\\) implies that \\(S'\\), \\(R'\\), and \\(G = PA \\cap BQ\\) are collinear. However, there exists a unique point \\(Q\\) such that \\(HQ\\), \\(AC\\), and \\(RG\\) are concurrent. By If, \\(Q\\) must be the midpoint of \\(\\overset{\\frown}{DC}\\) in order for the concurrency to occur; hence, \\(R' \\notin RS\\). Then \\(R'G \\cap BC = S' \\neq S\\), since \\(RG \\cap BC = S\\). However, this is a contradiction, so therefore \\(TV\\) cannot be parallel to \\(RG\\) and \\(PQRS\\) is not cyclic.\n\nSolution by TheBoomBox77" ]
USAMO-2009-6
https://artofproblemsolving.com/wiki/index.php/2009_USAMO_Problems/Problem_6
Let \(s_1, s_2, s_3, \ldots\) be an infinite, nonconstant sequence of rational numbers, meaning it is not the case that \(s_1 = s_2 = s_3 = \cdots.\) Suppose that \(t_1, t_2, t_3, \ldots\) is also an infinite, nonconstant sequence of rational numbers with the property that \((s_i - s_j)(t_i - t_j)\) is an integer for all \(i\) and \(j\). Prove that there exists a rational number \(r\) such that \((s_i - s_j)r\) and \((t_i - t_j)/r\) are integers for all \(i\) and \(j\).
[ "Suppose the \\(s_i\\) can be represented as \\(\\frac{a_i}{b_i}\\) for every \\(i\\), and suppose \\(t_i\\) can be represented as \\(\\frac{c_i}{d_i}\\). Let's start with only the first two terms in the two sequences, \\(s_1\\) and \\(s_2\\) for sequence \\(s\\) and \\(t_1\\) and \\(t_2\\) for sequence \\(t\\). Then by the conditions of the problem, we have \\((s_2 - s_1)(t_2 - t_1)\\) is an integer, or \\((\\frac{a_2}{b_2} - \\frac{a_1}{b_1})(\\frac{c_2}{d_2} - \\frac{c_1}{d_1})\\) is an integer. Now we can set \\(r = \\frac{b_1 b_2}{d_1 d_2}\\), because the least common denominator of \\(s_2 - s_1\\) is \\(b_1 b_2\\) and of \\(t_2 - t_1\\) is \\(d_1 d_2\\), and multiplying or dividing appropriately by \\(\\frac{b_1 b_2}{d_1 d_2}\\) will always give an integer.\n\nNow suppose we kept adding \\(s_i\\) and \\(t_i\\) until we get to \\(s_m = \\frac{a_m}{b_m}\\) in sequence \\(s\\) and \\(t_m = \\frac{c_m}{d_m}\\) in sequence \\(t\\) so that \\((t_m - t_i)(s_m - s_i)\\) is an integer for all \\(i\\) with \\(1 \\le i < m\\), where \\(m\\) is a positive integer. At this point, we will have \\(r\\) = \\(\\frac{\\prod_{n=1}^{m}b_n}{\\prod_{n=1}^{m}d_n}\\), because these are the least common denominators of the two sequences up to \\(m\\). As we keep adding \\(s_i\\) and \\(t_i\\), \\(r\\) will always have value \\(\\frac{\\prod_{n=1}^{m}b_n}{\\prod_{n=1}^{m}d_n}\\), and we are done." ]
USAMO-2010-1
https://artofproblemsolving.com/wiki/index.php/2010_USAMO_Problems/Problem_1
Let \(AXYZB\) be a convex pentagon inscribed in a semicircle of diameter \(AB\). Denote by \(P, Q, R, S\) the feet of the perpendiculars from \(Y\) onto lines \(AX, BX, AZ, BZ\), respectively. Prove that the acute angle formed by lines \(PQ\) and \(RS\) is half the size of \(\angle XOZ\), where \(O\) is the midpoint of segment \(AB\).
[ "Angles are directed mod 180 {don't worry}(take so that the solution can be applied to any configuration) Let the intersection of pq and rs is M.\n\n\\(m\\angle ARM=m\\angle ZRS =m \\angle ZYS =m\\angle YZR = m\\angle YZA = m \\angle YXP = m \\angle XYQ =m \\angle XPQ = m \\angle APM\\) This implies APRM is cyclic So: \\(m \\angle PMR = m \\angle PAR =m \\angle XAZ = \\frac{m\\angle XOZ}{2}\\) and we are done. For any concern mail on harshsahu1098@gmail.com", "Let \\(\\alpha = \\angle BAZ\\), \\(\\beta = \\angle ABX\\). Since \\(XY\\) is a chord of the circle with diameter \\(AB\\), \\(\\angle XAY = \\angle XBY = \\gamma\\). From the chord \\(YZ\\), we conclude \\(\\angle YAZ = \\angle YBZ = \\delta\\).\n\n\\[\n[asy] import olympiad; // Scale unitsize(1inch); real r = 1.75; // Semi-circle: centre O, radius r, diameter A--B. pair O = (0,0); dot(O); label(\"$O$\", O, plain.S); pair A = r * plain.W; dot(A); label(\"$A$\", A, unit(A)); pair B = r * plain.E; dot(B); label(\"$B$\", B, unit(B)); draw(arc(O, r, 0, 180)--cycle); // points X, Y, Z real alpha = 22.5; real beta = 15; real delta = 30; pair X = r * dir(180 - 2*beta); dot(X); label(\"$X$\", X, unit(X)); pair Y = r * dir(2*(alpha + delta)); dot(Y); label(\"$Y$\", Y, unit(Y)); pair Z = r * dir(2*alpha); dot(Z); label(\"$Z$\", Z, unit(Z)); // Feet of perpendiculars from Y pair P = foot(Y, A, X); dot(P); label(\"$P$\", P, unit(P-Y)); dot(P); pair Q = foot(Y, B, X); dot(P); label(\"$Q$\", Q, unit(A-Q)); dot(Q); pair R = foot(Y, B, Z); dot(R); label(\"$S$\", R, unit(R-Y)); dot(R); pair S = foot(Y, A, Z); dot(S); label(\"$R$\", S, unit(B-S)); dot(S); pair T = foot(Y, A, B); dot(T); label(\"$T$\", T, unit(T-Y)); dot(T); // Segments draw(B--X); draw(B--Y); draw(B--R); draw(A--Z); draw(A--Y); draw(A--P); draw(Y--P); draw(Y--Q); draw(Y--R); draw(Y--S); draw(R--T); draw(P--T); // Right angles draw(rightanglemark(A, X, B, 3)); draw(rightanglemark(A, Y, B, 3)); draw(rightanglemark(A, Z, B, 3)); draw(rightanglemark(A, P, Y, 3)); draw(rightanglemark(Y, R, B, 3)); draw(rightanglemark(Y, S, A, 3)); draw(rightanglemark(B, Q, Y, 3)); // Acute angles import markers; void langle(pair A, pair B, pair C, string l=\"\", real r=40, int n=1, int nm = 0) { string sl = \"$\\scriptstyle{\" + l + \"}$\"; marker m = (nm > 0) ? marker(markinterval(stickframe(n=nm, 2mm), true)) : nomarker; markangle(Label(sl), radius=r, n=n, A, B, C, m); } langle(B, A, Z, \"\\alpha\" ); langle(X, B, A, \"\\beta\", n=2); langle(Y, A, X, \"\\gamma\", nm=1); langle(Y, B, X, \"\\gamma\", nm=1); langle(Z, A, Y, \"\\delta\", nm=2); langle(Z, B, Y, \"\\delta\", nm=2); langle(R, S, Y, \"\\alpha+\\delta\", r=23); langle(Y, Q, P, \"\\beta+\\gamma\", r=23); langle(R, T, P, \"\\chi\", r=15); [/asy]\n\\]\n\nTriangles \\(BQY\\) and \\(APY\\) are both right-triangles, and share the angle \\(\\gamma\\), therefore they are similar, and so the ratio \\(PY : YQ = AY : YB\\). Now by Thales' theorem the angles \\(\\angle AXB = \\angle AYB = \\angle AZB\\) are all right-angles. Also, \\(\\angle PYQ\\), being the fourth angle in a quadrilateral with 3 right-angles is again a right-angle. Therefore \\(\\triangle PYQ \\sim \\triangle AYB\\) and \\(\\angle YQP = \\angle YBA = \\gamma + \\beta\\). Similarly, \\(RY : YS = AY : YB\\), and so \\(\\angle YRS = \\angle YAB = \\alpha + \\delta\\).\n\nNow \\(RY\\) is perpendicular to \\(AZ\\) so the direction \\(RY\\) is \\(\\alpha\\) counterclockwise from the vertical, and since \\(\\angle YRS = \\alpha + \\delta\\) we see that \\(SR\\) is \\(\\delta\\) clockwise from the vertical. (Draw an actual vertical line segment if necessary.)\n\nSimilarly, \\(QY\\) is perpendicular to \\(BX\\) so the direction \\(QY\\) is \\(\\beta\\) clockwise from the vertical, and since \\(\\angle YQP\\) is \\(\\gamma + \\beta\\) we see that \\(QY\\) is \\(\\gamma\\) counterclockwise from the vertical.\n\nTherefore the lines \\(PQ\\) and \\(RS\\) intersect at an angle \\(\\chi = \\gamma + \\delta\\). Now by the central angle theorem \\(2\\gamma = \\angle XOY\\) and \\(2\\delta = \\angle YOZ\\), and so \\(2(\\gamma + \\delta) = \\angle XOZ\\), and we are done.\n\nNote that \\(RTQY\\) is a quadrilateral whose angles sum to 360°; can you find a faster approach using this fact?\n\n## Footnote\n\nWe can prove a bit more. Namely, the extensions of the segments \\(RS\\) and \\(PQ\\) meet at a point on the diameter \\(AB\\) that is vertically below the point \\(Y\\).\n\nSince \\(YR = AY \\sin(\\delta)\\) and is inclined \\(\\alpha\\) counterclockwise from the vertical, the point \\(S\\) is \\(AY \\sin(\\delta) \\sin(\\alpha)\\) horizontally to the right of \\(Y\\).\n\nNow \\(AS = AY \\cos(\\delta)\\), so \\(S\\) is \\(AS \\sin(\\alpha) = AY \\cos(\\delta)\\sin(\\alpha)\\) vertically above the diameter \\(AB\\). Also, the segment \\(SR\\) is inclined \\(\\delta\\) clockwise from the vertical, so if we extend it down from \\(S\\) towards the diameter \\(AB\\) it will meet the diameter at a point which is \\(AY \\cos(\\delta)\\sin(\\alpha)\\tan(\\delta) = AY \\sin(\\delta)\\sin(\\alpha)\\) horizontally to the left of \\(S\\). This places the intersection point of \\(RS\\) and \\(AB\\) vertically below \\(Y\\).\n\nSimilarly, and by symmetry the intersection point of \\(PQ\\) and \\(AB\\) is directly below \\(Y\\) on \\(AB\\), so the lines through \\(PQ\\) and \\(RS\\) meet at a point \\(T\\) on the diameter that is vertically below \\(Y\\).\n\n## Footnote to the Footnote\n\nThe Footnote's claim is more easily proved as follows.\n\nNote that because \\(\\angle{QPY}\\) and \\(\\angle{YAB}\\) are both complementary to \\(\\beta + \\gamma\\), they must be equal. Now, let \\(PQ\\) intersect diameter \\(AB\\) at \\(T'\\). Then \\(PYT'A\\) is cyclic and so \\(\\angle{YT'A} = 180^\\circ - \\angle{APY} = 90^\\circ\\). Hence \\(T'YSB\\) is cyclic as well, and so we deduce that \\(\\angle{YST'} = \\angle{YBT'} = 90^\\circ - \\alpha - \\delta = \\angle{YSR}.\\) Hence \\(S, R, T'\\) are collinear and so \\(T = T'\\). This proves the Footnote.\n\n## Footnote to the Footnote to the Footnote\n\nThe Footnote's claim can be proved even more easily as follows.\n\nDrop an altitude from \\(Y\\) to \\(AB\\) at point \\(T\\). Notice that \\(P, Q, T\\) are collinear because they form the Simson line of \\(\\triangle AXB\\) from \\(Y\\). Also notice that \\(P, Q, T\\) are collinear because they form the Simson line of \\(\\triangle AZB\\) from \\(Y\\). Since \\(T\\) is at the diameter \\(AB\\), lines \\(PQ\\) and \\(SR\\) must intersect at the diameter.\n\n## Footnote to the Fourth Power\n\nThere is another, simpler solution using Simson lines. Can you find it?\n\n## Operation Diagram\n\nOf course, as with any geometry problem, DRAW A HUGE DIAGRAM spanning at least one page. And label all your right angles, noting rectangles \\(PXQY\\) and \\(YSZR\\). It looks like there are a couple of key angles we need to diagram. Let's take \\(\\angle{ZAB} = \\alpha, \\angle{XBA} = \\beta, \\angle{YAZ} = \\angle{YBZ} = \\delta\\). From there \\(\\angle{XOZ}=180^\\circ - \\angle{XOA}-\\angle{ZOB}=180-2(\\beta + \\alpha)\\).\n\nMove on to the part about the intersection of \\(PQ\\) and \\(RS\\). Call the intersection \\(J\\). Note that by Simson Lines from point \\(Y\\) to \\(\\triangle{ABX}\\) and \\(\\triangle{AZB}\\), \\(YJ\\) is perpendicular to \\(AB\\) and \\(J\\) lies on \\(AB\\). Immediately note that we are trying to show that \\(\\angle{PJS} = 90 - \\beta - \\alpha\\).\n\nIt suffices to show that referencing quadrilateral \\(QR~J\\), where \\(~\\) represents the intersection of \\(XB, AZ\\), we have reflex \\(\\angle{Q~R} + \\angle{BQJ} + \\angle{ARJ} = 270 + \\alpha + \\beta\\). Note that the reflex angle is \\(180^\\circ + \\angle{A~X} = 180^\\circ + (90^\\circ - \\angle{XA*}) = 270^\\circ - ((90 - \\beta) - \\alpha) = 180 ^\\circ + \\alpha + \\beta\\), therefore it suffices to show that \\(\\angle{BQJ} + \\angle{ARJ} = 90^\\circ\\). To make this proof more accessible, note that via (cyclic) rectangles \\(PXQY\\) and \\(YSZR\\), it suffices to prove \\(\\angle{YPJ} + \\angle{YSJ} = 90^\\circ\\).\n\nNote \\(\\angle{YPJ} = \\angle{YPQ} = \\angle{YXQ} = \\angle{YXB} = \\angle{YAB} = \\alpha + \\delta\\). Note \\(\\angle{YSJ} = \\angle{YSR} = \\angle{YZR} = \\angle{YZA} = \\angle{YBA} = \\angle{YBX} + \\angle{XBA} = ((90^\\circ - \\alpha) - \\delta - \\beta) + \\beta = 90^\\circ - \\alpha - \\delta\\), which completes the proof.\n\n## Footnote to Operation Diagram\n\nFor reference/feasibility records: took expiLnCalc ~56 minutes (consecutively). During the problem expiLnCalc realized that the inclusion of \\(\\delta\\) was necessary when trying to show that \\(\\angle{YSJ}+\\angle{YPJ}=90^\\circ\\). Don't be afraid to attempt several different strategies, and always be humble!", "\\[\n[asy] currentpicture=new picture; size(12cm); pair O, A, B, X, Y, Z, P, Q, R, SS, T; O=(0, 0); A=(-1, 0); B=(1, 0); X=(Cos(144), Sin(144)); Y=(Cos(105), Sin(105)); Z=(Cos(27), Sin(27)); P=foot(Y, A, X); Q=foot(Y, B, X); R=foot(Y, A, Z); SS=foot(Y, B, Z); T=foot(Y, A, B); dot(O); dot(A); dot(B); dot(X); dot(Y); dot(Z); dot(P); dot(Q); dot(R); dot(SS); dot(T); draw(arc(O, 1, 0, 180)); draw(circumcircle(T, A, Y), dotted); draw(circumcircle(T, B, Y), dotted); draw(A -- B); draw(Z -- O -- X -- A -- Z -- B -- X); draw(A -- Y -- B); draw(P -- T -- SS); draw(P -- Y -- Q); draw(R -- Y -- SS); draw(X -- P); draw(Z -- SS); draw(Y -- T); draw(rightanglemark(Y, T, B, 1.25)); draw(rightanglemark(Y, P, A, 1.25)); draw(rightanglemark(Y, Q, X, 1.25)); draw(rightanglemark(Y, R, Z, 1.25)); draw(rightanglemark(Y, SS, B, 1.25)); draw(rightanglemark(A, X, B, 1.25)); draw(rightanglemark(A, Y, B, 1.25)); draw(rightanglemark(A, Z, B, 1.25)); label(\"$O$\", O, S); label(\"$A$\", A, SW); label(\"$B$\", B, SE); label(\"$X$\", X, (X-B)/length(X-B)); label(\"$Y$\", Y, Y); label(\"$Z$\", Z, (Z-A)/length(Z-A)); label(\"$P$\", P, (P-T)/length(P-T)); label(\"$Q$\", Q, SW); label(\"$R$\", R, SE); label(\"$S$\", SS, (SS-T)/length(SS-T)); label(\"$T$\", T, S); [/asy]\n\\]\n\nLet \\(T\\) be the foot of the perpendicular from \\(Y\\) to \\(\\overline{AB}\\), let \\(O\\) be the center of the semi-circle.\n\nSince we have a semi-circle, if we were to reflect it over \\(\\overline {AB}\\), we would have a full circle, with \\(\\triangle{AXB}\\) and \\(\\triangle{AZB}\\) inscribed in it. Now, notice that \\(Y\\) is a point on that full circle, so we can say that \\(T\\) lies on the Simson Line \\(\\overline{PQ}\\) from \\(Y\\) to \\(\\triangle AXB\\) and that it also lies on the Simson line \\(\\overline {RS}\\) from \\(Y\\) to \\(\\triangle AZB\\). Thus, \\(T\\) lies on two distinct lines in a plane, which means that \\(T=\\overline{PQ}\\cap\\overline{RS}\\). Therefore, it suffices to show that \\(\\angle PTS=\\tfrac{1}{2}\\angle XOZ\\).\n\nSince \\(m\\angle YTA + m\\angle YPA = 90^\\circ + 90^\\circ = 180^\\circ\\) and \\(m \\angle YTB + m \\angle YSB = 90^\\circ + 90^\\circ = 180^\\circ\\), we know that \\(TAPY\\) and \\(TBSY\\) are cyclic quadrilaterals.\n\nWe use this fact to get\n\n\\[\n\\angle PTS=\\angle PTY+\\angle YTS=\\angle PAY+\\angle YBS=\\angle XAY+\\angle YBZ. \\space \\space (1)\n\\]\n\n\\\\ Now note that \\(\\angle XAY\\) is the inscribed angle of minor arc \\(\\overset{\\huge\\frown}{PY}\\), and \\(\\angle XOY\\) is the central angle of minor arc \\(\\overset{\\huge\\frown}{AB}\\), so \\(\\angle XAY = \\frac{\\overset{\\huge\\frown}{PY}}{2} = \\frac{\\angle XOY}{2}\\). Similarly, \\(\\angle YBZ = \\frac{\\overset{\\huge\\frown}{YZ}}{2}=\\frac{\\angle YOZ}{2}\\). Thus we can say\n\n\\[\n\\angle XAY + \\angle YBZ = \\frac{\\angle XOY}{2} + \\frac{\\angle YOZ}{2}=\\frac{\\angle XOY + \\angle YOZ}{2} = \\frac{\\angle XOZ}{2}. \\space \\space (2)\n\\]\n\nCombining statements \\((1)\\) and \\((2)\\), we can say that \\(\\angle PTS = \\frac{\\angle XOZ}{2}\\), as desired. \\(\\square\\)\n\n~thinker123" ]
USAMO-2010-2
https://artofproblemsolving.com/wiki/index.php/2010_USAMO_Problems/Problem_2
There are \(n\) students standing in a circle, one behind the other. The students have heights \(h_1 < h_2 < \ldots < h_n\). If a student with height \(h_k\) is standing directly behind a student with height \(h_{k-2}\) or less, the two students are permitted to switch places. Prove that it is not possible to make more than \(\binom{n}{3}\) such switches before reaching a position in which no further switches are possible.
[ "We adopt the usual convention that \\(\\binom{i}{j} = 0\\) unless \\(0 \\le j \\le i\\). With this, the binomial coefficients are defined for all integers via the recursion:\n\n\\[\n\\binom{0}{0} = 1, \\quad \\binom{n}{k} = \\binom{n-1}{k} + \\binom{n-1}{k-1}\n\\]\n\nIt is clear that the circle is oriented and all the students are facing in same direction (clockwise or counterclockwise). We'll call this direction forward.\n\nIn any switch consider the taller student to have moved forward and the shorter student to have remained stationary. No backward motion is allowed. With this definition of forward motion, the first two students with heights \\(h_1\\) and \\(h_2\\) are always stationary, while other students potentially move past them.\n\nFor \\(k > 2\\), the student with height \\(h_k\\) can never switch places with the student with height \\(h_{k-1}\\), and the former can make at most \\(k-2\\) more forward moves than the latter (when all the students of heights \\(h_1, \\ldots h_{k-2}\\) are between \\(h_k\\) and \\(h_{k-1}\\) in the forward direction).\n\nTherefore, if the \\((k-1)^{\\mathrm{st}}\\) student can make \\(s_{k-1}\\) forward steps, the \\(k^{\\mathrm{th}}\\) student can make at most \\(s_{k-1} + k - 2\\) steps. With \\(s_1 = s_2 = 0\\) and \\(s_3 = 0 + (3-2) = 1\\), and a constant second difference of \\(1\\), we quickly see that \\(s_k = \\binom{k-1}{2}\\).\n\nWith \\(n\\) students in all, the total number of steps is therefore at most \\(\\sum_{i=3}^{n}\\binom{i-1}{2} = \\binom{n}{3}\\), by the Hockey Stick Theorem", "WLOG, let \\(h_k=k\\). Now, we find the end arrangement with the most switches possible. We claim that the arrangement will be \\(n,n-1,n-2,\\dots ,1\\), where the left to right direction is the \"backwards\" direction. To prove this makes the most switches, we show that there is always at least one more switch that can be done for any other arrangement. This is elementary to show. There will always be one height \\(x\\) such that the number to its right is \\(2\\) less than \\(x\\), unless every number has \\(x-1\\) to the right of \\(x\\) (other than \\(2\\) and \\(1\\)). The exception occurs at our claim, so our claim is proven. Now we want to find the maximum ways we can \"undo\" our arrangement. But undoing a switch is just doing a switch from our arrangement in the opposite direction. So, the start arrangement with the most possible switches is the reverse of the end arrangement or \\(1,2,3,\\dots ,n\\). We want to find how many switches must be done to get from the start arrangement to the end arrangement. We start by switching \\(n\\) around until it cannot be switched anymore. We find that we can switch \\(n\\) \\(n-2\\) times. When we are switching (or, in other words, moving to the right) the number \\(k\\), we can switch it \\(k-2\\) times before it is to the left of \\(k-1\\). But then we can also switch each of \\(k+1,k+2,\\dots ,n\\) (in that order) \\(k-2\\) times before they get stopped again. Let \\(x=n-k+2\\). Then, the sum of the switches is \\(\\sum\\limits_{x=2}^{n-1} (n-x)(x-1)=\\sum\\limits_{x=1}^{n} (n-x)(x-1)\\).\n\nRearranging and summing, we get\n\n\\[\n\\sum\\limits_{x=1}^{n} x(n+1)-\\sum\\limits_{x=1}^{n} n -\\sum\\limits_{x=1}^{n} x^2=\\frac{n(n+1)^2}2-n^2-\\frac{n(n+1)(2n+1)}6=\\frac{3n(n^2+1)}6-\\frac{n(n+1)(2n+1)}6=\\frac{n(n^2-3n+2)}6=\\frac{n(n-1)(n-2)}{3!}=\\binom{n}{3}\n\\]\n\n## Comment\n\nThis process of switching is very similar to Bubble Sort, except that the terms must be in consecutive decreasing order, and wraps around \\(2,1,n,n-1\\dots\\). This answer seems to disagree, though, because the worst case efficiency is \\(\\mathcal{O}(n^2)\\), not \\(\\mathcal{O}(n^3)\\).\n\nhttps://en.wikipedia.org/wiki/Bubble_sort" ]
USAMO-2010-3
https://artofproblemsolving.com/wiki/index.php/2010_USAMO_Problems/Problem_3
The \(2010\) positive numbers \(a_1, a_2, \ldots , a_{2010}\) satisfy the inequality \(a_ia_j \le i+j\) for all distinct indices \(i, j\). Determine, with proof, the largest possible value of the product \(a_1a_2\cdots a_{2010}\).
[ "The largest possible value is\n\n\\[\n\\prod_{i=1}^{1005}(4i-1) = 3\\times 7 \\times \\ldots \\times 4019.\n\\]\n\n## Proof\n\nNo larger value is possible, since for each consecutive pair of elements: \\((a_{2i-1},a_{2i}), 1\\le i \\le 1005\\), the product is at most \\((2i-1) + 2i = 4i - 1\\), and so the product of all the pairs is at most:\n\n\\[\n\\prod_{i=1}^{1005}(4i-1).\n\\]\n\nIf we can demonstrate a sequence in which for all \\(1 \\le i \\le 1005\\) the product \\(a_{2i-1}a_{2i} = 4i-1\\), and all the inequalities are satisfied, the above upper bound will be achieved and the proof complete.\n\nWe will construct sequences of an arbitrarily large even length \\(2n \\ge 4\\), in which:\n\n\\[\n\\begin{cases} a_ia_j = i+j = 2i+1 & j=i+1, \\\\ a_ia_j = i+j = 2n-2 & j = i+2 = 2n, \\\\ a_ia_j < i+j & i \\ne 2n-2, \\mbox{ and } i < j-1. \\end{cases}\n\\]\n\nGiven \\(a_1\\), from the equations \\(a_ia_{i+1} = 2i+1,\\; 1\\le i\\le 2n-1\\), we obtain the whole sequence recursively: \\(a_1 = a_1,\\; a_2 = 3/a_1,\\; a_3 = 5/a_2 = 5a_1/3,\\; a_4 = 7/a_3 = (3\\cdot 7)/(5a_1) \\ldots.\\) And as a result:\n\n\\[\na_{i+2} = a_i \\cdot \\frac{2i+3}{2i+1}\\quad 1 \\le i \\le 2n-2.\n\\]\n\nThe same equations \\(a_ia_{i+1} = 2i+1\\) can be used to compute the whole sequence from any other known term.\n\nWe will often need to compare fractions in which the numerator and denominator are both positive, with fractions in which a positive term is added to both. Suppose \\(p, q, r\\) are three positive real numbers, then:\n\n\\[\n\\begin{align*} p &> q \\implies & \\frac{p+r}{q+r} &< \\frac{p}{q} \\\\ p &< q \\implies & \\frac{p+r}{q+r} &> \\frac{p}{q} \\\\ & \\mbox{and always} & \\frac{p+r}{q} &> \\frac{p}{q} \\end{align*}\n\\]\n\nReturning to the problem in hand, for \\(i < j\\), \\(a_ia_j \\le i+j \\implies a_ia_{j+2} < i+j+2\\). If it were otherwise, we would have for some \\(i < j\\):\n\n\\[\n\\frac{a_{j+2}}{a_j} \\ge \\frac{i+j+2}{i+j} > \\frac{j+j+2}{j+j} = \\frac{2j+2}{2j} > \\frac{2j+3}{2j+1} = \\frac{a_{j+2}}{a_j},\n\\]\n\nso our assumption is impossible.\n\nTherefore, we need only verify inequalities with an index difference of \\(1\\) or \\(2\\), as these imply the rest.\n\nNow, when the indices differ by \\(1\\) we have ensured equality (and hence the desired inequalities) by construction. So, we only need to prove the inequalities for successive even index and successive odd index pairs, i.e. for every index \\(i > 2\\), prove \\(a_{i-2}a_i \\le 2i-2\\).\n\nWe now compare \\(a_ia_{i+2}/(2i+2)\\) with \\(a_{i+2}a_{i+4}/(2i+6)\\). By our recurrence relations:\n\n\\[\n\\begin{align*} \\frac{a_{i+2}a_{i+4}}{2i+6} &= \\frac{a_ia_{i+2}}{2i+2} \\cdot \\frac{2i+2}{2i+6} \\cdot \\frac{a_{i+4}}{a_{i+2}}\\cdot \\frac{a_{i+2}}{a_i} \\\\ &= \\frac{a_ia_{i+2}}{2i+2}\\cdot \\frac{i+1}{i+3}\\cdot \\frac{2i+7}{2i+5}\\cdot \\frac{2i+3}{2i+1} \\\\ &= \\frac{a_ia_{i+2}}{2i+2}\\cdot \\frac{(i+1)(2i+7)(2i+3)}{(i+3)(2i+5)(2i+1)} \\\\ &= \\frac{a_ia_{i+2}}{2i+2}\\cdot \\frac{4i^3+24i^2+41i+21}{4i^3+24i^2+41i+15} \\\\ &= \\frac{a_ia_{i+2}}{2i+2}\\cdot \\left(1+\\frac{6}{4i^3+24i^2+41i+15}\\right) \\\\ &> \\frac{a_ia_{i+2}}{2i+2}. \\end{align*}\n\\]\n\nSo, for both odd and even index pairs, the strict inequality \\(a_ia_{i+2} < 2i+2\\) follows from \\(a_{i+2}a_{i+4} \\le 2i+6\\) and we need only prove the inequalities \\(a_{2n-3}a_{2n-1} \\le 4n-4\\) and \\(a_{2n-2}a_{2n} \\le 4n-2\\), the second of which holds (as an equality) by construction, so only the first remains.\n\nWe have not yet used the equation \\(a_{2n-2}a_{2n} = 4n-2\\), with this we can solve for the last three terms (or equivalently their squares) and thus compute the whole sequence. From the equations:\n\n\\[\n\\begin{align*} a_{2n-1}a_{2n} &= 4n-1 \\\\ a_{2n-2}a_{2n} &= 4n-2 \\\\ a_{2n-2}a_{2n-1} &= 4n-3 \\end{align*}\n\\]\n\nmultiplying any two and dividing by the third, we get:\n\n\\[\n\\begin{align*} a_{2n}^2 &= \\frac{(4n-2)(4n-1)}{4n-3} \\\\ a_{2n-1}^2 &= \\frac{(4n-3)(4n-1)}{4n-2} \\\\ a_{2n-2}^2 &= \\frac{(4n-3)(4n-2)}{4n-1} \\end{align*}\n\\]\n\nfrom which,\n\n\\[\na_{2n-3}^2 = \\frac{(4n-5)^2}{a_{2n-2}^2} = \\frac{(4n-5)^2(4n-1)}{(4n-3)(4n-2)}\n\\]\n\nWith the squares of the last four terms in hand, we can now verify the only non-redundant inequality:\n\n\\[\n\\begin{align*} a_{2n-3}^2a_{2n-1}^2 &= \\frac{(4n-5)^2}{a_{2n-2}^2}a_{2n-1}^2 \\\\ &= \\left(\\frac{(4n-5)(4n-1)}{4n-2}\\right)^2 \\\\ &= \\left(\\frac{16n^2 -24n + 5}{4n-2}\\right)^2 \\\\ &< \\left(\\frac{16n^2 -24n + 8}{4n-2}\\right)^2 \\\\ &= (4n-4)^2. \\end{align*}\n\\]\n\nThe inequality above follows because the numerator and denominator are both positive for \\(n > 1\\).\n\nThis completes the construction and the proof of all the inequalities, which miraculously reduced to just one inequality for the last pair of odd indices.\n\n## Additional observations\n\nIf we choose a different first term, say \\(a_1' = M\\cdot a_1\\), the sequence \\(a_i'\\) will have the form:\n\n\\[\n\\begin{align*} \ta_{2i-1}' &= Ma_{2i-1} \\\\ \ta_{2i}' &= \\frac{a_{2i}}{M} \\end{align*}\n\\]\n\nthe same holds if we have a longer sequence, at every index of the shorter sequence, the longer sequence will be a constant multiple (for all the odd terms) or dividend (for all the even terms) of the corresponding term of shorter sequence.\n\nWe observe that our solution is not unique, indeed for any \\(k>0\\), the same construction with \\(2n+2k\\) terms, truncated to just the first \\(2n\\) terms, yields a sequence \\(a'_i\\) which also satisfies all the required conditions, but in this case \\(a'_{2n-2}a'{2n} < 4n-2\\).\n\nWe could have constructed this alternative solution directly, by replacing the right hand side in the equation \\(a_{2n-2}a_{2n} = 4n-2\\) with any smaller value for which we still get \\(a_{2n-3}a_{2n-1} \\le 4n-4\\).\n\nIn the modified construction, for some constant \\(M > 1\\), we have:\n\n\\[\n\\begin{align*} \ta_{2n}^2 &= \\frac{1}{M}\\cdot \\frac{(4n-2)(4n-1)}{4n-3} \\\\ \ta_{2n-1}^2 &= M\\cdot \\frac{(4n-3)(4n-1)}{4n-2} \\\\ \ta_{2n-2}^2 &= \\frac{1}{M}\\cdot \\frac{(4n-3)(4n-2)}{4n-1} \\end{align*}\n\\]\n\nand so:\n\n\\[\n\\begin{align*} \ta_{2n-3}^2a_{2n-1}^2 &= \\frac{(4n-5)^2}{a_{2n-2}^2}a_{2n-1}^2 \\\\ \t &= M^2\\left(\\frac{(4n-5)(4n-1)}{4n-2}\\right)^2 \\\\ \t &= M^2\\left(\\frac{16n^2 -24n + 5}{4n-2}\\right)^2 \\\\ \t &= M^2\\left(\\frac{16n^2 -24n + 5}{16n^2 -24n + 8}\\right)^2 \t\t \\cdot \\left(\\frac{16n^2 -24n + 8}{4n-2}\\right)^2 \\\\ \t &= M^2\\left(\\frac{16n^2 -24n + 5}{16n^2 -24n + 8}\\right)^2 \t\t\\cdot (4n-4)^2. \\end{align*}\n\\]\n\nwhich satisfies the required inequality provided:\n\n\\[\n\\begin{align*} M^2 &\\le \\frac{16n^2 -24n + 8}{16n^2 -24n + 5} \\\\ &= \\frac{(4n-4)(4n-2)}{(4n-5)(4n-1)} = M_{\\mathrm{max}}^2. \\end{align*}\n\\]\n\nThe ratio \\(M_{\\mathrm{max}}\\), between the largest and smallest possible value of \\(a_{2n-3}\\) is in fact the ratio between the largest and smallest values of \\(a_1\\) that yield a sequence that meets the conditions for at least \\(2n\\) terms.\n\nIn the \\(n=2\\) case, the equation for \\(a_{2n-3}\\) gives: \\(a_1^2 = \\frac{21}{10}\\). We will next consider what happens to \\(a_1^2\\), and the sequence of squares in general, as \\(n\\) increases.\n\nLet \\(A_{n,2i-1}, A_{n,2i}\\) denote the \\(i^{\\mathrm{th}}\\) odd and even terms, respectively, of the unique sequence which satisfies our original equations and has \\(2n\\) terms in total. Let \\(A_{n+1,2i-1}, A_{n+1,2i}\\) be the odd and even terms of the solution with \\(2n+2\\) terms. We already noted that there must exist a constant \\(M_n\\) (that depends on \\(n\\), but not on \\(i\\)), such that:\n\n\\[\n\\begin{cases} \tA_{n+1,2i-1} = M_n\\cdot A_{n,2i-1},& 1 \\le i \\le n \\\\ \tA_{n+1,2i} = \\dfrac{A_{n,2i}}{M_n},& 1 \\le i \\le n \\end{cases}\n\\]\n\nThis constant is found explicitly by comparing the squares of the last term \\(A_{n,2n}\\) of the solution of length \\(2n\\) with the square of the third last term \\(A_{n+1,2n}\\) of the solution of length \\(2n+2\\):\n\n\\[\n\\begin{align*} \t(1/M_n)^2 &= \\frac{A_{n+1,2n}^2}{A_{n,2n}^2} \\\\ \t &= \\frac{(4n+1)(4n+2)}{4n+3} \\cdot \\frac{4n-3}{(4n-2)(4n-1)} \\\\ \t &= \\frac{32n^3 - 14n - 3}{32n^3 - 14n + 3} \\end{align*}\n\\]\n\nClearly \\(M_n > 1\\) for all positive \\(n\\), and so for fixed \\(i\\), the odd index terms \\(A_{n,2i-1}\\) strictly increase with \\(n\\), while the even index terms \\(A_{n,2i}\\) decrease with \\(n\\).\n\nTherefore, for \\(n \\ge 2\\),\n\n\\[\nA_{n,1}^2 = \\frac{21}{10} \\prod_{k=2}^{n-1} M_k^2 = \\frac{21}{10} \\cdot \\prod_{k=2}^{n-1} \t\\frac{(k+\\frac{3}{4})(k-\\frac{1}{2})(k-\\frac{1}{4})} \t {(k-\\frac{3}{4})(k+\\frac{1}{2})(k+\\frac{1}{4})}\n\\]\n\nThe product converges to a finite value even if taken infinitely far, and we can conclude (by a simple continuity argument) that there is a unique infinite positive sequence \\(A_\\omega\\), in which \\(A_{\\omega,i}A_{\\omega,i+1} = 2i+1\\), that satisfies all the inequalities \\(A_{\\omega,i}A_{\\omega,j} < i+j,\\; i \\le j - 2\\). The square of the first term of the infinite sequence is:\n\n\\[\n\\begin{align*} A_{\\omega,1}^2 &= \\frac{21}{10} \\cdot \\lim_{n \\to \\infty} \\prod_{k=2}^{n-1} \t\\frac{(k+\\frac{3}{4})(k-\\frac{1}{2})(k-\\frac{1}{4})} \t {(k-\\frac{3}{4})(k+\\frac{1}{2})(k+\\frac{1}{4})} \\\\ &= \\frac{21}{10} \\cdot \\lim_{n \\to \\infty} \\left( \\frac{\\Gamma(\\frac{5}{4}) \\Gamma(\\frac{5}{2}) \\Gamma(\\frac{9}{4})} \t {\\Gamma(\\frac{11}{4}) \\Gamma(\\frac{3}{2}) \\Gamma(\\frac{7}{4})}\\cdot \\frac{\\Gamma(n+\\frac{3}{4})\\Gamma(n-\\frac{1}{2})\\Gamma(n-\\frac{1}{4})} \t {\\Gamma(n-\\frac{3}{4})\\Gamma(n+\\frac{1}{2})\\Gamma(n+\\frac{1}{4})} \\right) \\\\ &= \\frac{1}{4} \\cdot \\frac{\\Gamma^2(\\frac{1}{4})}{\\Gamma^2(\\frac{3}{4})} \\cdot \\lim_{n \\to \\infty} \t\\frac{\\Gamma(n+\\frac{3}{4})\\Gamma(n-\\frac{1}{2})\\Gamma(n-\\frac{1}{4})} \t {\\Gamma(n-\\frac{3}{4})\\Gamma(n+\\frac{1}{2})\\Gamma(n+\\frac{1}{4})} \\\\ &= \\frac{1}{4} \\cdot \\frac{\\Gamma^2(\\frac{1}{4})}{\\Gamma^2(\\frac{3}{4})} \\cdot 1 \\quad \\mbox{(Stirling's approximation)}\\\\ &= \\frac{\\Gamma^4(\\frac{1}{4})}{8\\pi^2} = \\frac{\\pi}{\\mathrm{AGM}^2(\\sqrt{2}, 1)} \\approx 2.1884396152264766\\ldots \\end{align*}\n\\]\n\nIn summary, if we set \\(a_1 = \\frac{\\sqrt{\\pi}}{\\mathrm{AGM}(\\sqrt{2}, 1)}\\), and then recursively set \\(a_{i+1} = (2i + 1)/a_i\\), we get an infinite sequence that, for all \\(n \\ge 1\\), yields the maximum possible product \\(a_1a_2\\cdots a_{2n}\\), subject to the conditions \\(a_ia_j \\le i+j,\\; 1 \\le i < j \\le 2n\\)." ]
USAMO-2010-4
https://artofproblemsolving.com/wiki/index.php/2010_USAMO_Problems/Problem_4
Let \(ABC\) be a triangle with \(\angle A = 90^{\circ}\). Points \(D\) and \(E\) lie on sides \(AC\) and \(AB\), respectively, such that \(\angle ABD = \angle DBC\) and \(\angle ACE = \angle ECB\). Segments \(BD\) and \(CE\) meet at \(I\). Determine whether or not it is possible for segments \(AB, AC, BI, ID, CI, IE\) to all have integer lengths.
[ "We know that angle \\(BIC = 135^{\\circ}\\), as the other two angles in triangle \\(BIC\\) add to \\(45^{\\circ}\\). Assume that only \\(AB, AC, BI\\), and \\(CI\\) are integers. Using the Law of Cosines on triangle BIC,\n\n\\[\n[asy] import olympiad; // Scale unitsize(1inch); // Shape real h = 1.75; real w = 2.5; // Points void ldot(pair p, string l, pair dir=p) { dot(p); label(l, p, unit(dir)); } pair A = origin; ldot(A, \"$A$\", plain.SW); pair B = w * plain.E; ldot(B, \"$B$\", plain.SE); pair C = h * plain.N; ldot(C, \"$C$\", plain.NW); pair D = extension(B, bisectorpoint(C, B, A), A, C); ldot(D, \"$D$\", D-B); pair E = extension(C, bisectorpoint(A, C, B), A, B); ldot(E, \"$E$\", E-C); pair I = extension(B, D, C, E); ldot(I, \"$I$\", A-I); // Segments draw(A--B); draw(B--C); draw(C--A); draw(C--E); draw(B--D); // Angles import markers; draw(rightanglemark(B, A, C, 4)); markangle(Label(\"$\\scriptstyle{\\frac{\\theta}{2}}$\"), radius=40, I, B, E); markangle(Label(\"$\\scriptstyle{\\frac{\\theta}{2}}$\"), radius=40, C, B, I); markangle(Label(\"$\\scriptstyle{\\frac{\\pi}{4} - \\frac{\\theta}{2}}$\"), radius=40, I, C, B); markangle(Label(\"$\\scriptstyle{\\frac{\\pi}{4} - \\frac{\\theta}{2}}$\"), radius=40, D, C, I); markangle(Label(\"$\\scriptstyle{\\frac{3\\pi}{4}}$\"), radius=10, B, I, C); [/asy]\n\\]\n\n\\(BC^2 = BI^2 + CI^2 - 2BI\\cdot CI \\cdot \\cos 135^{\\circ}\\). Observing that \\(BC^2 = AB^2 + AC^2\\) is an integer and that \\(\\cos 135^{\\circ} = -\\frac{\\sqrt{2}}{2},\\) we have\n\n\\[\nBC^2 - BI^2 - CI^2 = BI\\cdot CI\\cdot \\sqrt{2}\n\\]\n\nand therefore,\n\n\\[\n\\sqrt{2} = \\frac{BC^2 - BI^2 - CI^2}{BI\\cdot CI}\n\\]\n\nThe LHS (\\(\\sqrt{2}\\)) is irrational, while the RHS is the quotient of the division of two integers and thus is rational. Clearly, there is a contradiction. Therefore, it is impossible for \\(AB, AC, BI\\), and \\(CI\\) to all be integers, which invalidates the original claim that all six lengths are integers, and we are done.", "The answer is no.\n\nSuppose otherwise. It is easy to see (through simple angle chasing) that \\(\\angle DIC=45^{\\circ}\\). Also, since \\(I\\) is the incenter, we have \\(\\angle IAC = 45^{\\circ}\\). Using the Law of Cosines, we have\n\n\\[\nCD^2=IC^2+ID^2-\\sqrt{2}(IC)(ID),\n\\]\n\nso that \\(CD\\) is irrational. But \\(\\triangle IAC \\sim \\triangle DIC\\), thus \\(IC^2=CD\\cdot AC\\), implying that \\(CD\\) is rational, contradiction. \\(\\blacksquare\\)", "The result can be also proved without direct appeal to trigonometry, via just the angle bisector theorem and the structure of Pythagorean triples. (This is a lot more work).\n\nA triangle in which all the required lengths are integers exists if and only if there exists a triangle in which \\(AB\\) and \\(AC\\) are relatively-prime integers and the lengths of the segments \\(BI, ID, CI, IE\\) are all rational (we divide all the lengths by the \\(\\gcd(AB, AC)\\) or conversely multiply all the lengths by the least common multiple of the denominators of the rational lengths).\n\nSuppose there exists a triangle in which the lengths \\(AB\\) and \\(AC\\) are relatively-prime integers and the lengths \\(IB, ID, CI, IE\\) are all rational.\n\nSince \\(CE\\) is the bisector of \\(\\angle ACB\\), by the angle bisector theorem, the ratio \\(IB : ID = CB : CD\\), and since \\(BD\\) is the bisector of \\(\\angle ABC\\), \\(CB : CD = (AB + BC) : AC\\). Therefore, \\(IB : ID = (AB + BC) : AC\\). Now \\(IB : ID\\) is by assumption rational, so \\((AB + BC) : AC\\) is rational, but \\(AB\\) and \\(AC\\) are assumed integers so \\(BC : AC\\) must also be rational. Since \\(BC\\) is the hypotenuse of a right-triangle, its length is the square root of an integer, and thus either an integer or irrational, so \\(BC\\) must be an integer.\n\nWith \\(AB\\) and \\(AC\\) relatively-prime, we conclude that the side lengths of \\(\\triangle ABC\\) must be a Pythagorean triple: \\((2pq, p^2 - q^2, p^2 + q^2)\\), with \\(p > q\\) relatively-prime positive integers and \\(p+q\\) odd.\n\nWithout loss of generality, \\(AC = 2pq, AB = p^2 - q^2, BC = p^2+q^2\\). By the angle bisector theorem,\n\n\\[\n\\begin{align*} AE &= \\dfrac{AB \\cdot AC}{AC + CB} = \\dfrac{2pq(p^2-q^2)}{p^2 + q^2 + 2pq} = \\dfrac{2pq(p-q)}{p+q} \\end{align*}\n\\]\n\nSince \\(\\triangle CAE\\) is a right-triangle, we have:\n\n\\[\n\\begin{align*} CE^2 &= AC^2 + AE^2 = 4p^2q^2 + \\left(\\dfrac{2pq(p-q)}{p+q}\\right)^2 = 4p^2q^2\\left[1 + \\left(\\dfrac{p-q}{p+q}\\right)^2\\right] \\\\ &= \\frac{4p^2q^2}{(p+q)^2}\\left[(p+q)^2 + (p-q)^2\\right] = \\frac{4p^2q^2}{(p+q)^2}(2p^2 + 2q^2) \\end{align*}\n\\]\n\nand so \\(CE\\) is rational if and only if \\(2p^2 + 2q^2\\) is a perfect square.\n\nAlso by the angle bisector theorem,\n\n\\[\n\\begin{align*} AD &= \\dfrac{AB \\cdot AC}{AB + BC} = \\dfrac{2pq(p^2-q^2)}{p^2 + q^2 + p^2 - q^2} = \\dfrac{q(p^2-q^2)}{p} \\end{align*}\n\\]\n\nand therefore, since \\(\\triangle DAB\\) is a right-triangle, we have:\n\n\\[\n\\begin{align*} BD^2 &= AB^2 + AD^2 = (p^2-q^2)^2 + \\left(\\dfrac{q(p^2-q^2)}{p}\\right)^2 \\\\ &= (p^2-q^2)^2\\left[1 + \\frac{q^2}{p^2}\\right] = \\frac{(p^2-q^2)^2}{p^2}(p^2 + q^2) \\end{align*}\n\\]\n\nand so \\(BD\\) is rational if and only if \\(p^2 + q^2\\) is a perfect square.\n\nCombining the conditions on \\(CE\\) and \\(BD\\), we see that \\(2p^2+2q^2\\) and \\(p^2+q^2\\) must both be perfect squares. If it were so, their ratio, which is \\(2\\), would be the square of a rational number, but \\(\\sqrt{2}\\) is irrational, and so the assumed triangle cannot exist.", "We proceed by contradiction.\n\nFTSOC, let \\(AB, BI, ID\\) have integer lengths. Then \\(BD = BI + ID \\in \\mathbb{Z}\\) as well. By trigonometry,\n\n\\[\nBD = \\frac{AB}{\\cos{(\\frac{\\angle ABC}{2})}}.\n\\]\n\nRearranging we find \\(\\cos{(\\frac{\\angle ABC}{2})} = \\frac{AB}{BD} \\in \\mathbb{Q}\\). As \\(0 < \\angle ABC < \\frac{\\pi}{2}\\) so \\(0 < \\frac{\\angle ABC}{2} < \\frac{\\pi}{4}\\) but there are no possible angles in the interval that result in a rational cosine by Niven's theorem. So we have contradiction.\n\n~Aaryabhatta1." ]
USAMO-2010-5
https://artofproblemsolving.com/wiki/index.php/2010_USAMO_Problems/Problem_5
Let \(q = \dfrac{3p-5}{2}\) where \(p\) is an odd prime, and let \[ S_q = \frac{1}{2\cdot 3 \cdot 4} + \frac{1}{5\cdot 6 \cdot 7} + \cdots + \frac{1}{q\cdot (q+1) \cdot (q+2)}. \] Prove that if \(\dfrac{1}{p}-2S_q = \dfrac{m}{n}\) for integers \(m\) and \(n\), then \(m-n\) is divisible by \(p\).
[ "Since \\(p\\) is an odd prime, \\(p = 2r + 1\\), for a suitable positive integer \\(r\\), and consequently \\(q = 3r - 1\\).\n\nThe partial-fraction decomposition of the general term of \\(S_q\\) is:\n\n\\[\n\\begin{align*} \\frac{1}{(3k-1)3k(3k+1)} &= \\frac{1}{2}\\left(\\frac{1}{3k-1} - \\frac{2}{3k} + \\frac{1}{3k+1}\\right) \\\\ &= \\frac{1}{2}\\left(\\frac{1}{3k-1} + \\frac{1}{3k} - \\frac{1}{k} + \\frac{1}{3k+1}\\right) \\\\ &= \\frac{1}{2}\\left[\\left(\\frac{1}{3k-1} + \\frac{1}{3k} + \\frac{1}{3k+1}\\right) - \\frac{1}{k}\\right], \\end{align*}\n\\]\n\ntherefore\n\n\\[\n\\begin{align*} \\frac{1}{p} - 2S_q &= \\frac{1}{2r+1} - \\left(\\sum_{k=2}^{3r+1}\\frac{1}{k} - \\sum_{k=1}^{r} \\frac{1}{k}\\right) \\\\ &= \\frac{1}{2r+1} - \\left[\\left(\\sum_{k=r+1}^{3r+1}\\frac{1}{k}\\right)- 1\\right] \\\\ &= 1 - \\left(\\sum_{k=r+1}^{2r}\\frac{1}{k} + \\sum_{k=2r+2}^{3r+1}\\frac{1}{k}\\right) \\\\ &= 1 - \\sum_{k=1}^{r}\\left(\\frac{1}{(2r+1) - k} + \\frac{1}{(2r+1) + k}\\right) \\\\ &= 1 - \\sum_{k=1}^{r}\\frac{2p}{(p - k)(p + k)} \\\\ &= 1 - \\frac{a}{b} = \\frac{b - a}{b} \\end{align*}\n\\]\n\nwith \\(a\\) and \\(b\\) positive relatively-prime integers.\n\nSince \\(r < p\\) and \\(p\\) is a prime, in the final sum all the denominators are relatively prime to \\(p\\), but all the numerators are divisible by \\(p\\), and therefore the numerator \\(a\\) of the reduced fraction \\(\\frac{a}{b}\\) will be divisible by \\(p\\). Since the sought difference \\(m - n = (b-a) - b = -a\\), we conclude that \\(p\\) divides \\(m-n\\) as required.\n\n## Alternative Calculation\n\nWe can obtain the result in a slightly different way:\n\n\\[\n\\begin{align*} \\frac{1}{p} - 2S_q &= \\frac{1}{2r+1} - \\left(\\sum_{k=2}^{3r+1}\\frac{1}{k} - \\sum_{k=1}^{r} \\frac{1}{k}\\right) \\\\ &= \\frac{1}{2r+1} - \\left(\\sum_{k=r+1}^{3r+1}\\frac{1}{k} - 1\\right) \\\\ &= 1 - \\left(\\sum_{k=r+1}^{2r}\\frac{1}{k} + \\sum_{k=2r+2}^{3r+1}\\frac{1}{k}\\right) \\end{align*}\n\\]\n\nIn the above sum the denominators of the fractions represent each non-zero remainder \\(\\pmod p\\) exactly once. Multiplying all the denominators yields a number \\(N\\) that is \\(p-1 \\pmod p\\). The numerator \\(\\pmod p\\) is \\(N\\) times the sum of the \\(\\pmod p\\) inverses of each non-zero remainder, and since this sum is \\(0 \\pmod p\\), the numerator is \\(0 \\pmod p\\). The rest of the argument is as before." ]
USAMO-2010-6
https://artofproblemsolving.com/wiki/index.php/2010_USAMO_Problems/Problem_6
A blackboard contains 68 pairs of nonzero integers. Suppose that for each positive integer \(k\) at most one of the pairs \((k, k)\) and \((-k, -k)\) is written on the blackboard. A student erases some of the 136 integers, subject to the condition that no two erased integers may add to 0. The student then scores one point for each of the 68 pairs in which at least one integer is erased. Determine, with proof, the largest number \(N\) of points that the student can guarantee to score regardless of which 68 pairs have been written on the board.
[ "Answer: 43\n\nAttainability: Consider 8 distinct positive numbers. Let there be 5 pairs for each of the numbers including 2 clones of that number. Let there also be 28 pairs that include the negatives of those numbers such that each negative associates with another negative once and exactly once (in graph theoretic terms, a K8). Let x be the number of positives chosen out of these 8 (assume the other chosen numbers are negatives) for erasing; then, the number of points the student scores is \\(28 + 5x - {x \\choose 2}\\), which is maximized at x=5 and x=6, and the maximum value is \\(43\\). Choosing the first 5 numbers as positive and the other three as negative attains this. Hence, 43 is a possible maximum possible score.\n\nBounding: We use expected values. WLOG all the pairs of numbers with both numbers identical have only positive values. Consider flipping a weighted coin whether to choose the positive number or its negation for each positive number; it chooses positive with probability p. The pairs with both numbers same are chosen with probability p, the pairs (k, -k) are chosen with probability 1, and the pairs (x, y) for distinct x and y that sum to a nonzero number are chosen with probability \\(1-p^2\\). We are trying to minimize the expected value, so we can assume that no pairs (k, -k) exist. Let A be the number of (k, k) pairs, and 68-A be the number of (x, y) pairs. The expected number of points scored is \\(Ap + (68-A)(1-p^2)\\). We want to prove this is larger than 42 at all times for some choice of p. If \\(A < 36\\), \\(1/2\\) works for p to give this bound. If \\(A > 36\\), \\(5/8\\) works for p for p to give the desired bound. If \\(A = 36\\), we can use \\(3/5\\) for p to get the desired bound. Hence, in any case, the expected value for the number of points scored across all of these weighted processes is larger than 42, so there exists some case that gives a score of 43. Hence, bounding is complete. We are done with both parts. Q.E.D.\n\n-Solution by thanosaops" ]
USAMO-2011-1
https://artofproblemsolving.com/wiki/index.php/2011_USAMO_Problems/Problem_1
Let \(a\), \(b\), \(c\) be positive real numbers such that \(a^2 + b^2 + c^2 + (a + b + c)^2 \le 4\). Prove that \[ \frac{ab + 1}{(a + b)^2} + \frac{bc + 1}{(b + c)^2} + \frac{ca + 1}{(c + a)^2} \ge 3. \]
[ "Since\n\n\\[\n\\begin{align*} (a+b)^2 + (b+c)^2 + (c+a)^2 &= 2(a^2 + b^2 + c^2 + ab + bc + ca) \\\\ \t&= a^2 + b^2 + c^2 + (a + b + c)^2, \\end{align*}\n\\]\n\nit is natural to consider a change of variables:\n\n\\[\n\\begin{align*} \\alpha &= b + c \\\\ \\beta &= c + a \\\\ \\gamma &= a + b \\end{align*}\n\\]\n\nwith the inverse mapping given by:\n\n\\[\n\\begin{align*} a &= \\frac{\\beta + \\gamma - \\alpha}2 \\\\ b &= \\frac{\\alpha + \\gamma - \\beta}2 \\\\ c &= \\frac{\\alpha + \\beta - \\gamma}2 \\end{align*}\n\\]\n\nWith this change of variables, the constraint becomes\n\n\\[\n\\alpha^2 + \\beta^2 + \\gamma^2 \\le 4,\n\\]\n\nwhile the left side of the inequality we need to prove is now\n\n\\[\n\\begin{align*} & \\frac{\\gamma^2 - (\\alpha - \\beta)^2 + 4}{4\\gamma^2} + \\frac{\\alpha^2 - (\\beta - \\gamma)^2 + 4}{4\\alpha^2} + \\frac{\\beta^2 - (\\gamma - \\alpha)^2 + 4}{4\\beta^2} \\ge \\\\ & \\frac{\\gamma^2 - (\\alpha - \\beta)^2 + \\alpha^2 + \\beta^2 + \\gamma^2}{4\\gamma^2} + \\frac{\\alpha^2 - (\\beta - \\gamma)^2 + \\alpha^2 + \\beta^2 + \\gamma^2}{4\\alpha^2} + \\frac{\\beta^2 - (\\gamma - \\alpha)^2 + \\alpha^2 + \\beta^2 + \\gamma^2}{4\\beta^2} = \\\\ & \\frac{2\\gamma^2 + 2\\alpha\\beta}{4\\gamma^2} + \\frac{2\\alpha^2 + 2\\beta\\gamma}{4\\alpha^2} + \\frac{2\\beta^2 + 2\\gamma\\alpha}{4\\beta^2} = \\\\ & \\frac32 + \\frac{\\alpha\\beta}{2\\gamma^2} + \\frac{\\beta\\gamma}{2\\alpha^2} + \\frac{\\gamma\\alpha}{2\\beta^2}. \\end{align*}\n\\]\n\nTherefore it remains to prove that\n\n\\[\n\\frac{\\alpha\\beta}{2\\gamma^2} + \\frac{\\beta\\gamma}{2\\alpha^2} + \\frac{\\gamma\\alpha}{2\\beta^2} \\ge \\frac32.\n\\]\n\nWe note that the product of the three (positive) terms is 1/8, therefore by AM-GM their mean is at least 1/2, and thus their sum is at least 3/2 and we are done.", "Rearranging the condition yields that\n\n\\[\na^2 + b^2 + c^2 +ab+bc+ac \\le 2\n\\]\n\nNow note that\n\n\\[\n\\frac{2ab+2}{(a+b)^2} \\ge \\frac{2ab+a^2 + b^2 + c^2 +ab+bc+ac}{(a+b)^2}=\\frac{(a+b)^2 + (c+a)(c+b)}{(a+b)^2}\n\\]\n\nSumming this for all pairs of \\(\\{ a,b,c \\}\\) gives that\n\n\\[\n\\sum_{cyc} \\frac{2ab+2}{(a+b)^2} \\ge 3+ \\sum_{cyc}\\frac{(c+a)(c+b)}{(a+b)^2} \\ge 6\n\\]\n\nBy AM-GM. Dividing by \\(2\\) gives the desired inequality.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2011-2
https://artofproblemsolving.com/wiki/index.php/2011_USAMO_Problems/Problem_2
An integer is assigned to each vertex of a regular pentagon so that the sum of the five integers is 2011. A turn of a solitaire game consists of subtracting an integer \(m\) from each of the integers at two neighboring vertices and adding 2m to the opposite vertex, which is not adjacent to either of the first two vertices. (The amount \(m\) and the vertices chosen can vary from turn to turn.) The game is won at a certain vertex if, after some number of turns, that vertex has the number 2011 and the other four vertices have the number 0. Prove that for any choice of the initial integers, there is exactly one vertex at which the game can be won.
[ "Let \\(\\mathbb{F}_5\\) be the field of positive residues modulo 5. We label the vertices of the pentagon clockwise with the residues 0, 1, 2, 3 and 4. For each \\(i \\in \\mathbb{F}_5\\) let \\(n_i\\) be the integer at vertex \\(i\\) and let \\(r_i \\in \\mathbb{F}_5\\) be defined as:\n\n\\[\nr_i \\equiv 4n_{i+1} + 3n_{i+2} + 2n_{i+3} + n_{i+4} \\equiv \\sum_{k\\in\\mathbb{F}_5}(i-k)n_k \\pmod 5.\n\\]\n\nLet \\(s = \\sum_{i\\in\\mathbb{F}_5} n_i\\). A move in the game consists of\n\n\\[\n(n_i, n_{i+1}, n_{i+2}, n_{i+3}, n_{i+4}) \\mapsto (n_i + 2m, n_{i+1}, n_{i+2} - m, n_{i+3} - m, n_{i+4})\n\\]\n\nfor some vertex \\(i \\in \\mathbb{F}_5\\) and integer \\(m\\). We immediately see that \\(s\\) is an invariant of the game. After our move the new value of \\(r_i\\) is decreased by \\(3m + 2m \\equiv 0 \\pmod 5\\) as a result of the change in the \\(3n_{i+2}\\) and \\(2n_{i+3}\\) terms. So \\(r_i\\) does not change after a move at vertex \\(i\\).\n\nFor all \\(i \\in \\mathbb{F}_5\\) we have:\n\n\\[\nr_{i+1} - r_i \\equiv \\sum_{k\\in\\mathbb{F}_5} ((i+1-k)n_k - (i-k)n_k) \\equiv \\sum_{k\\in\\mathbb{F}_5} n_k \\equiv s \\pmod 5.\n\\]\n\nTherefore, the \\(r_i\\) form an arithmetic progression in \\(\\mathbb{F}_5\\) with a difference of \\(s\\). Since \\(r_k\\) is unchanged by a move at vertex \\(k\\), so are all the remaining \\(r_i\\) as the differences are constant.\n\nProvided \\(s \\not\\equiv 0 \\pmod 5\\), we see that the mapping \\(i \\mapsto r_i\\) is a bijection \\(\\mathbb{F}_5 \\to \\mathbb{F}_5\\) and exactly one vertex will have \\(r_i \\equiv 0 \\pmod 5\\). As \\(r_i\\) is an invariant, a winning vertex must have \\(r_i \\equiv 0\\), since in the final state each \\(n_k\\) with \\(k \\ne i\\) is zero. So, for \\(s \\not\\equiv 0 \\pmod 5\\), if a winning vertex exists, it is the unique vertex with \\(r_i \\equiv 0\\).\n\nWithout loss of generality, it remains to show that if \\(r_0 \\equiv 0 \\pmod 5\\), then 0 must be a winning vertex. To prove this, we perform the following moves:\n\n\\[\n\\begin{align*} (n_0, n_1, n_2, n_3, n_4) &\\mapsto (n_0-n_1, 0, n_2, n_3 + 2n_1, n_4), & (i, m) &= (3, n_1) \\\\ &\\mapsto (n_0-n_1-n_4, 0, n_2+2n_4, n_3 +2n_1, 0) & (i, m) &= (2, n_4) \\end{align*}\n\\]\n\nWe designate the new state \\((n'_0, 0, n'_2, n'_3, 0)\\). Since \\(r_0 \\equiv 0\\) is an invariant, and \\(n'_1 = n'_4 = 0\\), we now have \\(r_0 \\equiv n'_3 - n'_2 = 5p\\), for some integer \\(p\\). Our final set of moves is:\n\n\\[\n\\begin{align*} (n'_0, 0, n'_2, n'_3, 0) &\\mapsto (n'_0+p, p, n'_2, n'_3 - 2p, 0), & (i, m) &= (3, -p)\\\\ &\\mapsto (n'_0+p, 0, n'_2-p, n'_3 - 2p, 2p), & (i, m) &= (4, p) \\\\ &\\mapsto (n'_0-p, 0, n'_2+3p, n'_3 - 2p, 0), & (i, m) &= (2, 2p)\\\\ &\\mapsto (n'_0+2n'_2 + 5p, 0, 0, n'_3-n'_2 - 5p = 0, 0) & (i, m) &= (0, n'_2 + 3p) \\end{align*}\n\\]\n\nNow our chosen vertex 0 is the only vertex with a non-zero value, and since \\(s\\) is invariant, that value is \\(s\\) as required. Since a vertex \\(i\\) with \\(r_i \\equiv 0 \\pmod 5\\) is winnable, and with \\(s = 2011 \\not\\equiv 0 \\pmod 5\\) we always have a unique such vertex, we are done.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2011-3
https://artofproblemsolving.com/wiki/index.php/2011_USAMO_Problems/Problem_3
In hexagon \(ABCDEF\), which is nonconvex but not self-intersecting, no pair of opposite sides are parallel. The internal angles satisfy \(\angle A = 3\angle D\), \(\angle C = 3\angle F\), and \(\angle E = 3\angle B\). Furthermore \(AB=DE\), \(BC=EF\), and \(CD=FA\). Prove that diagonals \(\overline{AD}\), \(\overline{BE}\), and \(\overline{CF}\) are concurrent.
[ "Let \\(\\angle D = \\alpha\\), \\(\\angle F = \\gamma\\), and \\(\\angle B = \\beta\\), \\(AB=DE=p\\), \\(BC=EF=q\\), \\(CD=FA=r\\). Define the vectors:\n\n\\[\n\\vec{u} = \\vec{AB} + \\vec{DE}\n\\]\n\n\\[\n\\vec{v} = \\vec{BC} + \\vec{EF}\n\\]\n\n\\[\n\\vec{w} = \\vec{CD} + \\vec{FA}\n\\]\n\nClearly, \\(\\vec{u}+\\vec{v}+\\vec{w}=\\textbf{0}\\).\n\nLet \\(AB\\) intersect \\(DE\\) at \\(X\\). Note that \\(\\angle X = 360^\\circ - \\angle D - \\angle C - \\angle B = 360^\\circ - \\alpha - 3\\gamma - \\beta = 180^\\circ - 2\\gamma\\). Define the points \\(M\\) and \\(N\\) on lines \\(AB\\) and \\(DE\\) respectively so that \\(\\vec{MX} = \\vec{AB}\\) and \\(\\vec{XN} = \\vec{DE}\\). Then \\(\\vec{u} = \\vec{MN}\\). As \\(XMN\\) is isosceles with \\(XM = XN = p\\), the base angles are both \\(\\gamma\\). Thus, \\(|\\vec{u}|=2p \\cos \\gamma\\). Similarly, \\(|\\vec{v}|=2q \\cos \\alpha\\) and \\(|\\vec{w}| = 2r \\cos \\beta\\).\n\nNext we will find the angles between \\(\\vec{u}\\), \\(\\vec{v}\\), and \\(\\vec{w}\\). As \\(\\angle MNX = \\gamma\\), the angle between the vectors \\(\\vec{u}\\) and \\(\\vec{NE}\\) is \\(\\gamma\\). Similarly, the angle between \\(\\vec{EF}\\) and \\(\\vec{v}\\) is \\(\\alpha\\). Since the angle between \\(\\vec{NE}\\) and \\(\\vec{EF}\\) is \\(\\angle E = 3\\beta\\), the angle between \\(\\vec{u}\\) and \\(\\vec{v}\\) is \\(360^\\circ - \\gamma - 3\\beta - \\alpha = 180^\\circ - 2\\beta\\). Similarly, the angle between \\(\\vec{v}\\) and \\(\\vec{w}\\) is \\(180^\\circ - 2\\gamma\\), and the angle between \\(\\vec{w}\\) and \\(\\vec{u}\\) is \\(180^\\circ - 2\\alpha\\).\n\nAnd since \\(\\vec{u}+\\vec{v}+\\vec{w}=\\vec{0}\\), we can arrange the three vectors to form a triangle, so the triangle with sides of lengths \\(2p \\cos \\gamma\\), \\(2q \\cos \\alpha\\), and \\(2r \\cos \\beta\\) has opposite angles of \\(180^\\circ - 2\\gamma\\), \\(180^\\circ - 2\\alpha\\), and \\(180^\\circ - 2\\beta\\), respectively. So by the law of sines:\n\n\\[\n\\frac{2p \\cos \\gamma}{\\sin 2\\gamma} = \\frac{2q \\cos \\alpha}{\\sin 2\\alpha} = \\frac{2r \\cos \\beta}{\\sin 2\\beta}\n\\]\n\n\\[\n\\frac{p}{\\sin \\gamma} = \\frac{q}{\\sin \\alpha} = \\frac{r}{\\sin \\beta},\n\\]\n\nand the triangle with sides of length \\(p\\), \\(q\\), and \\(r\\) has corresponding angles of \\(\\gamma\\), \\(\\alpha\\), and \\(\\beta\\). It follows by SAS congruency that this triangle is congruent to \\(FAB\\), \\(BCD\\), and \\(DEF\\), so \\(FD=p\\), \\(BF=q\\), and \\(BD=r\\), and \\(D\\), \\(F\\), and \\(B\\) are the reflections of the vertices of triangle \\(ACE\\) about the sides. So \\(AD\\), \\(BE\\), and \\(CF\\) concur at the orthocenter of triangle \\(ACE\\).", "We work in the complex plane, where lowercase letters denote their corresponding point's poition. Let \\(P\\) denote hexagon \\(ABCDEF\\). Since \\(AB=DE\\), the condition \\(AB\\not\\parallel DE\\) is equivalent to \\(a-b+d-e\\ne 0\\).\n\nConstruct a \"phantom hexagon\" \\(P'=A'B'C'D'E'F'\\) as follows: let \\(A'C'E'\\) be a triangle with \\(\\angle{A'C'E'}=\\angle{F}\\), \\(\\angle{C'E'A'}=\\angle{B}\\), and \\(\\angle{E'A'C'}=\\angle{F}\\) (this is possible since \\(\\angle{B}+\\angle{D}+\\angle{F}=180^\\circ\\) by the angle conditions), and reflect \\(A',C',E'\\) over its sides to get points \\(D',F',B'\\), respectively. By rotation and reflection if necessary, we assume \\(A'B'\\parallel AB\\) and \\(P',P\\) have the same orientation (clockwise or counterclockwise), i.e. \\(\\frac{b-a}{b'-a'}\\in\\mathbb{R}^+\\). It's easy to verify that \\(\\angle{X'}=\\angle{X}\\) for \\(X\\in\\{A,B,C,D,E,F\\}\\) and opposite sides of \\(P'\\) have equal lengths. As the corresponding sides of \\(P\\) and \\(P'\\) must then be parallel, there exist positive reals \\(r,s,t\\) such that \\(r=\\frac{a-b}{a'-b'}=\\frac{d-e}{d'-e'}\\), \\(s=\\frac{b-c}{b'-c'}=\\frac{e-f}{e'-f'}\\), and \\(t=\\frac{c-d}{c'-d'}=\\frac{f-a}{f'-a'}\\). But then \\(0\\ne a-b+d-e=r(a'-b'+d'-e')\\), etc., so the non-parallel condition \"transfers\" directly from \\(P\\) to \\(P'\\) and\n\n\\[\n\\begin{align*} 0 &=(a-b+d-e)+(b-c+e-f)+(c-d+f-a) \\\\ &=r(a'-b'+d'-e')+s(b'-c'+e'-f')+t(c'-d'+f'-a') \\\\ &=(r-t)(a'-b'+d'-e')+(s-t)(b'-c'+e'-f'). \\end{align*}\n\\]\n\nIf \\(r-t=s-t=0\\), then \\(P\\) must be similar to \\(P'\\) and the conclusion is obvious.\n\nOtherwise, since \\(a'-b'+d'-e'\\ne0\\) and \\(b'-c'+e'-f'\\ne0\\), we must have \\(r-t\\ne0\\) and \\(s-t\\ne0\\). Now let \\(x=\\frac{a'+d'}{2}\\), \\(y=\\frac{c'+f'}{2}\\), \\(z=\\frac{e'+b'}{2}\\) be the feet of the altitudes in \\(\\triangle{A'C'E'}\\); by the non-parallel condition in \\(P'\\), \\(x,y,z\\) are pairwise distinct. But \\(\\frac{z-x}{z-y}=\\frac{s-t}{r-t}\\in\\mathbb{R}\\), whence \\(x,y,z\\) are three distinct collinear points, which is clearly impossible. (The points can only be collinear when \\(\\triangle{A'C'E'}\\) is a right triangle, but in this case two of \\(x,y,z\\) must coincide.)\n\nAlternatively (for the previous paragraph), WLOG assume that \\((A'C'E')\\) is the unit circle, and use the fact that \\(b'=a'+c'-\\frac{a'c'}{e'}\\), etc. to get simple expressions for \\(a'-b'+d'-e'\\) and \\(b'-c'+e'-f'\\).", "We work in the complex plane to give (essentially) a complete characterization when the parallel condition is relaxed.\n\nWLOG assume \\(a,b,c\\) are on the unit circle. It suffices to show that \\(a,b,c\\) uniquely determine \\(d,e,f\\), since we know that if we let \\(E\\) be the reflection of \\(B\\) over \\(AC\\), \\(D\\) be the reflection of \\(A\\) over \\(CE\\), and \\(F\\) be the reflection of \\(C\\) over \\(AE\\), then \\(ABCDEF\\) satisfies the problem conditions. (*)\n\nIt's easy to see with the given conditions that\n\n\\[\n\\begin{align*} (a-b)(c-d)(e-f) &= (b-c)(d-e)(f-a) \\Longleftrightarrow f=\\frac{(a-b)(c-d)e+(c-b)(e-d)a}{(a-b)(c-d)+(c-b)(e-d)} \\\\ \\frac{(e-a)(c-b)}{(a-b)(c-d)+(c-b)(e-d)} = \\frac{f-e}{d-e} &= \\left(\\frac{c-b}{a-b}\\right)^2 \\overline{\\left(\\frac{a-b}{c-b}\\right)} = \\frac{c-b}{a-b}\\cdot\\frac{c}{a} \\Longleftrightarrow d=\\frac{c[(a-b)c+(c-b)e]+a(a-e)(a-b)}{c[(a-b)+(c-b)]} \\\\ \\frac{(a-b)(c-d)+(c-b)(e-d)}{(a-e)(c-d)} = \\frac{b-a}{f-a} &= \\left(\\frac{e-d}{c-d}\\right)^2 \\overline{\\left(\\frac{c-d}{e-d}\\right)}. \\end{align*}\n\\]\n\nNote that\n\n\\[\n\\frac{e-d}{c-d}=\\frac{(a-b)[c(e-c)+a(e-a)]}{c(c-e)(c-b)-a(a-e)(a-b)},\n\\]\n\nso plugging into the third equation we have\n\n\\[\n\\begin{align*} \\frac{a(a-b)(2b-a-c)}{c(c-e)(c-b)-a(a-e)(a-b)} &=\\frac{(a-b)+(c-b)\\frac{(a-b)[c(e-c)+a(e-a)]}{c(c-e)(c-b)-a(a-e)(a-b)}}{(a-e)}\\\\ &=\\left(\\frac{(a-b)[c(e-c)+a(e-a)]}{c(c-e)(c-b)-a(a-e)(a-b)}\\right)^2\\overline{\\left(\\frac{c(c-e)(c-b)-a(a-e)(a-b)}{(a-b)[c(e-c)+a(e-a)]}\\right)}\\\\ &=\\left(\\frac{(a-b)[c(e-c)+a(e-a)]}{c(c-e)(c-b)-a(a-e)(a-b)}\\right)^2\\frac{\\frac{1}{c}\\left(\\overline{e}-\\frac{1}{c}\\right)\\frac{b-c}{bc}-\\frac{1}{a}\\left(\\overline{e}-\\frac{1}{a}\\right)\\frac{b-a}{ba}}{\\frac{b-a}{ab}\\left(\\frac{1}{c}\\left(\\frac{1}{c}-\\overline{e}\\right)+\\frac{1}{a}\\left(\\frac{1}{a}-\\overline{e}\\right)\\right)}\\\\ &=\\left(\\frac{(a-b)[c(e-c)+a(e-a)]}{c(c-e)(c-b)-a(a-e)(a-b)}\\right)^2\\frac{c^3(a\\overline{e}-1)(a-b)-a^3(c\\overline{e}-1)(c-b)}{c(a-b)[a^2(c\\overline{e}-1)+c^2(a\\overline{e}-1)]}\\\\ &=\\frac{(a-b)[c(e-c)+a(e-a)]^2}{[c(c-e)(c-b)-a(a-e)(a-b)]^2}\\frac{c^3(a\\overline{e}-1)(a-b)-a^3(c\\overline{e}-1)(c-b)}{c[a^2(c\\overline{e}-1)+c^2(a\\overline{e}-1)]}. \\end{align*}\n\\]\n\nSimplifying, this becomes\n\n\\[\n\\begin{align*} &ac(2b-a-c)[c(c-e)(c-b)-a(a-e)(a-b)][a^2(c\\overline{e}-1)+c^2(a\\overline{e}-1)]\\\\ &=[c(e-c)+a(e-a)]^2[c^3(a\\overline{e}-1)(a-b)-a^3(c\\overline{e}-1)(c-b)]. \\end{align*}\n\\]\n\nOf course, we can also \"conjugate\" this equation -- a nice way to do this is to note that if\n\n\\[\nx=\\frac{(a-b)[c(e-c)+a(e-a)]}{c(c-e)(c-b)-a(a-e)(a-b)},\n\\]\n\nthen\n\n\\[\n\\frac{a(2b-a-c)}{c(e-c)+a(e-a)}=\\frac{x}{\\overline{x}}=\\overline{\\left(\\frac{c(e-c)+a(e-a)}{a(2b-a-c)}\\right)}=\\frac{\\frac{1}{c}\\left(\\overline{e}-\\frac{1}{c}\\right)+\\frac{1}{a}\\left(\\overline{e}-\\frac{1}{a}\\right)}{\\frac{1}{a}\\left(\\frac{2}{b}-\\frac{1}{a}-\\frac{1}{c}\\right)},\n\\]\n\nwhence\n\n\\[\nac(2b-a-c)[2ac-b(a+c)]=b[c(e-c)+a(e-a)][a^2(c\\overline{e}-1)+c^2(a\\overline{e}-1)].\n\\]\n\nIf \\(a+c\\ne 0\\), then eliminating \\(\\overline{e}\\), we get\n\n\\[\ne\\in\\left\\{a+c-\\frac{ac}{b},a+\\frac{2c(c-b)}{a+c},c+\\frac{2a(a-b)}{a+c}\\right\\}.\n\\]\n\nThe first case corresponds to (*) (since \\(a,b,c,e\\) uniquely determine \\(d\\) and \\(f\\)), the second corresponds to \\(AB\\parallel DE\\) (or equivalently, since \\(AB=DE\\), \\(a-b=e-d\\)), and by symmetry, the third corresponds to \\(CB\\parallel FE\\).\n\nOtherwise, if \\(c=-a\\), then we easily find \\(b^2e=a^4\\overline{e}\\) from the first of the two equations in \\(e,\\overline{e}\\) (we actually don't need this, but it tells us that the locus of working \\(e\\) is a line through the origin). It's easy to compute \\(d=e+\\frac{a(a-b)}{b}\\) and \\(f=e+\\frac{a(a+b)}{b}\\), so \\(a-c=2a=f-d\\implies c-d=a-f\\implies CD\\parallel AF\\), and we're done.\n\nComment. It appears that taking \\((ABC)\\) the unit circle is nicer than, say \\(e=0\\) or \\((ACE)\\) the unit circle (which may not even be reasonably tractable).\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2011-4
https://artofproblemsolving.com/wiki/index.php/2011_USAMO_Problems/Problem_4
Consider the assertion that for each positive integer \(n \ge 2\), the remainder upon dividing \(2^{2^n}\) by \(2^n-1\) is a power of 4. Either prove the assertion or find (with proof) a counter-example.
[ "We will show that \\(n = 25\\) is a counter-example.\n\nSince \\(\\textstyle 2^n \\equiv 1 \\pmod{2^n - 1}\\), we see that for any integer \\(k\\), \\(\\textstyle 2^{2^n} \\equiv 2^{(2^n - kn)} \\pmod{2^n-1}\\). Let \\(0 \\le m < n\\) be the residue of \\(2^n \\pmod n\\). Note that since \\(\\textstyle m < n\\) and \\(\\textstyle n \\ge 2\\), necessarily \\(\\textstyle 2^m < 2^n -1\\), and thus the remainder in question is \\(\\textstyle 2^m\\). We want to show that \\(\\textstyle 2^m \\pmod {2^n-1}\\) is an odd power of 2 for some \\(\\textstyle n\\), and thus not a power of 4.\n\nLet \\(\\textstyle n=p^2\\) for some odd prime \\(\\textstyle p\\). Then \\(\\textstyle \\varphi(p^2) = p^2 - p\\). Since 2 is co-prime to \\(\\textstyle p^2\\), we have\n\n\\[\n{2^{\\varphi(p^2)} \\equiv 1 \\pmod{p^2}}\n\\]\n\nand thus\n\n\\[\n\\textstyle 2^{p^2} \\equiv 2^{(p^2 - p) + p} \\equiv 2^p \\pmod{p^2}.\n\\]\n\nTherefore, for a counter-example, it suffices that \\(\\textstyle 2^p \\pmod{p^2}\\) be odd. Choosing \\(\\textstyle p=5\\), we have \\(\\textstyle 2^5 = 32 \\equiv 7 \\pmod{25}\\). Therefore, \\(\\textstyle 2^{25} \\equiv 7 \\pmod{25}\\) and thus\n\n\\[\n\\textstyle 2^{2^{25}} \\equiv 2^7 \\pmod {2^{25} - 1}.\n\\]\n\nSince \\(\\textstyle 2^7\\) is not a power of 4, we are done.", "Lemma (useful for all situations): If \\(x\\) and \\(y\\) are positive integers such that \\(2^x - 1\\) divides \\(2^y - 1\\), then \\(x\\) divides \\(y\\). Proof: \\(2^y \\equiv 1 \\pmod{2^x - 1}\\). Replacing the \\(1\\) with a \\(2^x\\) and dividing out the powers of two should create an easy induction proof which will be left to the reader as an Exercise.\n\nConsider \\(n = 25\\). We will prove that this case is a counterexample via contradiction.\n\nBecause \\(4 = 2^2\\), we will assume there exists a positive integer \\(k\\) such that \\(2^{2^n} - 2^{2k}\\) divides \\(2^n - 1\\) and \\(2^{2k} < 2^n - 1\\). Dividing the powers of \\(2\\) from LHS gives \\(2^{2^n - 2k} - 1\\) divides \\(2^n - 1\\). Hence, \\(2^n - 2k\\) divides \\(n\\). Because \\(n = 25\\) is odd, \\(2^{24} - k\\) divides \\(25\\). Euler's theorem gives \\(2^{24} \\equiv 2^4 \\equiv 16 \\pmod{25}\\) and so \\(k \\ge 16\\). However, \\(2^{2k} \\geq 2^{32} > 2^{25} - 1\\), a contradiction. Thus, \\(n = 25\\) is a valid counterexample.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2011-5
https://artofproblemsolving.com/wiki/index.php/2011_USAMO_Problems/Problem_5
Let \(P\) be a given point inside quadrilateral \(ABCD\). Points \(Q_1\) and \(Q_2\) are located within \(ABCD\) such that \(\angle Q_1 BC = \angle ABP\), \(\angle Q_1 CB = \angle DCP\), \(\angle Q_2 AD = \angle BAP\), \(\angle Q_2 DA = \angle CDP\). Prove that \(\overline{Q_1 Q_2} \parallel \overline{AB}\) if and only if \(\overline{Q_1 Q_2} \parallel \overline{CD}\).
[ "Lemma. If \\(AB\\) and \\(CD\\) are not parallel, then \\(AB, CD, Q_1 Q_2\\) are concurrent.\n\nProof. Let \\(AB\\) and \\(CD\\) meet at \\(R\\). Notice that with respect to triangle \\(ADR\\), \\(P\\) and \\(Q_2\\) are isogonal conjugates (this can be proven by dropping altitudes from \\(Q_2\\) to \\(AB\\), \\(CD\\), and \\(AD\\) or \\(BC\\) depending on where \\(R\\) is). With respect to triangle \\(BCR\\), \\(P\\) and \\(Q_1\\) are isogonal conjugates. Therefore, \\(Q_1\\) and \\(Q_2\\) lie on the reflection of \\(RP\\) in the angle bisector of \\(\\angle{DRA}\\), so \\(R, Q_1, Q_2\\) are collinear. Hence, \\(AB, CD, Q_1 Q_2\\) are concurrent at \\(R\\).\n\nNow suppose \\(Q_1 Q_2 \\parallel AB\\) but \\(Q_1 Q_2\\) is not parallel to \\(CD\\). Then \\(AB\\) and \\(CD\\) are not parallel and thus intersect at a point \\(R\\). But then \\(Q_1 Q_2\\) also passes through \\(R\\), contradicting \\(Q_1 Q_2 \\parallel AB\\). A similar contradiction occurs if \\(Q_1 Q_2 \\parallel CD\\) but \\(Q_1 Q_2\\) is not parallel to \\(AB\\), so we can conclude that \\(Q_1 Q_2 \\parallel AB\\) if and only if \\(Q_1 Q_2 \\parallel CD\\).", "First note that \\(\\overline{Q_1 Q_2} \\parallel \\overline{AB}\\) if and only if the altitudes from \\(Q_1\\) and \\(Q_2\\) to \\(\\overline{AB}\\) are the same, or \\(|Q_1B|\\sin \\angle ABQ_1 =|Q_2A|\\sin \\angle BAQ_2\\). Similarly \\(\\overline{Q_1 Q_2} \\parallel \\overline{CD}\\) iff \\(|Q_1C|\\sin \\angle DCQ_1 =|Q_2D|\\sin \\angle CDQ_2\\).\n\nIf we define \\(S =\\frac{|Q_1B|\\sin \\angle ABQ_1}{|Q_2A|\\sin \\angle BAQ_2} \\times \\frac{|Q_2D|\\sin \\angle CDQ_2}{|Q_1C|\\sin \\angle DCQ_1}\\), then we are done if we can show that S=1.\n\nBy the law of sines, \\(\\frac{|Q_1B|}{|Q_1C|}=\\frac{\\sin\\angle Q_1CB}{\\sin\\angle Q_1BC}\\) and \\(\\frac{|Q_2D|}{|Q_2A|}=\\frac{\\sin\\angle Q_2AD}{\\sin\\angle Q_2DA}\\).\n\nSo, \\(S=\\frac{\\sin \\angle ABQ_1}{\\sin \\angle BAQ_2}\\cdot\\frac{\\sin \\angle CDQ_2}{\\sin \\angle DCQ_1}\\cdot\\frac{\\sin \\angle BCQ_1}{\\sin \\angle CBQ_1}\\cdot\\frac{\\sin \\angle DAQ_2}{\\sin \\angle ADQ_2}\\)\n\nBy the terms of the problem, \\(S=\\frac{\\sin \\angle PBC}{\\sin \\angle PAD}\\cdot\\frac{\\sin \\angle PDA}{\\sin \\angle PCB}\\cdot\\frac{\\sin \\angle PCD}{\\sin \\angle PBA}\\cdot\\frac{\\sin \\angle PAB}{\\sin \\angle PDC}\\). (If two subangles of an angle of the quadrilateral are equal, then their complements at that quadrilateral angle are equal as well.)\n\nRearranging yields \\(S= \\frac{\\sin \\angle PBC}{\\sin \\angle PCB}\\cdot\\frac{\\sin \\angle PDA}{\\sin \\angle PAD}\\cdot\\frac{\\sin \\angle PCD}{\\sin \\angle PDC}\\cdot\\frac{\\sin \\angle PAB}{\\sin \\angle PBA}\\).\n\nApplying the law of sines to the triangles with vertices at P yields \\(S=\\frac{|PC|}{|PB|}\\frac{|PA|}{|PD|}\\frac{|PD|}{|PC|}\\frac{|PB|}{|PA|}=1\\).", "Case 1 The lines \\(AB\\) and \\(CD\\) are not parallel. Denote \\(E = AB \\cap CD.\\)\n\n\\[\n\\angle Q_1 BC = \\angle ABP, \\angle Q_1 CB = \\angle DCP \\implies\n\\]\n\nPoint \\(Q_1\\) isogonal conjugate of a point \\(P\\) with respect to a triangle \\(\\triangle EBC \\implies\\)\n\n\\(EP\\) and \\(EQ_1\\) are isogonals with respect to \\(\\angle BEC.\\)\n\nSimilarly point \\(Q_2\\) isogonal conjugate of a point \\(P\\) with respect to a triangle \\(\\triangle EAD \\implies\\)\n\n\\(EP\\) and \\(EQ_2\\) are isogonals with respect to \\(\\angle BEC.\\)\n\nTherefore points \\(E, Q_1, Q_2\\) lies on the isogonal \\(EP\\) with respect to \\(\\angle BEC \\implies\\)\n\n\\(Q_1Q_2\\) is not parallel to \\(AB\\) or \\(CD.\\)\n\nCase 2 \\(AB||CD.\\) We use The isogonal theorem in case parallel lines and get\n\n\\[\nAB||Q_1Q_2 || CD. \\blacksquare\n\\]\n\nvladimir.shelomovskii@gmail.com, vvsss" ]
USAMO-2011-6
https://artofproblemsolving.com/wiki/index.php/2011_USAMO_Problems/Problem_6
Let \(A\) be a set with \(|A| = 225\), meaning that \(A\) has 225 elements. Suppose further that there are eleven subsets \(A_1\), \(\dots\), \(A_{11}\) of \(A\) such that \(|A_i | = 45\) for \(1 \le i \le 11\) and \(|A_i \cap A_j| = 9\) for \(1 \le i < j \le 11\). Prove that \(|A_1 \cup A_2 \cup \dots \cup A_{11}| \ge 165\), and give an example for which equality holds.
[ "## Existence\n\nNote that \\(\\textstyle \\binom{11}3 = 165,\\) and so it is natural to consider placing one element in each intersection of three of the 11 sets. Since each pair of sets is in 9 3-way intersections—one with each of the 9 remaining sets—any two sets will have 9 elements in common. Since \\(\\textstyle \\binom{10}2 = 45,\\) each set is in 45 triples and thus will have 45 elements. We can now throw in 60 more elements outside the union of the \\(A_i\\) and we are done.\n\n## Minimality\n\nAs in the proof of PIE, let \\(I\\) be a finite set. Let \\(\\textstyle U = \\bigcup_{i\\in I} A_i\\). For \\(i\\in I\\), let \\(\\chi_i\\) be the characteristic function of \\(A_i\\), that is, for \\(\\alpha \\in U,\\)\n\n\\[\n\\chi_i(\\alpha) = \\begin{cases} 1 & \\alpha \\in A_i \\\\ 0 & \\alpha \\not\\in A_i \\end{cases}\n\\]\n\nFor each \\(\\alpha \\in U\\) let \\(\\textstyle n_\\alpha = \\sum_{i \\in I} \\chi_i(\\alpha)\\), that is the number of subsets \\(A_i\\) of which \\(\\alpha\\) is an element.\n\nIf \\(S \\subset I\\), let \\(\\textstyle A_S = \\bigcap_{i \\in S}A_i\\). Then the characteristic function of \\(A_S\\) is \\(\\textstyle \\prod_{i \\in S} \\chi_i\\). The number of elements of \\(A_S\\) is simply the sum of its characteristic function over all the elements of \\(U\\):\n\n\\[\n|A_S| = \\sum_{\\alpha \\in U} \\prod_{i \\in S} \\chi_i(\\alpha).\n\\]\n\nFor \\(0 \\le k \\le |I|\\), consider the sum \\(N_k\\) of \\(|A_S|\\) over all \\(S\\subset I\\) with \\(|S|= k\\). This is:\n\n\\[\nN_k = \\sum_{\\substack{S\\subset I\\\\|S| = k}} \\sum_{\\alpha \\in U} \\prod_{i \\in S} \\chi_i(\\alpha) = \\sum_{\\substack{S\\subset I\\\\|S| = k}} |A_S|\n\\]\n\nreversing the order summation, as an element \\(\\alpha\\) that appears in \\(n_\\alpha\\) of the \\(A_i\\), will appear in exactly \\(\\textstyle \\binom{n_\\alpha}{k}\\) intersections of \\(k\\) subsets, we get:\n\n\\[\n\\sum_{\\substack{S\\subset I\\\\|S| = k}} |A_S| = \\sum_{\\alpha \\in U} \\sum_{\\substack{S\\subset I\\\\|S| = k}} \\prod_{i \\in S} \\chi_i(\\alpha) = \\sum_{\\alpha \\in U} \\binom{n_\\alpha}k.\n\\]\n\nApplying the above with \\(I = \\{1, 2, \\ldots, 11\\},\\) and \\(k=1\\), since each of the 11 \\(A_i\\) has 45 elements, we get:\n\n\\[\nN_1 = \\sum_{\\alpha \\in U} n_\\alpha = 11 \\cdot 45,\n\\]\n\nand for \\(k = 2\\), since each of the 55 pairs \\(A_i \\cap A_j\\) has 9 elements, we get:\n\n\\[\nN_2 = \\sum_{\\alpha \\in U} \\binom{n_\\alpha}2 = 9 \\cdot 55 = 11 \\cdot 45.\n\\]\n\nTherefore\n\n\\[\n\\begin{align*} s_1 &= \\sum_{\\alpha \\in U} n_\\alpha = N_1 = 11\\cdot45 \\\\ s_2 &= \\sum_{\\alpha \\in U} n_\\alpha^2 = 2N_2 + N_1 = 3\\cdot11\\cdot45. \\end{align*}\n\\]\n\nLet \\(n = |U|\\) be the number of elements of \\(U\\). Since for any set of real numbers the mean value of the squares is greater than or equal to the square of the mean value, we have:\n\n\\[\n\\frac{s_2}n \\ge \\left(\\frac{s_1}n\\right)^2 \\implies n \\ge \\frac{s_1^2}{s_2} = \\frac{11^2\\cdot 45^2}{3\\cdot11\\cdot45} = \\frac{11\\cdot45}3 = 165.\n\\]\n\nThus \\(|U| \\ge 165\\) as required.", "We will count the number of ordered triples, \\((A_i,A_j,x)\\), where \\(1\\le i,j\\le11\\) and \\(x\\in A_i \\cap A_j\\). We know this is equal to \\(990=11\\cdot10\\cdot9\\)­. We can also find that this is \\(\\sum_{k=1}^{225}b_k^2-b_k\\), where \\(b_k\\) is the number of the \\(11\\) subsets the \\(k^{\\text{th}}\\) element of \\(A\\) is in. Since \\(\\sum_{k=1}^{225}b_k=45\\cdot11=495\\), we know \\(\\sum_{k=1}^{225}b_k^2=990+495=1485\\). Let \\(n=|A_1 \\cup A_2 \\cup \\dots \\cup A_{11}|\\). \\(n\\) is equal to the number of \\(b_k>0\\), where \\(1\\le k\\le225\\). By the QM-AM inequality, we know \\(\\frac{\\sum_{k=1}^{225}b_k^2}n\\ge\\Bigg(\\frac{\\sum_{k=1}^{225}b_k}n\\Bigg)^2\\implies\\frac{1485}n\\ge\\Big(\\frac{495}n\\Big)^2\\implies n\\ge165\\) and that equality occurs when \\(b_1=b_2=b_3=\\cdots=b_{165}=3,b_{166}=b_{167}=\\cdots=b_{225}=0\\). \\(\\square\\)\n\nSolution by randomdude10807", "Define \\(x_i\\) to be the amount of elements in \\(A\\) such that the element is contained exactly \\(i\\) times among the \\(11\\) subsets \\(A_j\\). We are trying to prove that \\(\\sum_{i=1}^{11}x_i\\ge 165\\). Now, note that \\(\\sum_{i=1}^{11}ix_i\\) is equivalent to the total amount of not necessarily distinct elements among the subsets \\(A_j\\), thus \\(\\sum_{i=1}^{11}ix_i=11|A_j|=495\\). Furthermore, note that \\(\\sum_{i=1}^{11}\\binom{i}{2}x_i\\) is equivalent to the total amount of not necessarily distinct pairs of subsets such that the subsets share an element. In other words, each subset pair should be counted \\(9\\) times in this sum, and there are \\(\\binom{11}{2}\\) total distinct pairs of subsets, so \\(\\sum_{i=1}^{11}\\binom{i}{2}x_i=9*\\binom{11}{2}=495\\). Noting that \\(2\\binom{i}{2}+i=i^2\\) for all nonnegative integers, we note\n\n\\[\n\\sum_{i=1}^{11}i^2x_i=2(\\sum_{i=1}^{11}\\binom{i}{2}x_i)+(\\sum_{i=1}^{11}ix_i)=495*2+495=495*3\n\\]\n\nFinally, CS inequality gives us\n\n\\[\n(\\sum_{i=1}^{11}i^2x_i)(\\sum_{i=1}^{11}x_i)\\ge(\\sum_{i=1}^{11}ix_i)^2\n\\]\n\n\\[\n(495*3)(\\sum_{i=1}^{11}x_i)\\ge(495^2)\n\\]\n\n\\[\n\\boxed{(\\sum_{i=1}^{11}x_i)\\ge 165}\n\\]\n\nas desired.\n\nOne equality case occurs when \\(x_i=0\\) if \\(i\\) is not \\(3\\), and \\(x_3=165\\). Choose any \\(165\\) elements, and represent each with a unique combination of \\(3\\) of the \\(11\\) subsets (\\(\\binom{11}{3}=165\\)). It is then obvious that each subset has \\(45\\) distinct elements, because after each subset is chosen, there are \\(\\binom{10}{2}=45\\) ways to choose the other two subsets to produce a unique element. Similarly, each pair of subsets shares exactly \\(9\\) elements, because after choosing the two subsets, we can choose one of the remaining \\(9\\) subsets to produce the unique element. Hence, proved. \\(\\blacksquare{}\\)\n\n~SigmaPiE\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2012-1
https://artofproblemsolving.com/wiki/index.php/2012_USAMO_Problems/Problem_1
Find all integers \(n \ge 3\) such that among any \(n\) positive real numbers \(a_1\), \(a_2\), \(\dots\), \(a_n\) with \[ \max(a_1, a_2, \dots, a_n) \le n \cdot \min(a_1, a_2, \dots, a_n), \] there exist three that are the side lengths of an acute triangle.
[ "Without loss of generality, assume that the set \\(\\{a\\}\\) is ordered from least to greatest so that the bounding condition becomes \\(a_n \\le n \\cdot a_1.\\) Now set \\(b_i \\equiv \\frac{a_i}{a_1},\\) and since a triangle with sidelengths from \\(\\{a\\}\\) will be similar to the corresponding triangle from \\(\\{b\\},\\) we simply have to show the existence of acute triangles in \\(\\{b\\}.\\) Note that \\(b_1 = 1\\) and for all \\(i\\), \\(b_i \\le n.\\)\n\nNow three arbitrary sidelengths \\(x\\), \\(y\\), and \\(z\\), with \\(x \\le y \\le z,\\) will form a valid triangle if and only if \\(x+y>z.\\) Furthermore, this triangle will be acute if and only if \\(x^2 + y^2 > z^2.\\) However, the first inequality can actually be inferred from the second, since \\(x+y>z \\longleftrightarrow x^2 + y^2 +2xy > z^2\\) and \\(2xy\\) is trivially greater than \\(0.\\) So we just need to find all \\(n\\) such that there is necessarily a triplet of \\(b\\)'s for which \\(b_i^2 + b_j^2 > b_k^2\\) (where \\(b_i < b_j < b_k\\)).\n\nWe now make another substitution: \\(c_i \\equiv b_i ^2.\\) So \\(c_1 = 1\\) and for all \\(i\\), \\(c_i \\le n^2.\\) Now we examine the smallest possible sets \\(\\{c\\}\\) for small \\(n\\) for which the conditions of the problem are not met. Note that by smallest, we mean the set whose greatest element is as small as possible. If \\(n=3\\), then the smallest possible set, call it \\(\\{s_3\\},\\) is trivially \\(\\{1,1,2\\}\\), since \\(c_1\\) and \\(c_2\\) are obviously minimized and \\(c_3\\) follows as minimal. Using this as the base case, we see inductively that in general \\(\\{s_n\\}\\) is the set of the first \\(n\\) Fibonacci numbers. To show this note that if \\(\\{s_n\\} = \\{F_0, F_1, ... F_n\\}\\), then \\(\\{s_{n+1}\\} = \\{F_0, F_1, ... F_n, c_{n+1}\\}.\\) The smallest possible value for \\(c_{n+1}\\) is the sum of the two greatest values of \\(\\{s_n\\}\\) which are \\(F_{n-1}\\) and \\(F_n\\). But these sum to \\(F_{n+1}\\) so \\(\\{s_{n+1}\\} = \\{F_0, F_1, ... F_{n+1}\\}\\) and our induction is complete.\n\nNow since we know that the Fibonacci set is the smallest possible set which does not satisfy the conditions of the problem, then any set \\(\\{c\\}\\) whose greatest term is less than \\(F_{n-1}\\) must satisfy the conditions. And since \\(\\{c\\}\\) is bounded between \\(1\\) and \\(n^2\\), then the conditions of the problem are met if and only if \\(F_{n-1} > n^2\\). The first \\(n\\) for which this restriction is satisfied is \\(n=13\\) and the exponential behavior of the Fibonacci numbers ensure that every \\(n\\) greater than \\(13\\) will also satisfy this restriction. So the final solution set is \\(\\boxed{\\{n \\ge 13\\}}\\).", "Outline:\n\n1. Define the Fibonacci numbers to be \\(F_1 = F_2 = 1\\) and \\(F_k = F_{k-1} + F_{k-2}\\) for \\(k \\ge 3\\).\n\n2. If the chosen \\(n\\) is such that \\(F_n \\le n^2\\), then choose the sequence \\(a_n\\) such that \\(a_k = \\sqrt{F_k}\\) for \\(1 \\le k \\le n\\). It is easy to verify that such a sequence satisfies the condition that the largest term is less than or equal to \\(n\\) times the smallest term. Also, because for any three terms \\(x = \\sqrt{F_a}, y = \\sqrt{F_b}, z = \\sqrt{F_c}\\) with \\(a<b<c\\), \\(x^2 + y^2 = F_a + F_b \\le F_{b-1} + F_b = F_{b+1} \\le F_c = z^2\\), x, y, z do not form an acute triangle. Thus, all \\(n\\) such that \\(F_n \\le n^2\\) do not work.\n\n3. It is easy to observe via a contradiction argument that all \\(n\\) such that \\(F_n > n^2\\) produce an acute triangle. (If, without loss of generality, \\(a_n\\) is an increasing sequence, such that no three (in particular, consecutive) terms form an acute triangle, then \\(a_2^2 \\ge F_1a_1^2, a_3^2 \\ge a_2^2 + a_1^2 \\ge F_2a^2\\), and by induction \\(a_n^2 > F_na_1^2\\), a contradiction to the condition's inequality.)\n\n4. Note that \\(F_{12} = 144 = 12^2\\) and \\(F_{13} = 233 > 169 = 13^2\\). It is easily to verify through strong induction that all \\(n\\) greater than 12 make \\(F_n > n^2\\). Thus, \\(\\boxed{n \\ge 13}\\) is the desired solution set." ]
USAMO-2012-2
https://artofproblemsolving.com/wiki/index.php/2012_USAMO_Problems/Problem_2
A circle is divided into 432 congruent arcs by 432 points. The points are colored in four colors such that some 108 points are colored Red, some 108 points are colored Green, some 108 points are colored Blue, and the remaining 108 points are colored Yellow. Prove that one can choose three points of each color in such a way that the four triangles formed by the chosen points of the same color are congruent.
[ "If you rotate the red points 431 times, they will overlap with blue points \\(108\\times 108\\) times, for an average of \\(\\frac{108\\times 108}{431}\\) per rotation. Note that this average is slightly greater than 27. Therefore at some point 28 red points overlap with blue points. In other words, there exist 28 red and blue points such that the convex 28-gons formed by them are congruent.\n\nRotate these 28 red points 431 times. They will overlap with green points \\(108\\times 28\\) times, for an average of \\(\\frac{108\\times 28}{431}\\) per rotation. This average is slightly greater than 7, so at some point 8 of those red points overlap with green points. In other words, there exist 8 red points, 8 blue points, and 8 green points such that the convex octagons formed by them are congruent.\n\nRotate these 8 red points 431 times. They will overlap with yellow points \\(108\\times 8\\) times, for an average of \\(\\frac{108\\cdot 8}{431}\\) per rotation. This average is slightly greater than 2, so at some point 3 of those red points will overlap with yellow points. In other words, there exist 3 red points, 3 blue points, 3 green points, and 3 yellow points such that the triangles formed by them are congruent.\n\n## Note\n\nHuge bash of Pigeonhole Principle:\n\n\\[\n108\\longrightarrow 28\\longrightarrow 8\\longrightarrow 3\n\\]\n\n\\[\n\\left \\lceil{\\left \\lceil{\\left \\lceil{108\\cdot \\frac{108}{431}}\\right \\rceil \\cdot \\frac{108}{431}}\\right \\rceil \\cdot \\frac{108}{431}}\\right \\rceil =\\left \\lceil{\\left \\lceil{28\\cdot \\frac{108}{431}}\\right \\rceil \\cdot \\frac{108}{431}}\\right \\rceil =\\left \\lceil{8\\cdot \\frac{108}{431}}\\right \\rceil =3\n\\]" ]
USAMO-2012-3
https://artofproblemsolving.com/wiki/index.php/2012_USAMO_Problems/Problem_3
Determine which integers \(n > 1\) have the property that there exists an infinite sequence \(a_1\), \(a_2\), \(a_3\), \(\dots\) of nonzero integers such that the equality \[ a_k + 2a_{2k} + \dots + na_{nk} = 0 \] holds for every positive integer \(k\).
[ "For \\(n\\) equal to any odd prime \\(p\\), the sequence \\(\\left\\{a_i = \\left(\\frac{1-n}{2}\\right)^{m_p\\left(i\\right)}\\right\\}\\), where \\(p^{m_p\\left(i\\right)}\\) is the greatest power of \\(p\\) that divides \\(i\\), gives a valid sequence. Therefore, the set of possible values for \\(n\\) is at least the set of odd primes.", "(Since Bertrand's is well known and provable using elementary techniques, I see nothing wrong with this-tigershark22)\n\nFor \\(n=2\\), \\(|a_1| = 2 |a_2| = \\cdots = 2^m |a_{2^m}|\\) implies that for any positive integer \\(m\\), \\(|a_1| \\ge 2^m\\), which is impossible.\n\nWe proceed to prove that the infinite sequence exists for all \\(n\\ge 3\\).\n\nFirst, one notices that if we have \\(a_{xy} = a_x a_y\\) for any integers \\(x\\) and \\(y\\), then it is suffice to define all \\(a_x\\) for \\(x\\) prime, and one only needs to verify the equation (*)\n\n\\[\na_1+2a_2+\\cdots+na_n=0\n\\]\n\nfor the other equations will be automatically true.\n\nTo proceed with the construction, I need the following fact: for any positive integer \\(m>2\\), there exists a prime \\(p\\) such that \\(\\frac{m}{2} <p \\le m\\).\n\nTo prove this, I am going to use Bertrand's Postulate ([1]) without proof. The Theorem states that, for any integer \\(n>1\\), there exists a prime \\(p\\) such that \\(n<p\\le 2n-1\\). In other words, for any positive integer \\(m>2\\), if \\(m=2n\\) with \\(n>1\\), then there exists a prime \\(p\\) such that \\(\\frac{m}{2} < p < m\\), and if \\(m=2n-1\\) with \\(n>1\\), then there exists a prime \\(p\\) such that \\(\\frac{m+1}{2} <p\\le m\\), both of which guarantees that for any integer \\(m>2\\), there exists a prime \\(p\\) such that \\(\\frac{m}{2} <p \\le m\\).\n\nGo back to the problem. Suppose \\(n\\ge 3\\). Let the largest two primes not larger than \\(n\\) are \\(P\\) and \\(Q\\), and that \\(n\\ge P > Q\\). By the fact stated above, one can conclude that \\(2P > n\\), and that \\(4Q = 2(2Q) \\ge 2P > n\\). Let's construct \\(a_n\\):\n\nLet \\(a_1=1\\). There will be three cases: (i) \\(Q>\\frac{n}{2}\\), (ii) \\(\\frac{n}{2} \\ge Q > \\frac{n}{3}\\), and (iii) \\(\\frac{n}{3} \\ge Q > \\frac{n}{4}\\).\n\nCase (i): \\(2Q>n\\). Let \\(a_x = 1\\) for all prime numbers \\(x<Q\\), and \\(a_{xy}=a_xa_y\\), then (*) becomes:\n\n\\[\nPa_P + Qa_Q = C_1\n\\]\n\nCase (ii): \\(2Q\\le n\\) but \\(3Q > n\\). In this case, let \\(a_2=-1\\), and \\(a_x = 1\\) for all prime numbers \\(2<x<Q\\), and \\(a_{xy}=a_xa_y\\), then (*) becomes:\n\n\\[\nPa_P + Qa_Q - Qa_{2Q} = C_2\n\\]\n\nor\n\n\\[\nPa_P - Qa_Q = C_2\n\\]\n\nCase (iii): \\(3Q\\le n\\). In this case, let \\(a_2=3\\), \\(a_3=-2\\), and \\(a_x = 1\\) for all prime numbers \\(3<x<Q\\), and \\(a_{xy}=a_xa_y\\), then (*) becomes:\n\n\\[\nPa_P + Qa_Q + 3Qa_{2Q} - 2Qa_{3Q} = C_3\n\\]\n\nor\n\n\\[\nPa_P + Qa_Q = C_3\n\\]\n\nIn each case, by Bezout's Theorem, there exists non zero integers \\(a_P\\) and \\(a_Q\\) which satisfy the equation. For all other primes \\(p > P\\), just let \\(a_p=1\\) (or any other non-zero integer).\n\nThis construction is correct because, for any \\(k> 1\\),\n\n\\[\na_k + 2 a_{2k} + \\cdots n a_{nk} = a_k (1 + 2 a_2 + \\cdots n a_n ) = 0\n\\]\n\nSince Bertrand's Theorem is not elementary, we still need to wait for a better proof.\n\n--Lightest 21:24, 2 May 2012 (EDT)" ]
USAMO-2012-4
https://artofproblemsolving.com/wiki/index.php/2012_USAMO_Problems/Problem_4
Find all functions \(f : \mathbb{Z}^+ \to \mathbb{Z}^+\) (where \(\mathbb{Z}^+\) is the set of positive integers) such that \(f(n!) = f(n)!\) for all positive integers \(n\) and such that \(m - n\) divides \(f(m) - f(n)\) for all distinct positive integers \(m\), \(n\).
[ "By the first condition we have \\(f(1)=f(1!)=f(1)!\\) and \\(f(2)=f(2!)=f(2)!\\), so \\(f(1)=1\\) or \\(2\\) and similarly for \\(f(2)\\). By the second condition, we have\n\n\\[\nn\\cdot n!=(n+1)!-n! \\mid f(n+1)!-f(n)! \\qquad \\qquad (1)\n\\]\n\nfor all positive integers \\(n\\).\n\nSuppose that for some \\(n \\geq 2\\) we have \\(f(n) = 1\\). We claim that \\(f(k)=1\\) for all \\(k\\ge n\\). Indeed, from Equation (1) we have \\(f(n+1)!\\equiv 1 \\mod n\\cdot n!\\), and this is only possible if \\(f(n+1)=1\\); the claim follows by induction.\n\nWe now divide into cases:\n\nCase 1: \\(f(1)=f(2)=1\\)\n\nThis gives \\(f(n)=1\\) always from the previous claim, which is a solution.\n\nCase 2: \\(f(1)=2, f(2)=1\\)\n\nThis implies \\(f(n)=1\\) for all \\(n\\ge 2\\), but this does not satisfy the initial conditions. Indeed, we would have\n\n\\[\n3-1 \\mid f(3)-f(1)\n\\]\n\nand so \\(2\\mid -1\\), a contradiction.\n\nCase 3: \\(f(1)=1\\), \\(f(2)=2\\)\n\nWe claim \\(f(n)=n\\) always by induction. The base cases are \\(n = 1\\) and \\(n = 2\\). Fix \\(k > 1\\) and suppose that \\(f(k)=k\\). By Equation (1) we have that\n\n\\[\nf(k+1)! \\equiv k! \\mod k\\cdot k! .\n\\]\n\nThis implies \\(f(k+1)<2k\\) (otherwise \\(f(k+1)!\\equiv 0 \\mod k\\cdot k!\\)). Also we have\n\n\\[\n(k+1)-1 \\mid f(k+1)-f(1)\n\\]\n\nso \\(f(k+1)\\equiv 1 \\mod k\\). This gives the solutions \\(f(k+1)=1\\) and \\(f(k+1)=k+1\\). The first case is obviously impossible, so \\(f(k + 1) = k + 1\\), as desired. By induction, \\(f(n) = n\\) for all \\(n\\). This also satisfies the requirements.\n\nCase 4: \\(f(1)=f(2)=2\\)\n\nWe claim \\(f(n)=2\\) by a similar induction. Again if \\(f(k)=2\\), then by (1) we have\n\n\\[\nf(k+1)\\equiv 2 \\mod k\\cdot k!\n\\]\n\nand so \\(f(k+1)<2k\\). Also note that\n\n\\[\nk+1-1 \\mid f(k+1)-2\n\\]\n\nand\n\n\\[\nk+1-2 \\mid f(k+1)-2\n\\]\n\nso \\(f(k+1)\\equiv 2 \\mod k(k-1)\\). Then the only possible solution is \\(f(k+1)=2\\). By induction, \\(f(n) = 2\\) for all \\(n\\), and this satisfies all requirements.\n\nIn summary, there are three solutions: \\(\\boxed{f(n)=1, f(n)=2, f(n)=n}\\).\n\n## Summary\n\nThree functions: \\(f(x)=x\\) since \\(x!=x!\\), \\(1!=1\\) and \\(2!=2\\), fixed points on the \\(x!\\) function.\n\nNo elegant\\(\\mod\\) argument needed; \\(f(m)-f(n)\\textcolor{red}{\\textbf{ = }}m-n\\) so of course \\(m-n\\mid f(m)-f(n)\\)!" ]
USAMO-2012-5
https://artofproblemsolving.com/wiki/index.php/2012_USAMO_Problems/Problem_5
Let \(P\) be a point in the plane of triangle \(ABC\), and \(\gamma\) a line passing through \(P\). Let \(A'\), \(B'\), \(C'\) be the points where the reflections of lines \(PA\), \(PB\), \(PC\) with respect to \(\gamma\) intersect lines \(BC\), \(AC\), \(AB\), respectively. Prove that \(A'\), \(B'\), \(C'\) are collinear.
[ "By the sine law on triangle \\(AB'P\\),\n\n\\[\n\\frac{AB'}{\\sin \\angle APB'} = \\frac{AP}{\\sin \\angle AB'P},\n\\]\n\nso\n\n\\[\nAB' = AP \\cdot \\frac{\\sin \\angle APB'}{\\sin \\angle AB'P}.\n\\]\n\n\\[\n[asy] import graph; import geometry; unitsize(0.5 cm); pair[] A, B, C; pair P, R; A[0] = (2,12); B[0] = (0,0); C[0] = (14,0); P = (4,5); R = 5*dir(70); A[1] = extension(B[0],C[0],P,reflect(P + R,P - R)*(A[0])); B[1] = extension(C[0],A[0],P,reflect(P + R,P - R)*(B[0])); C[1] = extension(A[0],B[0],P,reflect(P + R,P - R)*(C[0])); draw((P - R)--(P + R),red); draw(A[1]--B[1]--C[1]--cycle,blue); draw(A[0]--B[0]--C[0]--cycle); draw(A[0]--P); draw(B[0]--P); draw(C[0]--P); draw(P--A[1]); draw(P--B[1]); draw(P--C[1]); draw(A[1]--B[0]); draw(A[1]--B[0]); label(\"$A$\", A[0], N); label(\"$B$\", B[0], S); label(\"$C$\", C[0], SE); dot(\"$A'$\", A[1], SW); dot(\"$B'$\", B[1], NE); dot(\"$C'$\", C[1], W); dot(\"$P$\", P, SE); label(\"$\\gamma$\", P + R, N); [/asy]\n\\]\n\nSimilarly,\n\n\\[\n\\begin{align*} B'C &= CP \\cdot \\frac{\\sin \\angle CPB'}{\\sin \\angle CB'P}, \\\\ CA' &= CP \\cdot \\frac{\\sin \\angle CPA'}{\\sin \\angle CA'P}, \\\\ A'B &= BP \\cdot \\frac{\\sin \\angle BPA'}{\\sin \\angle BA'P}, \\\\ BC' &= BP \\cdot \\frac{\\sin \\angle BPC'}{\\sin \\angle BC'P}, \\\\ C'A &= AP \\cdot \\frac{\\sin \\angle APC'}{\\sin \\angle AC'P}. \\end{align*}\n\\]\n\nHence,\n\n\\[\n\\begin{align*} &\\frac{AB'}{B'C} \\cdot \\frac{CA'}{A'B} \\cdot \\frac{BC'}{C'A} \\\\ &= \\frac{\\sin \\angle APB'}{\\sin \\angle AB'P} \\cdot \\frac{\\sin \\angle CB'P}{\\sin \\angle CPB'} \\cdot \\frac{\\sin \\angle CPA'}{\\sin \\angle CA'P} \\cdot \\frac{\\sin \\angle BA'P}{\\sin \\angle BPA'} \\cdot \\frac{\\sin \\angle BPC'}{\\sin \\angle BC'P} \\cdot \\frac{\\sin \\angle AC'P}{\\sin \\angle APC'}. \\end{align*}\n\\]\n\nSince angles \\(\\angle AB'P\\) and \\(\\angle CB'P\\) are supplementary or equal, depending on the position of \\(B'\\) on \\(AC\\),\n\n\\[\n\\sin \\angle AB'P = \\sin \\angle CB'P.\n\\]\n\nSimilarly,\n\n\\[\n\\begin{align*} \\sin \\angle CA'P &= \\sin \\angle BA'P, \\\\ \\sin \\angle BC'P &= \\sin \\angle AC'P. \\end{align*}\n\\]\n\nBy the reflective property, \\(\\angle APB'\\) and \\(\\angle BPA'\\) are supplementary or equal, so\n\n\\[\n\\sin \\angle APB' = \\sin \\angle BPA'.\n\\]\n\nSimilarly,\n\n\\[\n\\begin{align*} \\sin \\angle CPA' &= \\sin \\angle APC', \\\\ \\sin \\angle BPC' &= \\sin \\angle CPB'. \\end{align*}\n\\]\n\nTherefore,\n\n\\[\n\\frac{AB'}{B'C} \\cdot \\frac{CA'}{A'B} \\cdot \\frac{BC'}{C'A} = 1,\n\\]\n\nso by Menelaus's theorem, \\(A'\\), \\(B'\\), and \\(C'\\) are collinear.", "We will perform barycentric coordinates on the triangle \\(PCC'\\), with \\(P=(1,0,0)\\), \\(C'=(0,1,0)\\), and \\(C=(0,0,1)\\). Set \\(a = CC'\\), \\(b = CP\\), \\(c = C'P\\) as usual. Since \\(A\\), \\(B\\), \\(C'\\) are collinear, we will define \\(A = (p : k : q)\\) and \\(B = (p : \\ell : q)\\).\n\nClaim: Line \\(\\gamma\\) is the angle bisector of \\(\\angle APA'\\), \\(\\angle BPB'\\), and \\(\\angle CPC'\\). This is proved by observing that since \\(A'P\\) is the reflection of \\(AP\\) across \\(\\gamma\\), etc.\n\nThus \\(B'\\) is the intersection of the isogonal of \\(B\\) with respect to \\(\\angle P\\) with the line \\(CA\\); that is,\n\n\\[\nB' = \\left( \\frac pk \\frac{b^2}{\\ell}: \\frac{b^2}{\\ell} : \\frac{c^2}{q} \\right).\n\\]\n\nAnalogously, \\(A'\\) is the intersection of the isogonal of \\(A\\) with respect to \\(\\angle P\\) with the line \\(CB\\); that is,\n\n\\[\nA' = \\left( \\frac{p}{\\ell} \\frac{b^2}{k} : \\frac{b^2}{k} : \\frac{c^2}{q} \\right).\n\\]\n\nThe ratio of the first to third coordinate in these two points is both \\(b^2pq : c^2k\\ell\\), so it follows \\(A'\\), \\(B'\\), and \\(C'\\) are collinear.\n\n~peppapig_", "We will use Cartesian coordinates. Without loss of generality, let \\(P = (0,0)\\) and let \\(\\gamma\\) be the line \\(x = 0\\). Let \\(A = (x_1,y_1)\\), \\(B = (x_2,y_2)\\), and \\(C = (x_3,y_3)\\). Then \\(A'\\) is the intersection of the lines \\begin{align*} y &= -\\frac{y_1}{x_1}x \\\\ y - y_2 &= \\frac{y_2 - y_3}{x_2 - x_3}(x - x_2). \\end{align*} Solving the system of equations, we see it is\n\n\\[\n\\left(\\frac{x_1(x_2y_3 - y_2x_3)}{x_1y_2 + y_1x_2 - x_1y_3 - x_3y_1},\\frac{y_1(y_2x_3 - x_2y_3)}{x_1y_2 + y_1x_2 - x_1y_3 - x_3y_1}\\right).\n\\]\n\nTo check collinearity, we need the determinant of the matrix\n\n\\[\n\\begin{bmatrix} x_1(x_2y_3 - y_2x_3) & y_1(y_2x_3 - x_2y_3) & x_1y_2 + y_1x_2 - x_1y_3 - x_3y_1 \\\\ x_2(x_3y_1 - y_3x_1) & y_2(y_3x_1 - x_3y_1) & x_2y_3 + y_2x_3 - x_2y_1 - x_1y_2 \\\\ x_3(x_1y_2 - y_1x_2) & y_3(y_1x_2 - x_1y_2) & x_3y_1 + y_3x_1 - x_3y_2 - x_2y_3 \\end{bmatrix}\n\\]\n\nto be zero, which is true because the columns sum to zero.\n\n~knowingant" ]
USAMO-2012-6
https://artofproblemsolving.com/wiki/index.php/2012_USAMO_Problems/Problem_6
For integer \(n \ge 2\), let \(x_1\), \(x_2\), \(\dots\), \(x_n\) be real numbers satisfying \[ x_1 + x_2 + \dots + x_n = 0, \quad \text{and} \quad x_1^2 + x_2^2 + \dots + x_n^2 = 1. \] For each subset \(A \subseteq \{1, 2, \dots, n\}\), define \[ S_A = \sum_{i \in A} x_i. \] (If \(A\) is the empty set, then \(S_A = 0\).) Prove that for any positive number \(\lambda\), the number of sets \(A\) satisfying \(S_A \ge \lambda\) is at most \(2^{n - 3}/\lambda^2\). For what choices of \(x_1\), \(x_2\), \(\dots\), \(x_n\), \(\lambda\) does equality hold?
[ "For convenience, let \\(N=\\{1,2,\\dots,n\\}\\).\n\nNote that \\(2\\sum_{1\\leq i<j\\leq n} x_ix_j=\\left(\\sum_{i=1}^{n}x_i\\right)^2-\\left(\\sum_{j=1}^{n} x_i^2\\right)=-1\\), so the sum of the \\(x_i\\) taken two at a time is \\(-1/2\\). Now consider the following sum:\n\n\\[\n\\sum_{A\\subseteq N}S_A^2=2^{n-1}\\left(\\sum_{j=1}^{n} x_i^2\\right)+2^{n-1}\\left(\\sum_{1\\leq i<j\\leq n} x_ix_j\\right)=2^{n-2}.\n\\]\n\nSince \\(S_A^2\\geq 0\\), it follows that at most \\(2^{n-2}/\\lambda^2\\) sets \\(A\\subseteq N\\) have \\(|S_A|\\geq \\lambda\\).\n\nNow note that \\(S_A+S_{N/A}=0\\). It follows that at most half of the \\(S_A\\) such that \\(|S_A|\\geq\\lambda\\) are positive. This shows that at most \\(2^{n-3}/\\lambda^2\\) sets \\(A\\subseteq N\\) satisfy \\(S_A\\geq \\lambda\\).\n\nNote that if equality holds, every subset \\(A\\) of \\(N\\) has \\(S_A\\in\\{-\\lambda,0,\\lambda\\}\\). It immediately follows that \\((x_1,x_2,\\ldots , x_n)\\) is a permutation of \\((\\lambda,-\\lambda,0,0,\\ldots , 0)\\). Since we know that \\(\\sum_{i=1}^{n} x_i^2=1\\), we have that \\(\\lambda=1/\\sqrt{2}\\).", "Let \\(x_i=x_1,x_2,...,x_n\\) It is evident that \\(x_i = \\frac{(-1)^i}{\\sqrt{n}}\\) for evens because of the second equation and \\(x_i=\\frac{(-1)^i}{\\sqrt{n-1}}\\) for odds(one term will be 0 to maintain the first condition). We may then try and get an expression for the maximum number of sets that satisfy this which occur when \\(\\lambda = \\frac{1}{\\sqrt{n}}\\): since it will be\n\n\\[\nx_1 + x_2 + ... + x_n\n\\]\n\nfor any choice of A we pick, it will have to be greater than \\(\\frac{1}{\\sqrt{n}}\\) which means we can either pick 0 negative \\(x_m\\) or \\(j-1\\) negatives for j positive terms. Since we also have that there are \\(\\frac{n}{2}\\) positive and negative terms for evens. Which then gives us:\n\n\\[\n\\sum_{k=1}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{k}\\sum_{i=0}^{k-1} \\binom{\\frac{n}{2}}{i}\n\\]\n\nand\n\n\\[\n\\sum_{k=1}^{\\frac{n}{2}}\\binom{\\frac{n}{2}}{k}\\left(2^{\\frac{n}{2}} - \\sum_{i=k}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{i}\\right)\n\\]\n\nFor odd values, let it be the same as the last even valued sequence where n is even(i.e. the same as the sequence before it but with an extra 0 in one of the spots). Then, the following is apparent:\n\n\\[\n\\sum_{k=1}^{\\lfloor\\frac{n}{2}\\rfloor}\\binom{\\lfloor \\frac{n}{2}\\rfloor}{k}\\left(2^{\\lfloor\\frac{n}{2}\\rfloor} - \\sum_{i=k}^{\\lfloor\\frac{n}{2}\\rfloor}\\binom{\\lfloor \\frac{n}{2}\\rfloor}{i}\\right) \\le n2^{n-3}\n\\]\n\nThus, we may say that this holds to be true for all \\(n \\ge 2\\) since \\(n2^{n-3}\\) grows faster than the sum. Note that equality holds when \\(S_A \\in \\{\\lambda,0,-\\lambda\\}\\) for all i which occurs when \\(x_i= \\frac{1}{\\sqrt{2}},-\\frac{1}{\\sqrt{2}},0,....\\) and \\(\\lambda = \\frac{1}{\\sqrt{2}}\\) since \\(S_A=\\frac{1}{\\sqrt{2}}\\) is the only choice for \\(S_A \\ge \\lambda\\) \\(\\blacksquare\\)\n\n## Note:\n\nThe proof to the last inequality is as follows: First we may rewrite this as being:\n\n\\[\n\\sum_{k=1}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{k} \\left( 2^{\\frac{n}{2}} - \\sum_{i=k}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{i}\\right) \\le n2^{n-3}\n\\]\n\nThus,\n\n\\[\n2^n - 2^{\\frac{n}{2}} - \\sum_{k=1}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{k} \\sum_{i=k}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{i} \\le n2^{n-3}\n\\]\n\n(the second equation is because the sum of the binomial coefficient is \\(2^{\\frac{n}{2}} - 1\\))\n\n\\[\n2^n - 2^{\\frac{n}{2}} \\le n2^{n-3} + \\sum_{k=1}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{k} \\sum_{i=k}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{i}\n\\]\n\nSince \\(2^n - 2^{\\frac{n}{2}} \\le n2^{n-3}\\) for all \\(n \\ge 8\\) and \\(\\sum_{k=1}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{k} \\sum_{i=k}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{i} \\ge 0\\) for all \\(k\\) and \\(i\\), it is apparent that:\n\n\\[\n\\sum_{k=1}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{k} \\sum_{i=0}^{k-1} \\binom{\\frac{n}{2}}{i} \\le n2^{n-3}\n\\]\n\nmust be true for all \\(n \\ge 8\\)(because if we rewrite this we get \\(2^{\\frac{n}{2}}\\left(8-n\\right) \\le 8\\).) For all \\(2 \\le n < 8\\) however, we may use some logic to first layout a plan. Since for \\(n=6,n=4,\\) and \\(n=2\\), \\(2^{n} - 2^{\\frac{n}{2}} = n2^{n-3} + 1\\), we may say that whole sum will be less than \\(n2^{n-3}\\) because \\(\\sum_{k=1}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{k} \\sum_{i=k}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{i} \\ge 1\\) for all \\(n \\ge 2\\) Plugging this inequality back in gives us:\n\n\\[\n2^{n} - 2^{\\frac{n}{2}} - \\sum_{k=1}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{k} \\sum_{i=k}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{i} \\le 2^{n} - 2^{\\frac{n}{2}} - 1 = n2^{n-3}\n\\]\n\nbecause of the fact that \\(- \\sum_{k=1}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{k} \\sum_{i=k}^{\\frac{n}{2}} \\binom{\\frac{n}{2}}{i} \\le -1\\)" ]
USAMO-2013-1
https://artofproblemsolving.com/wiki/index.php/2013_USAMO_Problems/Problem_1
In triangle \(ABC\), points \(P,Q,R\) lie on sides \(BC,CA,AB\) respectively. Let \(\omega_A\), \(\omega_B\), \(\omega_C\) denote the circumcircles of triangles \(AQR\), \(BRP\), \(CPQ\), respectively. Given the fact that segment \(AP\) intersects \(\omega_A\), \(\omega_B\), \(\omega_C\) again at \(X,Y,Z\) respectively, prove that \(YX/XZ=BP/PC\).
[ "\\[\n[asy] /* DRAGON 0.0.9.6 Homemade Script by v_Enhance. */ import olympiad; import cse5; size(11cm); real lsf=0.8000; real lisf=2011.0; defaultpen(fontsize(10pt)); /* Initialize Objects */ pair A = (-1.0, 3.0); pair B = (-3.0, -3.0); pair C = (4.0, -3.0); pair P = (-0.6698198198198195, -3.0); pair Q = (1.1406465288818244, 0.43122416534181074); pair R = (-1.6269590345062048, 1.119122896481385); path w_A = circumcircle(A,Q,R); path w_B = circumcircle(B,P,R); path w_C = circumcircle(P,Q,C); pair O_A = midpoint(relpoint(w_A, 0)--relpoint(w_A, 0.5)); pair O_B = midpoint(relpoint(w_B, 0)--relpoint(w_B, 0.5)); pair O_C = midpoint(relpoint(w_C, 0)--relpoint(w_C, 0.5)); pair X = (2)*(foot(O_A,A,P))-A; pair Y = (2)*(foot(O_B,A,P))-P; pair Z = (2)*(foot(O_C,A,P))-P; pair M = 2*foot(P,relpoint(O_B--O_C,0.5-10/lisf),relpoint(O_B--O_C,0.5+10/lisf))-P; pair D = (2)*(foot(O_B,X,M))-M; pair E = (2)*(foot(O_C,X,M))-M; /* Draw objects */ draw(A--B, rgb(0.6,0.6,0.0)); draw(B--C, rgb(0.6,0.6,0.0)); draw(C--A, rgb(0.6,0.6,0.0)); draw(w_A, rgb(0.4,0.4,0.0)); draw(w_B, rgb(0.4,0.4,0.0)); draw(w_C, rgb(0.4,0.4,0.0)); draw(A--P, rgb(0.0,0.2,0.4)); draw(D--E, rgb(0.0,0.2,0.4)); draw(P--D, rgb(0.0,0.2,0.4)); draw(P--E, rgb(0.0,0.2,0.4)); draw(P--M, rgb(0.4,0.2,0.0)); draw(R--M, rgb(0.4,0.2,0.0)); draw(Q--M, rgb(0.4,0.2,0.0)); draw(B--M, rgb(0.0,0.2,0.4)); draw(C--M, rgb(0.0,0.2,0.4)); draw((abs(dot(unit(M-P),unit(B-P))) < 1/2011) ? rightanglemark(M,P,B) : anglemark(M,P,B), rgb(0.0,0.8,0.8)); draw((abs(dot(unit(M-R),unit(A-R))) < 1/2011) ? rightanglemark(M,R,A) : anglemark(M,R,A), rgb(0.0,0.8,0.8)); draw((abs(dot(unit(M-X),unit(A-X))) < 1/2011) ? rightanglemark(M,X,A) : anglemark(M,X,A), rgb(0.0,0.8,0.8)); draw((abs(dot(unit(D-X),unit(P-X))) < 1/2011) ? rightanglemark(D,X,P) : anglemark(D,X,P), rgb(0.0,0.8,0.8)); /* Place dots on each point */ dot(A); dot(B); dot(C); dot(P); dot(Q); dot(R); dot(X); dot(Y); dot(Z); dot(M); dot(D); dot(E); /* Label points */ label(\"$A$\", A, lsf * dir(110)); label(\"$B$\", B, lsf * unit(B-M)); label(\"$C$\", C, lsf * unit(C-M)); label(\"$P$\", P, lsf * unit(P-M) * 1.8); label(\"$Q$\", Q, lsf * dir(90) * 1.6); label(\"$R$\", R, lsf * unit(R-M) * 2); label(\"$X$\", X, lsf * dir(-60) * 2); label(\"$Y$\", Y, lsf * dir(45)); label(\"$Z$\", Z, lsf * dir(5)); label(\"$M$\", M, lsf * dir(M-P)*2); label(\"$D$\", D, lsf * dir(150)); label(\"$E$\", E, lsf * dir(5));[/asy]\n\\]\n\nIn this solution, all lengths and angles are directed.\n\nFirstly, it is easy to see by that \\(\\omega_A, \\omega_B, \\omega_C\\) concur at a point \\(M\\) (the Miquel point). Let \\(XM\\) meet \\(\\omega_B, \\omega_C\\) again at \\(D\\) and \\(E\\), respectively. Then by Power of a Point, we have\n\n\\[\nXM \\cdot XE = XZ \\cdot XP \\quad\\text{and}\\quad XM \\cdot XD = XY \\cdot XP\n\\]\n\nThusly\n\n\\[\n\\frac{XY}{XZ} = \\frac{XD}{XE}\n\\]\n\nBut we claim that \\(\\triangle XDP \\sim \\triangle PBM\\). Indeed,\n\n\\[\n\\measuredangle XDP = \\measuredangle MDP = \\measuredangle MBP = - \\measuredangle PBM\n\\]\n\nand\n\n\\[\n\\measuredangle DXP = \\measuredangle MXY = \\measuredangle MXA = \\measuredangle MRA = 180^\\circ - \\measuredangle MRB = \\measuredangle MPB = -\\measuredangle BPM\n\\]\n\nTherefore, \\(\\frac{XD}{XP} = \\frac{PB}{PM}\\). Analogously we find that \\(\\frac{XE}{XP} = \\frac{PC}{PM}\\) and we are done.\n\ncourtesy v_enhance, minor clarification by integralarefun", "Diagram Refer to the Diagram link.\n\nBy Miquel's Theorem, there exists a point at which \\(\\omega_A, \\omega_B, \\omega_C\\) intersect. We denote this point by \\(M.\\) Now, we angle chase:\n\n\\[\n\\angle YMX = 180^{\\circ} - \\angle YXM - \\angle XYM\n\\]\n\n\\[\n= 180^{\\circ} - \\angle AXM - \\angle PYM\n\\]\n\n\\[\n= \\left(180^{\\circ} - \\angle ARM\\right) - \\angle PRM\n\\]\n\n\\[\n= \\angle BRM - \\angle PRM\n\\]\n\n\\[\n= \\angle BRP = \\angle BMP.\n\\]\n\nIn addition, we have\n\n\\[\n\\angle ZMX = 180^{\\circ} - \\angle MZY - \\angle ZYM - \\angle YMX\n\\]\n\n\\[\n= 180^{\\circ} - \\angle MZP - \\angle PYM - \\angle BMP\n\\]\n\n\\[\n= 180^{\\circ} - \\angle MCP - \\angle PBM - \\angle BMP\n\\]\n\n\\[\n= \\left(180^{\\circ} - \\angle PBM - \\angle BMP\\right) - \\angle MCP\n\\]\n\n\\[\n= \\angle BPM - \\angle MCP\n\\]\n\n\\[\n= 180^{\\circ} - \\angle MPC - \\angle MCP\n\\]\n\n\\[\n= \\angle CMP.\n\\]\n\nNow, by the Ratio Lemma, we have\n\n\\[\n\\frac{XY}{XZ} = \\frac{MY}{MZ} \\cdot \\frac{\\sin \\angle YMX}{\\sin \\angle ZMX}\n\\]\n\n\\[\n= \\frac{\\sin \\angle YZM}{\\sin \\angle ZYM} \\cdot \\frac{\\sin \\angle BMP}{\\sin \\angle CMP}\n\\]\n\n(by the Law of Sines in \\(\\triangle MZY\\))\n\n\\[\n= \\frac{\\sin \\angle PZM}{\\sin \\angle PYM} \\cdot \\frac{\\sin \\angle BMP}{\\sin \\angle CMP}\n\\]\n\n\\[\n= \\frac{\\sin \\angle PCM}{\\sin \\angle PBM} \\cdot \\frac{\\sin \\angle BMP}{\\sin \\angle CMP}\n\\]\n\n\\[\n= \\frac{MB}{MC} \\cdot \\frac{\\sin \\angle BMP}{\\sin \\angle CMP}\n\\]\n\n(by the Law of Sines in \\(\\triangle MBC\\))\n\n\\[\n= \\frac{PB}{PC}\n\\]\n\nby the Ratio Lemma. The proof is complete.", "Use directed angles modulo \\(\\pi\\).\n\nLemma. \\(\\angle{XRY} \\equiv \\angle{XQZ}.\\)\n\nProof.\n\n\\[\n\\angle{XRY} \\equiv \\angle{XRA} - \\angle{YRA} \\equiv \\angle{XQA} + \\angle{YRB} \\equiv \\angle{XQA} + \\angle{CPY} = \\angle{XQA} + \\angle{AQZ} = \\angle{XQZ}.\n\\]\n\nNow, it follows that (now not using directed angles)\n\n\\[\n\\frac{XY}{YZ} = \\frac{\\frac{XY}{\\sin \\angle{XRY}}}{\\frac{YZ}{\\sin \\angle{XQZ}}} = \\frac{\\frac{RY}{\\sin \\angle{RXY}}}{\\frac{QZ}{\\sin \\angle{QXZ}}} = \\frac{BP}{PC}\n\\]\n\nusing the facts that \\(ARY\\) and \\(APB\\), \\(AQZ\\) and \\(APC\\) are similar triangles, and that \\(\\frac{RA}{\\sin \\angle{RXA}} = \\frac{QA}{\\sin \\angle{QXA}}\\) equals twice the circumradius of the circumcircle of \\(AQR\\).", "We will use some construction arguments to solve the problem. Let \\(\\angle BAC=\\alpha,\\) \\(\\angle ABC=\\beta,\\) \\(\\angle ACB=\\gamma,\\) and let \\(\\angle APB=\\theta.\\) We construct lines through the points \\(Q,\\) and \\(R\\) that intersect with \\(\\triangle ABC\\) at the points \\(Q\\) and \\(R,\\) respectively, and that intersect each other at \\(T.\\) We will construct these lines such that \\(\\angle CQV=\\angle ARV=\\theta.\\)\n\nNow we let the intersections of \\(AP\\) with \\(RV\\) and \\(QU\\) be \\(Y'\\) and \\(Z',\\) respectively. This construction is as follows.\n\n\\[\n[asy] import graph; size(12cm); real labelscalefactor = 1.9; pen dps = linewidth(0.7) + fontsize(10); defaultpen(dps); pen rvwvcq = rgb(0.08235294117647059,0.396078431372549,0.7529411764705882); pen wrwrwr = rgb(0.3803921568627451,0.3803921568627451,0.3803921568627451); draw((-3.6988888888888977,6.426666666666669)--(-7.61,-5)--(7.09,-5)--cycle); draw((-3.6988888888888977,6.426666666666669)--(-7.61,-5)); draw((-7.61,-5)--(7.09,-5)); draw((7.09,-5)--(-3.6988888888888977,6.426666666666669)); draw((-3.6988888888888977,6.426666666666669)--(-2.958888888888898,-5)); draw((-5.053354907372894,2.4694710603912564)--(1.7922953932137468,0.6108747864253139)); draw((0.5968131669050584,1.8770271258031248)--(-5.8024625203461,-5)); dot((-3.6988888888888977,6.426666666666669)); label(\"$A$\", (-3.6988888888888977,6.426666666666669), N * labelscalefactor); dot((-7.61,-5)); label(\"$B$\", (-7.61,-5), SW * labelscalefactor); dot((7.09,-5)); label(\"$C$\", (7.09,-5), SE * labelscalefactor); dot((-2.958888888888898,-5)); label(\"$P$\", (-2.958888888888898,-5), S * labelscalefactor); dot((0.5968131669050584,1.8770271258031248)); label(\"$Q$\", (0.5968131669050584,1.8770271258031248), NE * labelscalefactor); dot((-5.053354907372894,2.4694710603912564)); label(\"$R$\", (-5.053354907372894,2.4694710603912564), dir(165) * labelscalefactor); dot((-3.143912404905382,-2.142970212141873)); label(\"$Z'$\", (-3.143912404905382,-2.142970212141873), dir(170) * labelscalefactor); dot((-3.413789986031826,2.0243286531799747)); label(\"$Y'$\", (-3.413789986031826,2.0243286531799747), NE * labelscalefactor); dot((-3.3284001481939356,0.7057864725120093)); label(\"$X$\", (-3.3284001481939356,0.7057864725120093), dir(220) * labelscalefactor); dot((1.7922953932137468,0.6108747864253139)); label(\"$V$\", (1.7922953932137468,0.6108747864253139), NE * labelscalefactor); dot((-5.8024625203461,-5)); label(\"$U$\", (-5.8024625203461,-5), S * labelscalefactor); dot((-0.10264330299819162,1.125351256231488)); label(\"$T$\", (-0.10264330299819162,1.125351256231488), dir(275) * labelscalefactor); [/asy]\n\\]\n\nWe know that \\(\\angle BRY'=180^\\circ-\\angle ARY'=180^\\circ-\\theta.\\) Hence, we have,\n\n\\[\n\\begin{align*} \\angle BRY'+\\angle BPY' &=180^\\circ-\\theta+\\theta\\\\ &=180^\\circ. \\end{align*}\n\\]\n\nSince the opposite angles of quadrilateral \\(RY'PB\\) add up to \\(180^\\circ,\\) it must be cyclic. Similarly, we can also show that quadrilaterals \\(CQZ'P,\\) and \\(AQTR\\) are also cyclic.\n\nSince points \\(Y'\\) and \\(Z'\\) lie on \\(AP,\\) we know that,\n\n\\[\nY'=\\omega_B\\cap AP\n\\]\n\nand that\n\n\\[\nZ'=\\omega_C\\cap AP.\n\\]\n\nHence, the points \\(Y'\\) and \\(Z'\\) coincide with the given points \\(Y\\) and \\(Z,\\) respectively.\n\nSince quadrilateral \\(AQTR\\) is also cyclic, we have,\n\n\\[\n\\begin{align*} \\angle Y'TZ' &=180^\\circ-\\angle RTQ\\\\ &=180^\\circ-(180^\\circ-\\angle RAQ)\\\\ &=\\angle RAQ\\\\ &=\\alpha. \\end{align*}\n\\]\n\nSimilarly, since quadrilaterals \\(CQZ'P,\\) and \\(AQTR\\) are also cyclic, we have,\n\n\\[\n\\begin{align*} \\angle TY'Z' &=180^\\circ-\\angle RY'P\\\\ &=180^\\circ-(180^\\circ-\\angle RBP)\\\\ &=\\angle RBP\\\\ &=\\beta, \\end{align*}\n\\]\n\nand,\n\n\\[\n\\begin{align*} \\angle Y'Z'T &=180^\\circ-\\angle PZ'Q\\\\ &=180^\\circ-(180^\\circ-\\angle PCQ)\\\\ &=\\angle PCQ\\\\ &=\\gamma. \\end{align*}\n\\]\n\nSince these three angles are of \\(\\triangle TY'Z',\\) and they are equal to corresponding angles of \\(\\triangle ABC,\\) by AA similarity, we know that \\(\\triangle TY'Z'\\sim \\triangle ABC.\\)\n\nWe now consider the point \\(X=\\omega_c\\cap AC.\\) We know that the points \\(A,\\) \\(Q,\\) \\(T,\\) and \\(R\\) are concyclic. Hence, the points \\(A,\\) \\(T,\\) \\(X,\\) and \\(R\\) must also be concyclic.\n\nHence, quadrilateral \\(AQTX\\) is cyclic.\n\n\\[\n[asy] import graph; size(12cm); real labelscalefactor = 1.9; pen dps = linewidth(0.7) + fontsize(10); defaultpen(dps); pen rvwvcq = rgb(0.08235294117647059,0.396078431372549,0.7529411764705882); pen wrwrwr = rgb(0.3803921568627451,0.3803921568627451,0.3803921568627451); draw((-3.6988888888888977,6.426666666666669)--(-7.61,-5)--(7.09,-5)--cycle); draw((-3.6988888888888977,6.426666666666669)--(-7.61,-5)); draw((-7.61,-5)--(7.09,-5)); draw((7.09,-5)--(-3.6988888888888977,6.426666666666669)); draw((-3.6988888888888977,6.426666666666669)--(-2.958888888888898,-5)); draw((-5.053354907372894,2.4694710603912564)--(1.7922953932137468,0.6108747864253139)); draw((0.5968131669050584,1.8770271258031248)--(-5.8024625203461,-5)); draw((-3.3284001481939356,0.7057864725120093)--(-0.10264330299819162,1.125351256231488)); draw((-3.3284001481939356,0.7057864725120093)--(-5.053354907372894,2.4694710603912564)); draw((-3.6988888888888977,6.426666666666669)--(-0.10264330299819162,1.125351256231488)); dot((-3.6988888888888977,6.426666666666669)); label(\"$A$\", (-3.6988888888888977,6.426666666666669), N * labelscalefactor); dot((-7.61,-5)); label(\"$B$\", (-7.61,-5), SW * labelscalefactor); dot((7.09,-5)); label(\"$C$\", (7.09,-5), SE * labelscalefactor); dot((-2.958888888888898,-5)); label(\"$P$\", (-2.958888888888898,-5), S * labelscalefactor); dot((0.5968131669050584,1.8770271258031248)); label(\"$Q$\", (0.5968131669050584,1.8770271258031248), NE * labelscalefactor); dot((-5.053354907372894,2.4694710603912564)); label(\"$R$\", (-5.053354907372894,2.4694710603912564), dir(165) * labelscalefactor); dot((-3.143912404905382,-2.142970212141873)); label(\"$Z'$\", (-3.143912404905382,-2.142970212141873), dir(170) * labelscalefactor); dot((-3.413789986031826,2.0243286531799747)); label(\"$Y'$\", (-3.413789986031826,2.0243286531799747), NE * labelscalefactor); dot((-3.3284001481939356,0.7057864725120093)); label(\"$X$\", (-3.3284001481939356,0.7057864725120093), dir(220) * labelscalefactor); dot((1.7922953932137468,0.6108747864253139)); label(\"$V$\", (1.7922953932137468,0.6108747864253139), NE * labelscalefactor); dot((-5.8024625203461,-5)); label(\"$U$\", (-5.8024625203461,-5), S * labelscalefactor); dot((-0.10264330299819162,1.125351256231488)); label(\"$T$\", (-0.10264330299819162,1.125351256231488), dir(275) * labelscalefactor); [/asy]\n\\]\n\nSince the angles \\(\\angle ART\\) and \\(\\angle AXT\\) are inscribed in the same arc \\(\\overarc{AT},\\) we have,\n\n\\[\n\\begin{align*} \\angle AXT &=\\angle ART\\\\ &=\\theta. \\end{align*}\n\\]\n\nConsider by this result, we can deduce that the homothety that maps \\(ABC\\) to \\(TY'Z'\\) will map \\(P\\) to \\(X.\\) Hence, we have that,\n\n\\[\nY'X/XZ'=BP/PC.\n\\]\n\nSince \\(Y'=Y\\) and \\(Z'=Z\\) hence,\n\n\\[\nYX/XZ=BP/PC,\n\\]\n\nas required.", "\\[\n[asy] /* DRAGON 0.0.9.6 */ import olympiad; import cse5; size(11cm); real lsf=0.8000; real lisf=2011.0; defaultpen(fontsize(10pt)); /* Initialize Objects */ pair A = (-1.0, 3.0); pair B = (-3.0, -3.0); pair C = (4.0, -3.0); pair P = (-0.6698198198198195, -3.0); pair Q = (1.1406465288818244, 0.43122416534181074); pair R = (-1.6269590345062048, 1.119122896481385); path w_A = circumcircle(A,Q,R); path w_B = circumcircle(B,P,R); path w_C = circumcircle(P,Q,C); pair O_A = midpoint(relpoint(w_A, 0)--relpoint(w_A, 0.5)); pair O_B = midpoint(relpoint(w_B, 0)--relpoint(w_B, 0.5)); pair O_C = midpoint(relpoint(w_C, 0)--relpoint(w_C, 0.5)); pair X = (2)*(foot(O_A,A,P))-A; pair Y = (2)*(foot(O_B,A,P))-P; pair Z = (2)*(foot(O_C,A,P))-P; pair M = 2*foot(P,relpoint(O_B--O_C,0.5-10/lisf),relpoint(O_B--O_C,0.5+10/lisf))-P; /* Draw objects */ draw(A--B, rgb(0.6,0.6,0.0)); draw(B--C, rgb(0.6,0.6,0.0)); draw(C--A, rgb(0.6,0.6,0.0)); draw(w_A, rgb(0.4,0.4,0.0)); draw(w_B, rgb(0.4,0.4,0.0)); draw(w_C, rgb(0.4,0.4,0.0)); draw(A--P, rgb(0.0,0.2,0.4)); draw(P--M, rgb(0.0,0.2,0.4)); draw(R--M, rgb(0.4,0.2,0.0)); draw(Q--M, rgb(0.4,0.2,0.0)); draw(B--M, rgb(0.0,0.2,0.4)); draw(C--M, rgb(0.0,0.2,0.4)); draw(Z--M, rgb(0.0,0.2,0.4)); draw(X--M, rgb(0.0,0.2,0.4)); draw(Y--M, rgb(0.0,0.2,0.4)); draw(C--Z, rgb(0.0,0.2,0.4)); draw(B--Y, rgb(0.0,0.2,0.4)); /* Place dots on each point */ dot(A); dot(B); dot(C); dot(P); dot(Q); dot(R); dot(X); dot(Y); dot(Z); dot(M); /* Label points */ label(\"$A$\", A, lsf * dir(110)); label(\"$B$\", B, lsf * unit(B-M)); label(\"$C$\", C, lsf * unit(C-M)); label(\"$P$\", P, lsf * unit(P-M) * 1.8); label(\"$Q$\", Q, lsf * dir(90) * 1.6); label(\"$R$\", R, lsf * unit(R-M) * 2); label(\"$X$\", X, lsf * dir(-60) * 2); label(\"$Y$\", Y, lsf * dir(45)); label(\"$Z$\", Z, lsf * dir(5)); label(\"$M$\", M, lsf * dir(M-P)*2); [/asy]\n\\]\n\nWe begin again by noting that the three circumcircles intersect at point \\(M\\) by Miquel's theorem. In addition, we state that the angle \\(\\angle MQC = \\alpha\\), hence \\(\\angle MPC = \\angle MZC = 180 - \\alpha\\), as well as \\(\\angle AQM\\), from which follows that \\(\\angle ARM = \\alpha\\), so \\(\\angle BRM = 180 - \\alpha\\), and \\(\\angle BPM = \\alpha\\). We shall prove that the triangles \\(\\triangle MZC\\) and \\(\\triangle XMP\\) and \\(\\triangle YMB\\) are similar, which will imply a rotational homotethy with angle \\(\\angle APB\\) about the point \\(M\\), that takes \\(Y,X,Z\\) to \\(B,P,C\\), thus proving the problem. (In essence, just imagine we rotate \\(YMZX\\) around M and lengthen things out and get \\(MCPB\\) - the ratios will remain identical.)\n\nWe do this by angle chasing. Denote angle \\(\\angle ZMP = \\beta\\). From the angles labeled before, we now know \\(\\angle ZPB = \\alpha-\\beta = \\angle ZMC\\). In addition, \\(\\angle ZPM = \\angle ZCM\\). So the angles in triangle \\(\\triangle ZMC\\) are \\(\\alpha-\\beta\\), \\(\\beta\\) and thus \\(180-\\alpha\\), with \\(\\alpha-\\beta\\) at the point \\(M\\). In addition, \\(\\angle YRM = \\angle YPM = \\beta\\), so \\(\\angle ARY = \\alpha-\\beta\\), so \\(\\angle YMB = \\alpha - \\beta\\). Since \\(\\angle YBM = \\angle YPM = \\beta\\), the triangle \\(\\triangle YMB\\) has angles \\(\\alpha-\\beta, \\beta, 180-\\alpha\\) also, with \\(\\alpha-\\beta\\) at \\(M\\). Finally, In triangle \\(\\triangle XPM\\), we already know the angle \\(\\angle XPM\\) to be \\(\\beta\\); we also can find that \\(\\angle AQM = 180-\\alpha\\), so \\(\\angle AXM = \\alpha\\), so \\(\\angle MXP = 180-\\alpha\\), so \\(\\angle XMP = \\alpha-\\beta\\). Thus, the three triangles are similar have a common point \\(M\\), which proves that there is a rotational homotethy around \\(M\\) that maps \\(Y, X, Z\\) to \\(B, P, C\\) as desired.\n\nSolution by SimilarTriangle." ]
USAMO-2013-2
https://artofproblemsolving.com/wiki/index.php/2013_USAMO_Problems/Problem_2
For a positive integer \(n\geq 3\) plot \(n\) equally spaced points around a circle. Label one of them \(A\), and place a marker at \(A\). One may move the marker forward in a clockwise direction to either the next point or the point after that. Hence there are a total of \(2n\) distinct moves available; two from each point. Let \(a_n\) count the number of ways to advance around the circle exactly twice, beginning and ending at \(A\), without repeating a move. Prove that \(a_{n-1}+a_n=2^n\) for all \(n\geq 4\).
[ "We label the points in clockwise order as \\(1,2,3,\\ldots,n\\), where point \\(A\\) is the same as point \\(1\\). We start and end at point \\(1\\), and we must cross over it, either by visiting it again, or else by making the move from point \\(n\\) to point \\(2\\). We interpret each of these cases in terms of tiling. In each move, we either move one or two points clockwise, so we can think of each move as a \\(1\\times 1\\) or \\(1\\times 2\\) tile. If the point \\(1\\) is visited in the middle, then the first cycle around the circle can be thought of as a tiling of a \\(1\\times n\\) board, and the second cycle around the circle can also be thought of as a \\(1\\times n\\) board. We place this second board directly below the first board. Therefore, in this first case, we wish to find the number of tilings of two \\(1\\times n\\) boards, and to guarantee that no move is repeated, we cannot have two tiles of the same type lying directly atop each other in the \\(2\\times n\\) board. Suppose there are \\(u_n\\) such tilings. It can easily be computed that \\(u_3=2\\) and \\(u_4=6\\) as shown below.\n\n\\[\n[asy] unitsize(10); draw((0,0)--(3,0)--(3,2)--(0,2)--cycle^^(0,1)--(3,1),linewidth(2)); draw((1,0)--(1,1)^^(2,1)--(2,2)); draw(shift((0,-2.5))*((0,0)--(3,0)--(3,2)--(0,2)--cycle^^(0,1)--(3,1)),linewidth(2)); draw(shift((0,-2.5))*((1,0)--(1,1)^^(2,1)--(2,2))); draw(shift((7,0))*((0,0)--(4,0)--(4,2)--(0,2)--cycle^^(0,1)--(4,1)),linewidth(2)); draw(shift((7,0))*((1,1)--(1,2)^^(2,0)--(2,2)^^(3,1)--(3,2))); draw(shift((7,-2.5))*((0,0)--(4,0)--(4,2)--(0,2)--cycle^^(0,1)--(4,1)),linewidth(2)); draw(shift((7,-2.5))*((1,0)--(1,1)^^(2,0)--(2,2)^^(3,1)--(3,2))); draw(shift((12,0))*((0,0)--(4,0)--(4,2)--(0,2)--cycle^^(0,1)--(4,1)),linewidth(2)); draw(shift((12,0))*((1,1)--(1,2)^^(2,0)--(2,1)^^(3,1)--(3,2))); draw(shift((12,-2.5))*((0,0)--(4,0)--(4,2)--(0,2)--cycle^^(0,1)--(4,1)),linewidth(2)); draw(shift((12,-2.5))*((1,1)--(1,2)^^(2,0)--(2,2)^^(3,0)--(3,1))); draw(shift((17,0))*((0,0)--(4,0)--(4,2)--(0,2)--cycle^^(0,1)--(4,1)),linewidth(2)); draw(shift((17,0))*((1,0)--(1,1)^^(2,1)--(2,2)^^(3,0)--(3,1))); draw(shift((17,-2.5))*((0,0)--(4,0)--(4,2)--(0,2)--cycle^^(0,1)--(4,1)),linewidth(2)); draw(shift((17,-2.5))*((1,0)--(1,1)^^(2,0)--(2,2)^^(3,0)--(3,1))); [/asy]\n\\]\n\nIn the second case, where we pass by point \\(1\\) by moving from point \\(n\\) to point \\(2\\), we can similarly think about it in terms of tiling two rows of \\(1\\times n\\) boards, but we remove the last square in the first row and the first square in the second row to make sure that we jump from point \\(n\\) to point \\(2\\). Suppose that we can tile such boards in \\(s_n\\) ways. It can be easily computed that \\(s_3=3\\) and \\(s_4=5\\) as shown below.\n\n\\[\n[asy] unitsize(10); draw((1,0)--(3,0)--(3,1)--(2,1)--(2,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(2,1),linewidth(2)); draw(shift((0,-2.5))*((1,0)--(3,0)--(3,1)--(2,1)--(2,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(2,1)),linewidth(2)); draw(shift((0,-2.5))*((2,0)--(2,1))); draw(shift((4,0))*((1,0)--(3,0)--(3,1)--(2,1)--(2,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(2,1)),linewidth(2)); draw(shift((4,0))*((1,1)--(1,2))); draw(shift((11,0))*((1,0)--(4,0)--(4,1)--(3,1)--(3,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(3,1)),linewidth(2)); draw(shift((11,0))*((1,1)--(1,2)^^(2,1)--(2,2)^^(3,0)--(3,1))); draw(shift((11,-2.5))*((1,0)--(4,0)--(4,1)--(3,1)--(3,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(3,1)),linewidth(2)); draw(shift((11,-2.5))*((2,0)--(2,2))); draw(shift((16,0))*((1,0)--(4,0)--(4,1)--(3,1)--(3,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(3,1)),linewidth(2)); draw(shift((16,0))*((2,1)--(2,2)^^(3,0)--(3,1))); draw(shift((16,-2.5))*((1,0)--(4,0)--(4,1)--(3,1)--(3,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(3,1)),linewidth(2)); draw(shift((16,-2.5))*((1,1)--(1,2)^^(2,0)--(2,1)^^(3,0)--(3,1))); draw(shift((21,0))*((1,0)--(4,0)--(4,1)--(3,1)--(3,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(3,1)),linewidth(2)); draw(shift((21,0))*((1,1)--(1,2)^^(2,0)--(2,1))); [/asy]\n\\]\n\nSince these are the only two possible cases, we see that\n\n\\[\na_n=u_n+s_n\\tag{1}.\n\\]\n\nFor the sake of convenience in determining recurrence relations, we define another type of board with two \\(1\\times n\\) boards where a specific corner is removed (without loss of generality, we place this in the lower left hand corner). Let \\(t_n\\) be the number of ways to tile such a board without placeing two of the same type of tile atop each other. Once again, we compute that \\(t_3=3\\) and \\(t_4=5\\) as shown below.\n\n\\[\n[asy] unitsize(10); draw((1,0)--(3,0)--(3,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(3,1),linewidth(2)); draw((1,1)--(1,2)^^(2,0)--(2,1)); draw(shift((0,-2.5))*((1,0)--(3,0)--(3,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(3,1)),linewidth(2)); draw(shift((0,-2.5))*((1,1)--(1,2)^^(2,1)--(2,2))); draw(shift((4,0))*((1,0)--(3,0)--(3,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(3,1)),linewidth(2)); draw(shift((4,0))*((2,1)--(2,2))); draw(shift((11,0))*((1,0)--(4,0)--(4,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(4,1)),linewidth(2)); draw(shift((11,0))*((1,1)--(1,2)^^(2,1)--(2,2)^^(3,0)--(3,1))); draw(shift((11,-2.5))*((1,0)--(4,0)--(4,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(4,1)),linewidth(2)); draw(shift((11,-2.5))*((1,1)--(1,2)^^(2,0)--(2,1)^^(3,1)--(3,2))); draw(shift((16,0))*((1,0)--(4,0)--(4,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(4,1)),linewidth(2)); draw(shift((16,0))*((2,0)--(2,2)^^(3,1)--(3,2))); draw(shift((16,-2.5))*((1,0)--(4,0)--(4,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(4,1)),linewidth(2)); draw(shift((16,-2.5))*((2,1)--(2,2)^^(3,0)--(3,1))); draw(shift((21,0))*((1,0)--(4,0)--(4,2)--(0,2)--(0,1)--(1,1)--cycle^^(1,1)--(4,1)),linewidth(2)); draw(shift((21,0))*((2,0)--(2,2)^^(3,0)--(3,1))); [/asy]\n\\]\n\nWe can determine reccurence relations for \\(s_n\\), \\(t_n\\), and \\(u_n\\) in terms of each other. For \\(s_n\\), note that a tiling can end in one of the three following ways such that the rest of the board can be tiled without restriction (the placed tiles are shaded).\n\n\\[\n[asy] path unitrect=(0,0)--(2,0)--(2,1)--(0,1)--cycle; unitsize(10); fill(shift((3,0))*unitsquare,mediumgray); draw(shift((0,0))*((0,0)--(4,0)--(4,1)--(0,1)^^(3,1)--(3,2)--(0,2)),linewidth(2)); draw(shift((0,0))*((3,0)--(3,1))); label(\"$\\dots$\",(-1,1)); fill(shift((10,1))*unitsquare,mediumgray); fill(shift((10,0))*unitrect,mediumgray); draw(shift((8,0))*((0,0)--(4,0)--(4,1)--(0,1)^^(3,1)--(3,2)--(0,2)),linewidth(2)); draw(shift((8,0))*((2,0)--(2,2))); label(\"$\\dots$\",(8-1,1)); fill(shift((18,0))*unitrect,mediumgray); fill(shift((17,1))*unitrect,mediumgray); draw(shift((16,0))*((0,0)--(4,0)--(4,1)--(0,1)^^(3,1)--(3,2)--(0,2)),linewidth(2)); draw(shift((16,0))*((2,0)--(2,1)^^(1,1)--(1,2))); label(\"$\\dots$\",(16-1,1)); [/asy]\n\\]\n\nIn the first, second, and third cases, we see that we can tile the rest of the board in \\(t_{n-1}\\), \\(t_{n-2}\\), and \\(s_{n-2}\\) ways, respectively. Hence for \\(n\\ge 5\\), we see that\n\n\\[\ns_n=t_{n-1}+t_{n-2}+s_{n-2}\\tag{2}.\n\\]\n\nFor \\(t_n\\), note that a tiling can end in one of the three following ways such that the rest of the board can be tiled without restriction.\n\n\\[\n[asy] path unitrect=(0,0)--(2,0)--(2,1)--(0,1)--cycle; unitsize(10); filldraw(shift((3,0))*unitsquare,mediumgray); filldraw(shift((2,1))*unitrect,mediumgray); draw(shift((0,0))*((0,0)--(4,0)--(4,1)--(0,1)^^(4,1)--(4,2)--(0,2)),linewidth(2)); label(\"$\\dots$\",(-1,1)); filldraw(shift((11,1))*unitsquare,mediumgray); filldraw(shift((10,0))*unitrect,mediumgray); filldraw(shift((9,1))*unitrect,mediumgray); draw(shift((8,0))*((0,0)--(4,0)--(4,1)--(0,1)^^(4,1)--(4,2)--(0,2)),linewidth(2)); label(\"$\\dots$\",(8-1,1)); filldraw(shift((19,1))*unitsquare,mediumgray); filldraw(shift((18,1))*unitsquare,mediumgray); filldraw(shift((18,0))*unitrect,mediumgray); draw(shift((16,0))*((0,0)--(4,0)--(4,1)--(0,1)^^(4,1)--(4,2)--(0,2)),linewidth(2)); label(\"$\\dots$\",(16-1,1)); [/asy]\n\\]\n\nIn the first, second, and third cases, we see that we can tile the rest of the board in \\(s_{n-1}\\), \\(s_{n-2}\\), and \\(t_{n-2}\\) ways, respectively. Hence for \\(n\\ge 5\\), we see that\n\n\\[\nt_n=s_{n-1}+s_{n-2}+t_{n-2}\\tag{3}.\n\\]\n\nFor \\(u_n\\), note that a tiling can end in one of the two following ways such that the rest of the board can be tiled without restriction.\n\n\\[\n[asy] path unitrect=(0,0)--(2,0)--(2,1)--(0,1)--cycle; unitsize(10); filldraw(shift((3,1))*unitsquare,mediumgray); filldraw(shift((2,0))*unitrect,mediumgray); draw(shift((0,0))*((0,0)--(4,0)--(4,1)--(0,1)^^(4,1)--(4,2)--(0,2)),linewidth(2)); label(\"$\\dots$\",(-1,1)); filldraw(shift((11,0))*unitsquare,mediumgray); filldraw(shift((10,1))*unitrect,mediumgray); draw(shift((8,0))*((0,0)--(4,0)--(4,1)--(0,1)^^(4,1)--(4,2)--(0,2)),linewidth(2)); label(\"$\\dots$\",(8-1,1)); [/asy]\n\\]\n\nIn either case, we are left with a board with a corner removed, hence we can tile the rest of the board in \\(t_{n-1}\\) ways in each case. Hence for \\(n\\ge 4\\), we see that\n\n\\[\nu_n=2t_{n-1}\\tag{4}.\n\\]\n\nSubtracting (3) from (2), we find that\n\n\\[\ns_n-t_n=s_{n-1}-t_{n-1}.\n\\]\n\nTherefore, if \\(s_{n-1}=t_{n-1}\\), then \\(s_n=t_n\\). Since \\(s_3=t_3\\) and \\(s_4=t_4\\), we see that \\(s_n=t_n\\) for all \\(n\\ge 3\\). Therefore, (2) can be rewritten as\n\n\\[\ns_n=s_{n-1}+2s_{n-2}\\tag{5},\n\\]\n\nand (4) can be rewritten as\n\n\\[\nu_n=2s_{n-1}\\tag{6}.\n\\]\n\nNow by (1), we know that\n\n\\[\na_{n}+a_{n-1}=u_n+s_n+u_{n-1}+s_{n-1}.\\tag{7}\n\\]\n\nIn particular, \\(a_4+a_3=6+5+2+3=2^4\\), so the statement is true for \\(n=4\\). Then by (7), and then substituting (5) and (6) (where these are valid for \\(n\\ge 5\\)), we find\n\n\\[\n\\begin{align*} a_n+a_{n-1}&=u_n+s_n+u_{n-1}+s_{n-1}\\\\ &=2s_{n-1}+(s_{n-1}+2s_{n-2})+u_{n-1}+s_{n-1}\\\\ &=4s_{n-1}+2s_{n-2}+u_{n-1}.\\tag{8} \\end{align*}\n\\]\n\nBut then by (5), and then substituting \\(4s_{n-3}=2u_{n-2}\\) and \\(2s_{n-2}=u_{n-1}\\) (by (6), and these are valid for \\(n\\ge 6\\)), we find\n\n\\[\n\\begin{align*} 2s_{n-1}&=2s_{n-2}+4s_{n-3}\\\\ &=u_{n-1}+2u_{n-2}. \\end{align*}\n\\]\n\nWe substitute this into (8), finding that for \\(n\\ge 6\\),\n\n\\[\n\\begin{align*} a_{n}+a_{n-1}&=2s_{n-1}+2s_{n-2}+u_{n-1}+(u_{n-1}+2u_{n-2})\\\\ &=2(s_{n-1}+s_{n-2}+u_{n-1}+u_{n-2})\\\\ &=2(a_{n-1}+a_{n-2}).\\tag{9} \\end{align*}\n\\]\n\nTherefore, if \\(a_{n-1}+a_{n-2}=2^{n-1}\\), then \\(a_{n}+a_{n-1}=2^n\\). We already know that \\(a_4+a_3=2^4\\). Also, we can compute that\n\n\\[\n\\begin{align*} s_5&=s_4+2s_3=5+2\\cdot 3=11\\\\ u_5&=2s_4=2\\cdot 5=10. \\end{align*}\n\\]\n\nHence \\(a_5=s_5+u_5=21\\). So as \\(a_4=s_4+u_4=11\\), we find that \\(a_5+a_4=2^5\\). Then we can use (9) for \\(n\\ge 6\\) to find by induction that \\(a_{n}+a_{n-1}=2^n\\) for all \\(n\\ge 4\\).", "We choose the tilings for the first and second pass around the circle simultaneously. As we tile both passes, our state is defined by whether or not we are in a continuation of a length 2 block on the first and second pass. We can't be in a continuation of a length 2 block on both passes, since that would mean we used the same move on the same block on both passes. Thus we have three states: (non-continuation, non-continuation), (non-continuation, continuation), and (continuation, non-continuation), where the elements of each tuple refer to the first and second pass respectively. Listing out the state transitions (while following the rule that we can't use the same move on both passes, and that a continuation must become a non-continuation), we get the transition matrix \\(T = \\begin{bmatrix} 0 & 1 & 1 \\\\ 1 & 0 & 1 \\\\ 1 & 1 & 0 \\end{bmatrix} = J-I\\), where \\(J\\) is the ones matrix, and \\(I\\) is the identity. We have that \\(a_n = \\left(T^n\\right)_{11} + \\left(T^n\\right)_{23}\\). The first term represents the case where we end the first pass at point \\(A\\), and the second term represents the case where we jump over point \\(A\\) when finishing the first pass. Thus it simply remains to compute \\(T^n\\). We have\n\n\\begin{align*} &\\mathrel{\\phantom{=}} T^n \\\\ &= (J-I)^n \\\\ &= \\sum_{k=0}^n J^k (-1)^{n-k} {n \\choose k} \\\\ &= J \\sum_{k=1}^n 3^{k-1} (-1)^{n-k} {n \\choose k} + (-1)^n I \\\\ &= \\frac{1}{3}J\\left(2^n-(-1)^n\\right) + (-1)^n I \\end{align*}\n\nThus \\(a_n = \\left(T^n\\right)_{11} + \\left(T^n\\right)_{23} = \\frac{2}{3}\\left(2^n-(-1)^n\\right) + (-1)^n\\). Thus, \\(a_{n-1}+a_n = 2^n\\), as desired.", "We split this into two cases. Case 1: The marker lands on \\(A\\) thrice. This implies that \\(A\\) is touched after one full cycle. Place any nodes not touched in the first cycle in the set \\(K\\). Lemma 1: Any unique configuration of \\(K\\) such that [Work In Progress] ~SigmaPiE\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2013-3
https://artofproblemsolving.com/wiki/index.php/2013_USAMO_Problems/Problem_3
Let \(n\) be a positive integer. There are \(\tfrac{n(n+1)}{2}\) marks, each with a black side and a white side, arranged into an equilateral triangle, with the biggest row containing \(n\) marks. Initially, each mark has the black side up. An operation is to choose a line parallel to the sides of the triangle, and flipping all the marks on that line. A configuration is called admissible if it can be obtained from the initial configuration by performing a finite number of operations. For each admissible configuration \(C\), let \(f(C)\) denote the smallest number of operations required to obtain \(C\) from the initial configuration. Find the maximum value of \(f(C)\), where \(C\) varies over all admissible configurations.
[ "Solution by inconsistent\n\nThis problem is headache-inducing... but also fun. Ouch.\n\nLet \\(n = 4k + b\\) where \\(b \\in {1, 2, 3, 4}\\). Then I claim the answer is \\(6k+1, 6k+2, 6k+3, 6k+6\\) in these four cases. First, note that the base cases \\(n = 1, 2, 3\\) are trivial by checking to be \\(1, 2, 3\\). Now note the case \\(n = 4\\) needs at least \\(6\\) to construct a black triangle with a single white mark at its center by checking. Call this size \\(4\\) triangle the evil triangle.\n\nTo show the lower bound, notice we can take the top \\(n\\) rows of a \\(n+4\\) equilateral triangle to be the limiting case for \\(n\\) and put two evil triangles on the bottom left and bottom right, respectively: these force \\(4\\) new rows to be occupied since each evil subcase requires at least two rows in every direction, which is a claim easy verified.\n\nTo show the upper bound, we reparameterize the triangles in terms of their rows. Notably, we can start counting rows from each corner, and give each mark a barycentric coordinate \\((a, b, c)\\) where \\(a, b, c\\) are integers denoting the \\(i\\)th row in a given direction the point lies in, and this satisfies \\(a+b+c = 2n+1\\) by Viviani's theorem.\n\nNow, consider the configuration \\(C\\) that minimizes \\(f(C)\\). We place the rows in boxes: odd rows of each direction, and even rows of each direction, to make six boxes total. Call the numbers of odd rows \\(a, b, c\\) and the even rows \\(d, e, f\\) such that \\((a, b, c)\\) and \\((d, e, f)\\) correspond to the same directions respectively in that order. Notice no row is operated on twice since row operations are commutative. Assume for sake of contradiction our claim is wrong, then we can construct a solution for the case one higher than our claim.\n\nWe now perform casework: for \\(b = 4\\), it follows one of the quantities \\(a+b+d+e, b+c+e+f, c+a+f+d\\) is greater than \\(n\\). In this case, invert the rows chosen in those four boxes (i.e. remove the ones chosen and choose the ones unchosen). This fixes the configuration and decreases \\(f(C)\\), giving contradiction.\n\nNow, consider the case \\(b = 1\\). Like the above paragraph, we know \\(a+d, b+e, c+f\\) are equal to \\(2k, 2k+1, 2k+1\\) in some order since if any pair sums to more than \\(4k+1\\) then we reach a contradiction. But then \\(2k+1 + 2k + 1 > 4k + 2\\), giving contradiction.\n\nNow, for the two remaining cases, we show a lemma: the xor of all the odd rows in two directions and all the even row in a third direction, or the sum of all the even rows in all directions are both \\(0\\). This result follows immediately from our earlier result via Vivani's theorem (it cannot be that all three claims are true, and it cannot be the case that two of the claims are false and one is true by parity).\n\nThus in the \\(b = 2\\) case it follows that \\(\\min(a+b+f, b+c+d, c+a+e, d+e+f) \\geq \\frac{12k+6}{4}\\) so one of the triples is at least \\(3k + 2\\). However inverting the three involved boxes, that total to \\(6k+3\\) rows and don't affect the configuration when all three are inverted, decreases the \\(f(C)\\), giving contradiction.\n\nFinally, in the \\(b = 3\\) it follows that \\((a+b+f)+(b+c+d) + (c+a+e) + (d+e+f) =12k+8\\). In particular, the odd/odd/even triples consist of \\(\\frac{n+1}{2} + \\frac{n+1}{2} + \\frac{n-1}{2} = 6k+5\\) rows each, while the even/even/even triple consists of \\(3 \\cdot \\frac{n-1}{2} = 6k+3\\) rows. For each of the triples to avoid decreases \\(f(C)\\) after being inverted, it needs to satisfy either \\(\\leq 3k + 2\\) or \\(\\leq 3k+1\\) depending on the two cases, so it follows that \\(LHS \\leq 12k + 7\\), giving contradiction.\n\nThus we have reached contradiction in all cases of counterexamples, proving our desired claim.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2013-4
https://artofproblemsolving.com/wiki/index.php/2013_USAMO_Problems/Problem_4
Find all real numbers \(x,y,z\geq 1\) satisfying \[ \min(\sqrt{x+xyz},\sqrt{y+xyz},\sqrt{z+xyz})=\sqrt{x-1}+\sqrt{y-1}+\sqrt{z-1}. \]
[ "The key Lemma is:\n\n\\[\n\\sqrt{a-1}+\\sqrt{b-1} \\le \\sqrt{ab}\n\\]\n\nfor all \\(a,b \\ge 1\\). Equality holds when \\((a-1)(b-1)=1\\).\n\nThis is proven easily.\n\n\\[\n\\sqrt{a-1}+\\sqrt{b-1} = \\sqrt{a-1}\\sqrt{1}+\\sqrt{1}\\sqrt{b-1} \\le \\sqrt{(a-1+1)(b-1+1)} = \\sqrt{ab}\n\\]\n\nby Cauchy.\n\nEquality then holds when \\(a-1 =\\frac{1}{b-1} \\implies (a-1)(b-1) = 1\\).\n\nNow assume that \\(x = \\min(x,y,z)\\). Now note that, by the Lemma,\n\n\\[\n\\sqrt{x-1}+\\sqrt{y-1}+\\sqrt{z-1} \\le \\sqrt{x-1} + \\sqrt{yz} \\le \\sqrt{x(yz+1)} = \\sqrt{xyz+x}\n\\]\n\n. So equality must hold in order for the condition in the problem statement to be met. So \\((y-1)(z-1) = 1\\) and \\((x-1)(yz) = 1\\). If we let \\(z = c\\), then we can easily compute that \\(y = \\frac{c}{c-1}, x = \\frac{c^2+c-1}{c^2}\\). Now it remains to check that \\(x \\le y, z\\).\n\nBut by easy computations, \\(x = \\frac{c^2+c-1}{c^2} \\le c = z \\Longleftrightarrow (c^2-1)(c-1) \\ge 0\\), which is obvious. Also \\(x = \\frac{c^2+c-1}{c^2} \\le \\frac{c}{c-1} = y \\Longleftrightarrow 2c \\ge 1\\), which is obvious, since \\(c \\ge 1\\).\n\nSo all solutions are of the form \\(\\boxed{\\left(\\frac{c^2+c-1}{c^2}, \\frac{c}{c-1}, c\\right)}\\), and all permutations for \\(c > 1\\).\n\nRemark: An alternative proof of the key Lemma is the following: By AM-GM,\n\n\\[\n(ab-a-b+1)+1 = (a-1)(b-1) + 1 \\ge 2\\sqrt{(a-1)(b-1)}\n\\]\n\n\\[\nab\\ge (a-1)+(b-1)+2\\sqrt{(a-1)(b-1)}\n\\]\n\n. Now taking the square root of both sides gives the desired. Equality holds when \\((a-1)(b-1) = 1\\).", "WLOG, assume that \\(x = \\min(x,y,z)\\). Let \\(a=\\sqrt{x-1},\\) \\(b=\\sqrt{y-1}\\) and \\(c=\\sqrt{z-1}\\). Then \\(x=a^2+1\\), \\(y=b^2+1\\) and \\(z=c^2+1\\). The equation becomes\n\n\\[\n(a^2+1)+(a^2+1)(b^2+1)(c^2+1)=(a+b+c)^2.\n\\]\n\nRearranging the terms, we have\n\n\\[\n(1+a^2)(bc-1)^2+[a(b+c)-1]^2=0.\n\\]\n\nTherefore \\(bc=1\\) and \\(a(b+c)=1.\\) Express \\(a\\) and \\(b\\) in terms of \\(c\\), we have \\(a=\\frac{c}{c^2+1}\\) and \\(b=\\frac{1}{c}.\\) Easy to check that \\(a\\) is the smallest among \\(a\\), \\(b\\) and \\(c.\\) Then \\(x=\\frac{c^4+3c^2+1}{(c^2+1)^2}\\), \\(y=\\frac{c^2+1}{c^2}\\) and \\(z=c^2+1.\\) Let \\(c^2=t\\), we have the solutions for \\((x,y,z)\\) as follows: \\((\\frac{t^2+3t+1}{(t+1)^2}, \\frac{t+1}{t}, t+1)\\) and permutations for all \\(t>0.\\)\n\n--J.Z.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2013-5
https://artofproblemsolving.com/wiki/index.php/2013_USAMO_Problems/Problem_5
Given positive integers \(m\) and \(n\), prove that there is a positive integer \(c\) such that the numbers \(cm\) and \(cn\) have the same number of occurrences of each non-zero digit when written in base ten.
[ "This solution is adopted from the official solution. Both the problem and the solution were suggested by Richard Stong.\n\nWithout Loss of Generality, suppose \\(m \\geq n \\geq 1\\). By prime factorization of \\(n\\), we can find a positive integer \\(c_1\\) such that \\(c_1n=10^s n_1\\) where \\(n_1\\) is relatively prime to \\(10\\). If a positive \\(k\\) is larger than \\(s\\), then \\((10^k c_1 m - c_1 n)= 10^s t\\), where \\(t=10^{k-s} c_1m-n_1>0\\) is always relatively prime to \\(10\\).\n\nChoose a \\(k\\) large enough so that \\(t\\) is larger than \\(c_1m\\). We can find an integer \\(b\\geq 1\\) such that \\(10^b-1\\) is divisible by \\(t\\), and also larger than \\(10c_1m\\). For example, let \\(b=\\varphi(t)\\) and use Euler's theorem. Now, let \\(c_2=(10^b-1)/t\\), and \\(c=c_1c_2\\). We claim that \\(c\\) is the desired number.\n\nIndeed, since both \\(c_1m\\) and \\(n_1\\) are less than \\(t\\), we see that the decimal expansion of both the fraction \\((c_1m)/t = (cm)/(c_2t) = (cm)/(10^b-1)\\) and \\(n_1/t=(c_2n_1)/(10^b-1)\\) are repeated in \\(b\\)-digit. And we also see that \\(10^k (c_1m)/t = (t+c_1n)/t= 1+10^s (n_1/t)\\), therefore the two repeated \\(b\\)-digit expansions are cyclic shift of one another.\n\nThis proves that \\(cm\\) and \\(c_2n_1\\) have the same number of occurrences of non-zero digits. Furthermore, \\(cn = c_2c_1n=10^s c_2n_1\\) also have the same number of occurrences of non-zero digits with \\(c_2n_1\\).", "This is a rephrasing of the above solution.\n\nIt is enough to solve the problem when \\(m,n\\) are replaced by \\(km,kn\\) for any positive integer \\(k\\). In particular, by taking \\(k=2^a5^b\\) for appropriate values of \\(a,b\\), we may assume \\(n=10^sn_1\\) where \\(n_1\\) is relatively prime to 10.\n\nFurthermore, adding or removing trailing zeros from \\(m\\) and \\(n\\) doesn't affect the claim, so we may further assume \\(\\gcd(n,10)=1\\) and that \\(m\\) has a xillion trailing zeros (enough to make \\(m\\) way bigger than \\(n\\), and also so that \\(m\\) has at least one trailing zero).\n\nFor clarity of exposition, we will also multiply \\(m,n\\) by a small number to make the units digit of \\(n\\) be \\(1\\) (though this is not necessary for the solution to work).\n\nThe point is that, for any positive integer \\(X\\), most nonzero digits appear the same number of times in \\(X\\) and \\(X+999\\dots999\\) if there are enough \\(9\\)s; In particular, if the units digits of \\(X\\) is \\(1\\), then all nonzero digits appear the same number of times as long as there are at least as many \\(9\\)s as digits in \\(X\\).\n\nSo we will pick \\(c\\) to satisfy:\n\n- \\(mc=nc+999\\dots999\\) where the number of \\(9\\)s is more than the number of digits of \\(nc\\)\n- The units digit of \\(nc\\) is \\(1\\).\n\nBecause we made the units of \\(n\\) to be \\(1\\), the second condition is equivalent to making the units digit of \\(c\\) to be \\(1\\).\n\nThe first condition is equivalent to \\((m-n)c=999\\dots999\\). Because \\(m\\) has at least one trailing 0, the units digit of \\(m-n\\) is 9, so \\(\\gcd(m-n,10)=1\\) and there is some \\(c\\) so that \\((m-n)c=999\\dots999\\), and the units digit of \\(c\\) must be \\(1\\) which agrees with the other condition.\n\nFinally, as \\(m\\gg n, (m-n)c\\approx mc\\gg nc\\) so it is possible to make the number of \\(9\\)s more than the number of digits in \\(nc\\).\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2013-6
https://artofproblemsolving.com/wiki/index.php/2013_USAMO_Problems/Problem_6
Let \(ABC\) be a triangle. Find all points \(P\) on segment \(BC\) satisfying the following property: If \(X\) and \(Y\) are the intersections of line \(PA\) with the common external tangent lines of the circumcircles of triangles \(PAB\) and \(PAC\), then \[ \left(\frac{PA}{XY}\right)^2+\frac{PB\cdot PC}{AB\cdot AC}=1. \]
[ "Let circle \\(PAB\\) (i.e. the circumcircle of \\(PAB\\)), \\(PAC\\) be \\(\\omega_1, \\omega_2\\) with radii \\(r_1\\), \\(r_2\\) and centers \\(O_1, O_2\\), respectively, and \\(d\\) be the distance between their centers.\n\nLemma. \\(XY = \\frac{r_1 + r_2}{d} \\sqrt{d^2 - (r_1 - r_2)^2}.\\)\n\nProof. Let the external tangent containing \\(X\\) meet \\(\\omega_1\\) at \\(X_1\\) and \\(\\omega_2\\) at \\(X_2\\), and let the external tangent containing \\(Y\\) meet \\(\\omega_1\\) at \\(Y_1\\) and \\(\\omega_2\\) at \\(Y_2\\). Then clearly \\(X_1 Y_1\\) and \\(X_2 Y_2\\) are parallel (for they are both perpendicular \\(O_1 O_2\\)), and so \\(X_1 Y_1 Y_2 X_2\\) is a trapezoid.\n\nNow, \\(X_1 X^2 = XA \\cdot XP = X_2 X^2\\) by Power of a Point, and so \\(X\\) is the midpoint of \\(X_1 X_2\\). Similarly, \\(Y\\) is the midpoint of \\(Y_1 Y_2\\). Hence, \\(XY = \\frac{1}{2} (X_1 Y_1 + X_2 Y_2).\\) Let \\(X_1 Y_1\\), \\(X_2 Y_2\\) meet \\(O_1 O_2\\) s at \\(Z_1, Z_2\\), respectively. Then by similar triangles and the Pythagorean Theorem we deduce that \\(X_1 Z_1 = \\frac{r_1 \\sqrt{d^2 - (r_1 - r_2)^2}}{d}\\) and \\(\\frac{r_2 \\sqrt{d^2 - (r_1 - r_2)^2}}{d}\\). But it is clear that \\(Z_1\\), \\(Z_2\\) is the midpoint of \\(X_1 Y_1\\), \\(X_2 Y_2\\), respectively, so \\(XY = \\frac{(r_1 + r_2)}{d} \\sqrt{d^2 - (r_1 - r_2)^2},\\) as desired.\n\nLemma 2. Triangles \\(O_1 A O_2\\) and \\(BAC\\) are similar.\n\nProof. \\(\\angle{AO_1 O_2} = \\frac{\\angle{PO_1 A}}{2} = \\angle{ABC}\\) and similarly \\(\\angle{AO_2 O_1} = \\angle{ACB}\\), so the triangles are similar by AA Similarity.\n\nAlso, let \\(O_1 O_2\\) intersect \\(AP\\) at \\(Z\\). Then obviously \\(Z\\) is the midpoint of \\(AP\\) and \\(AZ\\) is an altitude of triangle \\(A O_1 O_2\\).Thus, we can simplify our expression of \\(XY\\):\n\n\\[\nXY = \\frac{AB + AC}{BC} \\cdot \\frac{AP}{2 h_a} \\sqrt{BC^2 - (AB - AC)^2},\n\\]\n\nwhere \\(h_a\\) is the length of the altitude from \\(A\\) in triangle \\(ABC\\). Hence, substituting into our condition and using \\(AB = c, BC = a, CA = b\\) gives\n\n\\[\n\\left( \\frac{2a h_a}{(b+c) \\sqrt{a^2 - (b-c)^2}} \\right)^2 + \\frac{PB \\cdot PC}{bc} = 1.\n\\]\n\nUsing \\(2 a h_a = 4[ABC] = \\sqrt{(a + b + c)(a + b - c)(a - b + c)(-a + b + c)}\\) by Heron's Formula (where \\([ABC]\\) is the area of triangle \\(ABC\\), our condition becomes\n\n\\[\n\\frac{(a + b + c)(-a + b + c)}{(b + c)^2} + \\frac{PB \\cdot PC}{bc} = 1,\n\\]\n\nwhich by \\((a + b + c)(-a + b + c) = (b + c)^2 - a^2\\) becomes\n\n\\[\n\\frac{PB \\cdot PC}{bc} = \\frac{a^2 bc}{(b+c)^2}.\n\\]\n\nLet \\(PB = x\\); then \\(PC = a - x\\). The quadratic in \\(x\\) is\n\n\\[\nx^2 - ax + \\frac{a^2 bc}{(b+c)^2} = 0,\n\\]\n\nwhich factors as\n\n\\[\n\\left(x - \\frac{ab}{b+c}\\right)\\left(x - \\frac{ac}{b+c}\\right) = 0.\n\\]\n\nHence, \\(PB = \\frac{ab}{b+c}\\) or \\(\\frac{ac}{b+c}\\), and so the \\(P\\) corresponding to these lengths are our answer.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2014-1
https://artofproblemsolving.com/wiki/index.php/2014_USAMO_Problems/Problem_1
Let \(a,b,c,d\) be real numbers such that \(b-d \ge 5\) and all zeros \(x_1, x_2, x_3,\) and \(x_4\) of the polynomial \(P(x)=x^4+ax^3+bx^2+cx+d\) are real. Find the smallest value the product \((x_1^2+1)(x_2^2+1)(x_3^2+1)(x_4^2+1)\) can take.
[ "Using the hint we turn the equation into \\(\\prod_{k=1} ^4 (x_k-i)(x_k+i) \\implies P(i)P(-i) \\implies (b-d-1)^2 + (a-c)^2 \\implies \\boxed{16}\\). This minimum is achieved when all the \\(x_i\\) are equal to \\(1\\).\n\nA more detailed version goes as follows:\n\nObserve that \\(x^2+1=x^2-(-1)=x^2-i^2=(x-i)(x+i).\\) Now, notice that:\n\n\\[\n\\prod_{k=1} ^4 (x_k-i)(x_k+i)=\\prod_{k=1}^4(x_k-i)\\cdot\\prod_{k=1}^4(x_k+i).\n\\]\n\nThe definition of \\(P(x)\\) is \\((x-x_1)(x-x_2)(x-x_3)(x-x_4)\\) where the leading coeffecient is \\(1\\) since \\(P\\) is a monic polynomial as given in the problem. Then, substituting in \\(i,\\) we have:\n\n\\[\nP(i)=(i-x_1)\\cdots(i-x_4)=(-1)^4(x_1-i)\\cdots(x_4-i).\n\\]\n\nThis is exactly our first product. Our second product can be found as follows:\n\n\\[\nP(-i)=(-i-x_1)\\cdots(-i-x_4)=(-1)^4(x_1+i)\\cdots(x_4+i).\n\\]\n\nHence, what we want to find is \\(P(i)P(-i).\\) But substituting in \\(i\\) into our expressoin \\(x^4+ax^3+bx^2+cx+d\\) gets us \\(P(i)=(1-b+d)+(c-a)i,\\) and \\(P(-i)=(1-b+d)+(a-c)i.\\) Hence:\n\n\\[\nP(i)P(-i)=((1-b+d)+(c-a)i)((1-b+d)+(a-c)i)=(1-b+d)^2-(a-c)^2=(-1)^2(b-d-1)^2-(a-c)^2 \\ge 4^2 -0^2=16\n\\]\n\nwhere equality holds at \\(a=c\\) and \\(b-d=5 \\Rightarrow b=d+5.\\) Hence, the minimum is \\(\\boxed{16}.\\)\n\nTo finish this off we find a construction for this minimum. We know that \\(a=c\\) and \\(b=d+5.\\) Hence\n\n\\[\nP(x)=x^4+cx^3+(d+5)x^2+cx+d.\n\\]\n\nWe set \\(d=1\\) to get as much symmetry as possible within our polynomial. This leads to \\(x^4+cx^3+6x^2+cx+1.\\) However, note that \\(\\binom{4}{2}=6\\) and \\(\\binom{4}{0}=\\binom{4}{4}=1,\\) so with some wishful thinking this leads us to think about the binomial theorem. We can try \\((x\\pm 1)^4\\) and realize that those solutions do work.\n\nHence, \\(16\\) is obtained when all of \\(x_i\\) are equal to \\(1\\) or all equal to \\(-1.\\)\n\n~mathboy282" ]
USAMO-2014-2
https://artofproblemsolving.com/wiki/index.php/2014_USAMO_Problems/Problem_2
Let \(\mathbb{Z}\) be the set of integers. Find all functions \(f : \mathbb{Z} \rightarrow \mathbb{Z}\) such that \[ xf(2f(y)-x)+y^2f(2x-f(y))=\frac{f(x)^2}{x}+f(yf(y)) \] for all \(x, y \in \mathbb{Z}\) with \(x \neq 0\).
[ "Note: This solution is kind of rough. I didn't want to put my 7-page solution all over again. It would be nice if someone could edit in the details of the expansions.\n\nLemma 1: \\(f(0) = 0\\). Proof: Assume the opposite for a contradiction. Plug in \\(x = 2f(0)\\) (because we assumed that \\(f(0) \\neq 0\\)), \\(y = 0\\). What you get eventually reduces to:\n\n\\[\n4f(0)-2 = \\left( \\frac{f(2f(0))}{f(0)} \\right)^2\n\\]\n\nwhich is a contradiction since the LHS is divisible by 2 but not 4.\n\nThen plug in \\(y = 0\\) into the original equation and simplify by Lemma 1. We get:\n\n\\[\nx^2f(-x) = f(x)^2\n\\]\n\nThen:\n\n\\[\n\\begin{align*} x^6f(x) &= x^4\\bigl(x^2f(x)\\bigr)\\\\ &= x^4\\bigl((-x)^2f(-(-x))\\bigr)\\\\ &= x^4(-x)^2f(-(-x))\\\\ &= x^4f(-x)^2\\\\ &= f(x)^4 \\end{align*}\n\\]\n\nTherefore, \\(f(x)\\) must be 0 or \\(x^2\\).\n\nNow either \\(f(x)\\) is \\(x^2\\) for all \\(x\\) or there exists \\(a \\neq 0\\) such that \\(f(a)=0\\). The first case gives a valid solution. In the second case, we let \\(y = a\\) in the original equation and simplify to get:\n\n\\[\nxf(-x) + a^2f(2x) = \\frac{f(x)^2}{x}\n\\]\n\nBut we know that \\(xf(-x) = \\frac{f(x)^2}{x}\\), so:\n\n\\[\na^2f(2x) = 0\n\\]\n\nSince \\(a\\) is not 0, \\(f(2x)\\) is 0 for all \\(x\\) (including 0). Now either \\(f(x)\\) is 0 for all \\(x\\), or there exists some \\(m \\neq 0\\) such that \\(f(m) = m^2\\). Then \\(m\\) must be odd. We can let \\(x = 2k\\) in the original equation, and since \\(f(2x)\\) is 0 for all \\(x\\), stuff cancels and we get:\n\n\\[\ny^2f(4k - f(y)) = f(yf(y))\n\\]\n\nfor \\(\\mathbf{k \\neq 0}\\). Now, let \\(y = m\\) and we get:\n\n\\[\nm^2f(4k - m^2) = f(m^3)\n\\]\n\nNow, either both sides are 0 or both are equal to \\(m^6\\). If both are \\(m^6\\) then:\n\n\\[\nm^2(4k - m^2)^2 = m^6\n\\]\n\nwhich simplifies to:\n\n\\[\n4k - m^2 = \\pm m^2\n\\]\n\nSince \\(k \\neq 0\\) and \\(m\\) is odd, both cases are impossible, so we must have:\n\n\\[\nm^2f(4k - m^2) = f(m^3) = 0\n\\]\n\nThen we can let \\(k\\) be anything except 0, and get \\(f(x)\\) is 0 for all \\(x \\equiv 3 \\pmod{4}\\) except \\(-m^2\\). Also since \\(x^2f(-x) = f(x)^2\\), we have \\(f(x) = 0 \\Rightarrow f(-x) = 0\\), so \\(f(x)\\) is 0 for all \\(x \\equiv 1 \\pmod{4}\\) except \\(m^2\\). So \\(f(x)\\) is 0 for all \\(x\\) except \\(\\pm m^2\\). Since \\(f(m) \\neq 0\\), \\(m = \\pm m^2\\). Squaring, \\(m^2 = m^4\\) and dividing by \\(m\\), \\(m = m^3\\). Since \\(f(m^3) = 0\\), \\(f(m) = 0\\), which is a contradiction for \\(m \\neq 1\\). However, if we plug in \\(x = 1\\) with \\(f(1) = 1\\) and \\(y\\) as an arbitrary large number with \\(f(y) = 0\\) into the original equation, we get \\(0 = 1\\) which is a clear contradiction, so our only solutions are \\(f(x) = 0\\) and \\(f(x) = x^2\\).", "Given that the range of f consists entirely of integers, it is clear that the LHS must be an integer and \\(f(yf(y))\\) must also be an integer, therefore \\(\\frac{f(x)^2}{x}\\) is an integer. If \\(x\\) divides \\(f(x)^2\\) for all integers \\(x \\ne 0\\), then \\(x\\) must be a factor of \\(f(x)\\), therefore \\(f(0)=0\\). Now, by setting \\(y=0\\) in the original equation, this simplifies to \\(xf(-x)=\\frac{f(x)^2}{x}\\). Assuming \\(x \\ne 0\\), we have \\(x^2f(-x)=f(x)^2\\). Substituting in \\(-x\\) for \\(x\\) gives us \\(x^2f(x)=f(-x)^2\\). Substituting in \\(\\frac{f(x)^2}{x^2}\\) in for \\(f(-x)\\) in the second equation gives us \\(x^2f(x)=\\frac{f(x)^4}{x^4}\\), so \\(x^6f(x)=f(x)^4\\). In particular, if \\(f(x) \\ne 0\\), then we have \\(f(x)^3=x^6\\), therefore \\(f(x)\\) is equivalent to \\(0\\) or \\(x^2\\) for every integer \\(x\\). Now, we shall prove that if for some integer \\(t \\ne 0\\), if \\(f(t)=0\\), then \\(f(x)=0\\) for all integers \\(x\\). If we assume \\(f(y)=0\\) and \\(y \\ne 0\\) in the original equation, this simplifies to \\(xf(-x)+y^2f(2x)=\\frac{f(x)^2}{x}\\). However, since \\(x^2f(-x)=f(x)^2\\), we can rewrite this equation as \\(\\frac{f(x)^2}{x}+y^2f(2x)=\\frac{f(x)^2}{x}\\), \\(y^2f(2x)\\) must therefore be equivalent to \\(0\\). Since, by our initial assumption, \\(y \\ne 0\\), this means that \\(f(2x)=0\\), so, if for some integer \\(y \\ne 0\\), \\(f(y)=0\\), then \\(f(x)=0\\) for all integers \\(x\\). The contrapositive must also be true, i.e. If \\(f(x) \\ne 0\\) for all integers \\(x\\), then there is no integral value of \\(y \\ne 0\\) such that \\(f(y)=0\\), therefore \\(f(x)\\) must be equivalent for \\(x^2\\) for every integer \\(x\\), including \\(0\\), since \\(f(0)=0\\). Thus, \\(f(x)=0, x^2\\) are the only possible solutions.", "Let's assume \\(f(0)\\neq 0.\\) Substitute \\((x,y)=(2f(0),0)\\) to get\n\n\\[\n2f(0)^2=f(2f(0))^2/2f(0)+f(0)\n\\]\n\n\\[\n2f(0)^2(2f(0)-1)=f(2f(0))^2\n\\]\n\nThis means that \\(2(2f(0)-1)\\) is a perfect square. However, this is impossible, as it is equivalent to \\(2\\pmod{4}.\\) Therefore, \\(f(0)=0.\\) Now substitute \\(x\\neq 0, y=0\\) to get\n\n\\[\nxf(-x)=\\frac{f(x)^2}{x} \\implies x^2f(-x)=f(x)^2.\n\\]\n\nSimilarly,\n\n\\[\nx^2f(x)=f(-x)^2.\n\\]\n\nFrom these two equations, we can find either \\(f(x)=f(-x)=0,\\) or \\(f(x)=f(-x)=x^2.\\) Both of these are valid solutions on their own, so let's see if there are any solutions combining the two.\n\nLet's say we can find \\(f(x)=x^2, f(y)=0,\\) and \\(x,y\\neq 0.\\) Then\n\n\\[\nxf(-x)+y^2f(2x)=f(x)^2/x.\n\\]\n\n\\[\ny^2f(2x)=x-x^3.\n\\]\n\n(NEEDS FIXING: \\(f(x)^2/x= x^4/x = x^3\\), so the RHS is \\(0\\) instead of \\(x-x^3\\).)\n\nIf \\(f(2x)=4x^2,\\) then \\(y^2=\\frac{x-x^3}{4x^2}=\\frac{1-x^2}{4x},\\) which is only possible when \\(y=0.\\) This contradicts our assumption. Therefore, \\(f(2x)=0.\\) This forces \\(x=\\pm 1\\) due to the right side of the equation. Let's consider the possibility \\(f(2)=0, f(1)=1.\\) Substituting \\((x,y)=(2,1)\\) into the original equation yields\n\n\\[\n0=2f(0)+1f(2)=0+f(1)=1,\n\\]\n\nwhich is impossible. So \\(f(2)=f(-2)=4\\) and there are no solutions \"combining\" \\(f(x)=x^2\\) and \\(f(x)=0.\\)\n\nTherefore our only solutions are \\(\\boxed{f(x)=0}\\) and \\(\\boxed{f(x)=x^2.}\\)", "Let the given assertion be \\(P(x, y)\\). We try \\(P(x, 0)\\) and get \\(xf(2c-x)=f(x)^2/x+c\\), where \\(f(0)=c\\). We plug in \\(x=c\\) and get \\(cf(c)=f(c)^2/c+c\\). Rearranging and solving for \\(c^2\\) gives us \\(c^2=\\frac{f(c)^2}{f(c)-1}\\). Obviously, the only \\(c\\) that works such that the RHS is an integer is \\(c=0\\), and thus \\(f(0)=0\\).\n\nWe use this information on assertion \\(P(x,0)\\) and obtain \\(xf(-x)=f(x^2)/x\\), or \\(f(-x)=\\frac{f(x)^2}{x^2}\\). Thus, \\(f(x)\\) is an even function. It follows that \\(f(x)=0, x^2\\) for each \\(x\\). We now prove that \\(f(x)=x^2\\), f(x)=0$ are the only solutions. [in progress] ~SigmaPiE" ]
USAMO-2014-3
https://artofproblemsolving.com/wiki/index.php/2014_USAMO_Problems/Problem_3
Prove that there exists an infinite set of points \[ \ldots,\,\,\,\,P_{-3},\,\,\,\,P_{-2},\,\,\,\,P_{-1},\,\,\,\,P_0,\,\,\,\,P_1,\,\,\,\,P_2,\,\,\,\,P_3,\,\,\,\,\ldots \] in the plane with the following property: For any three distinct integers \(a,b,\) and \(c\), points \(P_a\), \(P_b\), and \(P_c\) are collinear if and only if \(a+b+c=2014\).
[ "Consider an elliptic curve with a generator \\(g\\), such that \\(g\\) is not a root of \\(0\\). By repeatedly adding \\(g\\) to itself under the standard group operation, with can build \\(g, 2g, 3g, \\ldots\\) as well as \\(-g, -2g, -3g, \\ldots\\). If we let\n\n\\[\nP_k = (3k-2014)g\n\\]\n\nthen we can observe that collinearity between \\(P_a\\), \\(P_b\\), and \\(P_c\\) occurs only if \\(P_a + P_b + P_c = 0\\) (by definition of the group operation), which is equivalent to \\((3a-2014)g + (3b-2014)g + (3c-2014)g = (3a+3b+3c-3*2014)g = 0\\), or \\(3a + 3b + 3c = 3*2014\\), or \\(a + b + c = 2014\\). We know that all these points \\(P_k\\) exist because \\(3k-2014\\) is never 0 for integer \\(k\\), so that none of these points need to be point at infinity (the identity element of the group).", "Consider letting \\(P_x\\) be the point \\((x, f(x))\\), where \\(f(x) = x^3 - 2014x^2\\). Then if three points \\(P_a, P_b, P_c\\) are on the same line \\(y = mx + p\\), they must be the solutions to the equation \\(x^3 - 2014x^2 = mx + p\\) (i.e. the intersection of \\(f\\) and the line). By Vieta's Formulas, the sum of the roots to this equation is 2014. Because each value of x corresponds to a different one of a, b, and c by definition of \\(P_x\\), \\(a + b + c = 2014\\). Conversely, if \\(a + b + c = 2014\\), they must be the solutions to \\((x-a)(x-b)(x-c) = x^3 - 2014x^2 - mx - p = 0\\) for some real \\(m\\) and \\(p\\). Clearly, then, \\(P_a, P_b, P_c\\) must all lie on the line \\(y = mx + p\\). Hence, our setting \\(P_x = f(x)\\) produces a valid infinite set of points.\n\nNote: We could have let \\(f(x) = ax^3 - 2014ax^2 + bx + c\\), where a, b, and c are arbitrary constants. (a is nonzero.)" ]
USAMO-2014-4
https://artofproblemsolving.com/wiki/index.php/2014_USAMO_Problems/Problem_4
Let \(k\) be a positive integer. Two players \(A\) and \(B\) play a game on an infinite grid of regular hexagons. Initially all the grid cells are empty. Then the players alternately take turns with \(A\) moving first. In his move, \(A\) may choose two adjacent hexagons in the grid which are empty and place a counter in both of them. In his move, \(B\) may choose any counter on the board and remove it. If at any time there are \(k\) consecutive grid cells in a line all of which contain a counter, \(A\) wins. Find the minimum value of \(k\) for which \(A\) cannot win in a finite number of moves, or prove that no such minimum value exists.
[ "We claim that the minimum \\(k\\) such that A cannot create a \\(k\\) in a row is \\(\\boxed{6}\\).\n\nIt is easy to verify that player A can create a 5 in a row.\n\nLet A place two counters anywhere, and B take off one of them. Then, A should create a \"triangle of hexagons\" by placing two adjacent counters also next to the unremoved one, so that no matter what B does there will be a line of two counters on adjacent hexagons.\n\nNote: A diagram and more conciseness is required in the following solution. Better maneuvers are always appreciated.\n\nThen, it is advisable for A to make a 4-in-a-row by placing 2 counters at the end of the line. B to prevent a 5-in-a-row must counter by removing a middle counter (a counter not on the edge of the 4-in-a-row). Player A then could reinstate the threat by placing a counter there and one adjacent to that hexagon (but not adjacent to the other middle counter). It is not in B's best interest to then remove the other middle counter, for then A can \"add\" a counter to both hexagons (adjacent to the removed counter) on the same side of the originally placed hexagon while preserving the threat. Now, we have two different threats B cannot both counter. Thus, assume B does not remove the other middle counter. Eventually, by \"adding\" counters and preserving the threat player A can create a web of hexagons surrounding the chosen middle counter. As we have proven, it will be disastrous for B to then remove the other middle counter; thus, B has to remove the chosen middle counter.\n\nNow, player A can administer the decisive winning maneuver as follows. Let X be two adjacent counters of the web, one of which is adjacent to the initial edge counter. Let Y be the other two counters that satisfy the condition. The line X should not have more than two counters on it. Player A should place his two counters on the first two hexagons not adjacent to any counter in X (or the web), so that B is forced to remove the outermost counter A played. Then, A should place two counters adjacent to his remaining counter in a direction parallel to line Y. B has to expunge the outermost one; otherwise A can place two counters to complete a 5-in-a-row. Let A place two more counters in the same direction; then, B has to remove the middle counter played. Now, turn in a direction parallel to line X; let A place a counter on the first two spaces adjacent to counters already placed. B will remove the second counter played. A's knockout blow occurs when he reinstates the lost counter and a counter adjacent to that counter and the remaining one he played last turn, and it is easy to verify he is guaranteed a 5-in-a-row. Therefore, A can (after an elongated siege) obtain a 5-in-a-row.\n\nTo prove that 6-in-a-row is impossible, tile the hexagon board in an A-B-C-A-B-C-A-B-C format. Define an \"A tile\" to be a tile engraved with the letter A. To prevent player A from obtaining a 6 in a row, all player B has to do is to remove counters placed on A tiles by player A. Because no two A tiles are adjacent, player A can play at most one counter on an A tile at a time; thus, there can at most be one counter on A tiles any time of the game. Clearly, then, a 6 in a row, which requires A-B-C-A-B-C (or two counters on A tiles), is impossible, completing the proof." ]
USAMO-2014-5
https://artofproblemsolving.com/wiki/index.php/2014_USAMO_Problems/Problem_5
Let \(ABC\) be a triangle with orthocenter \(H\) and let \(P\) be the second intersection of the circumcircle of triangle \(AHC\) with the internal bisector of the angle \(\angle BAC\). Let \(X\) be the circumcenter of triangle \(APB\) and \(Y\) the orthocenter of triangle \(APC\). Prove that the length of segment \(XY\) is equal to the circumradius of triangle \(ABC\).
[ "Let \\(O_1\\) be the center of \\((AHPC)\\), \\(O\\) be the center of \\((ABC)\\). Note that \\((O_1)\\) is the reflection of \\((O)\\) across \\(AC\\), so \\(AO=AO_1\\). Additionally\n\n\\[\n\\angle AYC=180-\\angle APC=180-\\angle AHC=\\angle B\n\\]\n\nso \\(Y\\) lies on \\((O)\\). Now since \\(XO,OO_1,XO_1\\) are perpendicular to \\(AB,AC,\\) and their bisector, \\(XOO_1\\) is isosceles with \\(XO=OO_1\\), and \\(\\angle XOO_1=180-\\angle A\\). Also\n\n\\[\n\\angle AOY=2\\angle ACY=2(90-\\angle PAC)=2(90-\\frac{A}{2})=180-\\angle A = \\angle XOO_1\n\\]\n\nBut \\(YO=OA\\) as well, and \\(\\angle YOX=\\angle AOO_1\\), so \\(\\triangle OYX\\cong \\triangle OAO_1\\). Thus \\(XY=AO_1=AO\\).", "Since \\(AHPC\\) is a cyclic quadrilateral, \\(\\angle AHC = \\angle APC\\). \\(\\angle AHC = 90^\\circ + \\angle ABC\\) and \\(\\angle APC = 90^\\circ + \\angle AYC\\), we find \\(\\angle ABC = \\angle AYC\\). That is, \\(ABYC\\) is a cyclic quadrilateral. Let \\(D\\) be mid-point of \\(\\overline{AB}\\). \\(O, X, D\\) are collinear and \\(OX \\perp AB\\). Let \\(M\\) be second intersection of \\(AP\\) with circumcircle of the triangle \\(ABC\\). Let \\(YP \\cap AC = E\\), \\(YM \\cap AB = F\\). Since \\(M\\) is mid-point of the arc \\(BC\\), \\(OM\\perp BC\\). Since \\(AYMC\\) is a cyclic quadrilateral, \\(\\angle CYM = \\angle CAM = \\angle BAC /2\\). Since \\(Y\\) is the orthocenter of triangle \\(APC\\), \\(\\angle PYC = \\angle CAP = \\angle BAC /2\\). Thus, \\(\\angle PYM = \\angle BAC\\) and \\(AEYF\\) is a cyclic quadrilateral. So, \\(YF \\perp AB\\) and \\(OX \\parallel MY\\). We will prove that \\(XYMO\\) is a parallelogram.\n\nhttps://wiki-images.artofproblemsolving.com//7/7b/Usamo2014-5.png (figure link)\n\nWe see that \\(YPM\\) is an isosceles triangle and \\(YM=YP\\). Also \\(XB=XP\\) and \\(\\angle BXP = 2\\angle BAP = \\angle BAC = \\angle PYM\\). Then, \\(BXP \\sim MYP\\). By spiral similarity, \\(BPM \\sim XPY\\) and \\(\\angle XYP = \\angle BMP = \\angle BCA\\). Hence, \\(\\angle XYP = \\angle BCA\\), \\(XY \\perp BC\\). Since \\(OM \\perp BC\\), we get \\(XYMO\\) is a parallelogram. As a result, \\(OM = XY\\).\n\n(Lokman GÖKÇE)" ]
USAMO-2014-6
https://artofproblemsolving.com/wiki/index.php/2014_USAMO_Problems/Problem_6
Prove that there is a constant \(c>0\) with the following property: If \(a, b, n\) are positive integers such that \(\gcd(a+i, b+j)>1\) for all \(i, j\in\{0, 1, \ldots n\}\), then \[ \min\{a, b\}>c^n\cdot n^{\frac{n}{2}}. \]
[ "The following solution is due to Gabriel Dospinescu and v_Enhance (also known as Evan Chen).\n\nLet \\(N = n+1\\) and assume \\(N\\) is (very) large. We construct an \\(N \\times N\\) with cells \\((i,j)\\) where \\(0 \\le i, j \\le n\\) and in each cell place a prime \\(p\\) dividing \\(\\gcd (a+i, b+j)\\).\n\nThe central claim is at least \\(50\\%\\) of the primes in this table exceed \\(0.001n^2\\). We count the maximum number of squares they could occupy:\n\n\\[\n\\sum_p \\left\\lceil \\frac{N}{p} \\right\\rceil^2 \t\t\\le \\sum_p \\left( \\frac Np + 1 \\right)^2 \\\\ \t\t= N^2 \\sum_p \\frac{1}{p^2} + 2N \\sum_p \\frac1p + \\sum_p 1.\n\\]\n\nHere the summation runs over primes \\(p \\le 0.001n^2\\).\n\nLet \\(r = \\pi(0.001n^2)\\) denote the number of such primes. Now we apply the three bounds:\n\n\\[\n\\sum_p \\frac{1}{p^2} < \\frac 12\n\\]\n\nwhich follows by adding all the primes directly with some computation,\n\n\\[\n\\sum_p \\frac 1p < \\sum_{k=1}^r \\frac 1k = O(\\ln r) < o(N)\n\\]\n\nusing the harmonic series bound, and\n\n\\[\n\\sum_p 1 < r \\sim O \\left( \\frac{N^2}{\\ln N} \\right) < o(N^2)\n\\]\n\nvia Prime Number Theorem. Hence the sum in question is certainly less than \\(\\tfrac 12 N^2\\) for \\(N\\) large enough, establishing the central claim.\n\nHence some column \\(a+i\\) has at least one half of its primes greater than \\(0.001n^2\\). Because this is greater than \\(n\\) for large \\(n\\), these primes must all be distinct, so \\(a+i\\) exceeds their product, which is larger than\n\n\\[\n\\left( 0.001n^2 \\right)^{N/2} > c^n \\cdot n^n\n\\]\n\nwhere \\(c\\) is some constant." ]
USAMO-2015-1
https://artofproblemsolving.com/wiki/index.php/2015_USAMO_Problems/Problem_1
Solve in integers the equation \[ x^2+xy+y^2 = \left(\frac{x+y}{3}+1\right)^3. \]
[ "We first notice that both sides must be integers, so \\(\\frac{x+y}{3}\\) must be an integer.\n\nWe can therefore perform the substitution \\(x+y = 3t\\) where \\(t\\) is an integer.\n\nThen:\n\n\\[\n(3t)^2 - xy = (t+1)^3\n\\]\n\n\\[\n9t^2 + x (x - 3t) = t^3 + 3t^2 + 3t + 1\n\\]\n\n\\[\n4x^2 - 12xt + 9t^2 = 4t^3 - 15t^2 + 12t + 4\n\\]\n\n\\[\n(2x - 3t)^2 = (t - 2)^2(4t + 1)\n\\]\n\n\\(4t+1\\) is therefore the square of an odd integer and can be replaced with \\((2n+1)^2 = 4n^2 + 4n +1\\)\n\nBy substituting using \\(t = n^2 + n\\) we get:\n\n\\[\n(2x - 3n^2 - 3n)^2 = [(n^2 + n - 2)(2n+1)]^2\n\\]\n\n\\[\n2x - 3n^2 - 3n = \\pm (2n^3 + 3n^2 -3n -2)\n\\]\n\n\\(x = n^3 + 3n^2 - 1\\) or \\(x = -n^3 + 3n + 1\\)\n\nUsing substitution we get the solutions: \\((n^3 + 3n^2 - 1, -n^3 + 3n + 1) \\cup (-n^3 + 3n + 1, n^3 + 3n^2 - 1)\\)", "Let \\(n = \\frac{x+y}{3}\\). Thus, \\(x+y = 3n\\). We have\n\n\\[\nx^2+xy+y^2 = \\left(\\frac{x+y}{3}+1\\right)^3 \\implies (x+y)^2 - xy = \\left(\\frac{x+y}{3}+1\\right)^3\n\\]\n\nSubstituting \\(n\\) for \\(\\frac{x+y}{3}\\), we have\n\n\\[\n9n^2 - x(3n-x) = (n+1)^3\n\\]\n\nTreating \\(x\\) as a variable and \\(n\\) as a constant, we have\n\n\\[\n9n^2 - 3nx + x^2 = (n+1)^3,\n\\]\n\nwhich turns into\n\n\\[\nx^2 - 3nx + (9n^2 - (n+1)^3) = 0,\n\\]\n\na quadratic equation. By the quadratic formula,\n\n\\[\nx = \\frac{1}{2} \\left(3n \\pm \\sqrt{9n^2 - 4(9n^2 - (n+1)^3)} \\right)\n\\]\n\nwhich simplifies to\n\n\\[\nx = \\frac{1}{2} \\left(3n \\pm \\sqrt{4(n+1)^3 - 27n^2} \\right)\n\\]\n\nSince we want \\(x\\) and \\(y\\) to be integers, we need \\(4(n+1)^3 - 27n^2\\) to be a perfect square. We can factor the aforementioned equation to be\n\n\\[\n(n-2)^2 (4n+1) = k^2\n\\]\n\nfor an integer \\(k\\). Since \\((n-2)^2\\) is always a perfect square, for \\((n-2)^2 (4n+1)\\) to be a perfect square, \\(4n + 1\\) has to be a perfect square as well. Since \\(4n + 1\\) is odd, the square root of the aforementioned equation must be odd as well. Thus, we have \\(4n + 1 = a^2\\) for some odd \\(a\\). Thus,\n\n\\[\nn = \\frac{a^2 - 1}{4},\n\\]\n\nin which by difference of squares it is easy to see that all the possible values for \\(n\\) are just \\(n = p(p-1)\\), where \\(p\\) is a positive integer. Thus,\n\n\\[\nx+y = 3n = 3p(p-1).\n\\]\n\nThus, the general form for\n\n\\[\nx = \\frac{1}{2} \\left(3p(p-1) \\pm \\sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \\right)\n\\]\n\nfor a positive integer \\(p\\). (This is an integer since \\(4(p(p-1)+1)^3 - 27(p(p-1))^2\\) is an even perfect square (since \\(4(p(p-1)+1)\\) is always even, as well as \\(27(p(p-1))^2\\) being always even) as established, and \\(3p(p-1)\\) is always even as well. Thus, the whole numerator is even, which makes the quantity of that divided by \\(2\\) always an integer.) Since \\(y = 3n - x\\), the general form for \\(y\\) is just\n\n\\[\ny = 3p(p-1) - \\frac{1}{2} \\left(3p(p-1) \\pm \\sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \\right)\n\\]\n\n(This is an integer since \\(4(p(p-1)+1)^3 - 27(p(p-1))^2\\) is an even perfect square (since \\(4(p(p-1)+1)\\) is always even, as well as \\(27(p(p-1))^2\\) being always even) as established, and \\(3p(p-1)\\) is always even as well. Thus, the whole numerator is even, which makes the quantity of that divided by \\(2\\) always an integer, which thus trivially makes\n\n\\[\n3p(p-1) - \\frac{1}{2} \\left(3p(p-1) \\pm \\sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \\right)\n\\]\n\nan integer.) for a positive integer \\(p\\). Thus, our general in integers \\((x, y)\\) is\n\n\\[\n(\\frac{1}{2} \\left(3p(p-1) \\pm \\sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \\right), 3p(p-1) - \\frac{1}{2} \\left(3p(p-1) \\pm \\sqrt{4(p(p-1)+1)^3 - 27(p(p-1))^2} \\right).\n\\]\n\n\\(\\boxed{}\\)\n\n-fidgetboss_4000" ]
USAMO-2015-2
https://artofproblemsolving.com/wiki/index.php/2015_USAMO_Problems/Problem_2
Quadrilateral \(APBQ\) is inscribed in circle \(\omega\) with \(\angle P = \angle Q = 90^{\circ}\) and \(AP = AQ < BP\). Let \(X\) be a variable point on segment \(\overline{PQ}\). Line \(AX\) meets \(\omega\) again at \(S\) (other than \(A\)). Point \(T\) lies on arc \(AQB\) of \(\omega\) such that \(\overline{XT}\) is perpendicular to \(\overline{AX}\). Let \(M\) denote the midpoint of chord \(\overline{ST}\). As \(X\) varies on segment \(\overline{PQ}\), show that \(M\) moves along a circle.
[ "We will use coordinate geometry.\n\nWithout loss of generality, let the circle be the unit circle centered at the origin,\n\n\\[\nA=(1,0) P=(1-a,b), Q=(1-a,-b)\n\\]\n\n, where \\((1-a)^2+b^2=1\\).\n\nLet angle \\(\\angle XAB=A\\), which is an acute angle, \\(\\tan{A}=t\\), then \\(X=(1-a,at)\\).\n\nAngle \\(\\angle BOS=2A\\), \\(S=(-\\cos(2A),\\sin(2A))\\). Let \\(M=(u,v)\\), then \\(T=(2u+\\cos(2A), 2v-\\sin(2A))\\).\n\nThe condition \\(TX \\perp AX\\) yields: \\((2v-\\sin(2A)-at)/(2u+\\cos(2A)+a-1)=\\cot A.\\) (E1)\n\nUse identities \\((\\cos A)^2=1/(1+t^2)\\), \\(\\cos(2A)=2(\\cos A)^2-1= 2/(1+t^2) -1\\), \\(\\sin(2A)=2\\sin A\\cos A=2t^2/(1+t^2)\\), we obtain \\(2vt-at^2=2u+a\\). (E1')\n\nThe condition that \\(T\\) is on the circle yields \\((2u+\\cos(2A))^2+ (2v-\\sin(2A))^2=1\\), namely \\(v\\sin(2A)-u\\cos(2A)=u^2+v^2\\). (E2)\n\n\\(M\\) is the mid-point on the hypotenuse of triangle \\(STX\\), hence \\(MS=MX\\), yielding \\((u+\\cos(2A))^2+(v-\\sin(2A))^2=(u+a-1)^2+(v-at)^2\\). (E3)\n\nExpand (E3), using (E2) to replace \\(2(v\\sin(2A)-u\\cos(2A))\\) with \\(2(u^2+v^2)\\), and using (E1') to replace \\(a(-2vt+at^2)\\) with \\(-a(2u+a)\\), and we obtain \\(u^2-u-a+v^2=0\\), namely \\((u-\\frac{1}{2})^2+v^2=a+\\frac{1}{4}\\), which is a circle centered at \\((\\frac{1}{2},0)\\) with radius \\(r=\\sqrt{a+\\frac{1}{4}}\\).", "Let the midpoint of \\(AO\\) be \\(K\\). We claim that \\(M\\) moves along a circle with radius \\(KP\\).\n\nWe will show that \\(KM^2 = KP^2\\), which implies that \\(KM = KP\\), and as \\(KP\\) is fixed, this implies the claim.\n\n\\(KM^2 = \\frac{AM^2+OM^2}{2}-\\frac{AO^2}{4}\\) by the median formula on \\(\\triangle AMO\\).\n\n\\(KP^2 = \\frac{AP^2+OP^2}{2}-\\frac{AO^2}{4}\\) by the median formula on \\(\\triangle APO\\).\n\n\\(KM^2-KP^2 = \\frac{1}{2}(AM^2+OM^2-AP^2-OP^2)\\).\n\nAs \\(OP = OT\\), \\(OP^2-OM^2 = MT^2\\) from right triangle \\(OMT\\). \\((1)\\)\n\nBy \\((1)\\), \\(KM^2-KP^2 = \\frac{1}{2}(AM^2-MT^2-AP^2)\\).\n\nSince \\(M\\) is the circumcenter of \\(\\triangle XTS\\), and \\(MT\\) is the circumradius, the expression \\(AM^2-MT^2\\) is the power of point \\(A\\) with respect to \\((XTS)\\). However, as \\(AX*AS\\) is also the power of point \\(A\\) with respect to \\((XTS)\\), this implies that \\(AM^2-MT^2=AX*AS\\). \\((2)\\)\n\nBy \\((2)\\), \\(KM^2-KP^2 = \\frac{1}{2}(AX*AS-AP^2)\\)\n\nFinally, \\(\\triangle APX \\sim \\triangle ASP\\) by AA similarity (\\(\\angle XAP = \\angle SAP\\) and \\(\\angle APX = \\angle AQP = \\angle ASP\\)), so \\(AX*AS = AP^2\\). \\((3)\\)\n\nBy \\((3)\\), \\(KM^2-KP^2=0\\), so \\(KM^2=KP^2\\), as desired. \\(QED\\)", "To begin with, we connect \\(\\overline{AT}\\) and we construct the nine-point circle of \\(\\triangle AST\\) centered at \\(N_9\\).\n\nLemma \\(1\\): \\(AX \\cdot AS = AP^2\\). We proceed on a directed angle chase. We get \\(\\measuredangle ASP = \\measuredangle AQP = \\measuredangle QPA\\), so \\(\\triangle PAS \\sim \\triangle XAP\\) and the desired result follows by side length ratios.\n\nLemma \\(2\\): The locus of \\(N_9\\) as \\(X\\) moves along \\(\\overline{PQ}\\) is a circle centered about \\(A\\). We add the midpoint of \\(\\overline{AS}\\), \\(N\\), and let the circumradius of \\(\\triangle AST\\) be \\(R\\). Taking the power of \\(A\\) with respect to \\((N_9)\\), we get\n\n\\[\nAN_9^2 - \\left(\\frac{1}{2} R\\right)^2 = \\text{Pow}_{(N_9)} A = AX \\cdot AN = \\frac{1}{2} AX \\cdot AS = \\frac{1}{2} AP^2.\n\\]\n\nHence, \\(AN_9 = \\sqrt{\\frac{1}{4}R^2 + \\frac{1}{2}AP^2}\\), which remains constant as \\(X\\) moves.\n\nNext, consider the homothety of scale factor \\(\\frac{2}{3}\\) about \\(O\\) mapping \\(N_9\\) to \\(G\\). This means that the locus of \\(G\\) is a circle as well.\n\nFinally, we take a homothety of scale factor \\(\\frac{3}{2}\\) about \\(A\\) mapping \\(G\\) to \\(M\\). Hence, the locus of \\(M\\) is a circle, as desired. - Spacesam" ]
USAMO-2015-3
https://artofproblemsolving.com/wiki/index.php/2015_USAMO_Problems/Problem_3
Let \(S = \{1, 2, ..., n\}\), where \(n \ge 1\). Each of the \(2^n\) subsets of \(S\) is to be colored red or blue. (The subset itself is assigned a color and not its individual elements.) For any set \(T \subseteq S\), we then write \(f(T)\) for the number of subsets of T that are blue. Determine the number of colorings that satisfy the following condition: for any subsets \(T_1\) and \(T_2\) of \(S\), \[ f(T_1)f(T_2) = f(T_1 \cup T_2)f(T_1 \cap T_2). \]
[ "Define function: \\(C(T)=1\\) if the set T is colored blue, and \\(C(T)=0\\) if \\(T\\) is colored red. Define the \\(\\text{Core} =\\text{intersection of all } T \\text{ where } C(T)=1\\).\n\nThe empty set is denoted as \\(\\varnothing\\), \\(\\cap\\) denotes intersection, and \\(\\cup\\) denotes union. Let \\(S_n=\\{n\\}\\) are one-element subsets.\n\nLet \\(m_{c_k}=\\dbinom{m}{k} = \\frac{m!}{k!(m-k)!}\\) denote m choose k.\n\n(Case I) \\(f(\\varnothing)=1\\). Then for distinct m and k, \\(f(S_m \\cup S_k)=f(S_m)f(S_k)\\), meaning only if \\(S_m\\) and \\(S_k\\) are both blue is their union blue. Namely \\(C(S_m \\cup S_k)=C(S_m)C(S_k).\\)\n\nSimilarly, for distinct \\(m,n,k\\), \\(f(S_m \\cup S_k \\cup Sn)=f(S_m \\cup S_k)f(S_n)\\), \\(C(S_m \\cup S_k \\cup S_n)=C(S_m)C(S_k)C(S_n)\\). This procedure of determination continues to \\(S\\). Therefore, if \\(T=\\{a_1,a_2, \\cdots a_k\\}\\), then \\(C(T)=C(S_{a_1})C(S_{a_2}) \\cdots C(S_{a_k})\\). All colorings thus determined by the free colors chosen for subsets of one single elements satisfy the condition. There are \\(2^n\\) colorings in this case.\n\n(Case II.) \\(f(\\varnothing)=0\\).\n\n(Case II.1) \\(\\text{Core}=\\varnothing\\). Then either (II.1.1) there exist two nonintersecting subsets A and B, \\(C(A)=C(B)=1\\), but f\\((A)f(B)=0\\), a contradiction, or (II.1.2) all subsets has \\(C(T)=0\\), which is easily confirmed to satisfy the condition \\(f(T_1)f(T_2)=f(T_1 \\cap T_2)f(T_1 \\cup T_2)\\). There is one coloring in this case.\n\n(Case II.2) Core = a subset of 1 element. WLOG, \\(C(S_1)=1\\). Then \\(f(S_1)=1\\), and subsets containing element 1 may be colored blue. \\(f(S_1 \\cup S_m)f(S_1\\cup S_n)=f(S_1 \\cup S_m \\cup S_n)\\), namely \\(C(S_1 \\cup S_m \\cup S_n)=C(S_m \\cup S_1)C(S_n \\cup S_1)\\). Now S_1 functions as the \\(\\varnothing\\) in case I, with \\(n-1\\) elements to combine into a base of \\(n-1\\) two-element sets, and all the other subsets are determined. There are \\(2^{n-1}\\) colorings for each choice of core. However, there are nC1 = n such cores. Hence altogether there are \\(n2^{n-1}\\) colorings in this case.\n\n(Case II.3) Core = a subset of 2 elements. WLOG, let \\(C(S_1 \\cup S_2)=1\\). Only subsets containing the core may be colored blue. With the same reasoning as in the preceding case, there are \\((nC2)2^{n-2}\\) colorings.\n\n\\[\n\\dots\n\\]\n\n(Case II.n+1) Core = S. Then \\(C(S)=1\\), with all other subsets \\(C(T)=0\\), there is \\(1=\\dbinom{n}{n}2^0\\)\n\nCombining all the cases, we have \\(1+\\left[2^n+\\dbinom{n}{1}2^{n-1}+\\dbinom{n}{2}2^{n-2}+ \\cdots + \\dbinom{n}{n}2^0\\right]=\\boxed{1+3^n}\\) colorings. sponsored by ALLEN" ]
USAMO-2015-4
https://artofproblemsolving.com/wiki/index.php/2015_USAMO_Problems/Problem_4
Steve is piling \(m\geq 1\) indistinguishable stones on the squares of an \(n\times n\) grid. Each square can have an arbitrarily high pile of stones. After he finished piling his stones in some manner, he can then perform stone moves, defined as follows. Consider any four grid squares, which are corners of a rectangle, i.e. in positions \((i, k), (i, l), (j, k), (j, l)\) for some \(1\leq i, j, k, l\leq n\), such that \(i<j\) and \(k<l\). A stone move consists of either removing one stone from each of \((i, k)\) and \((j, l)\) and moving them to \((i, l)\) and \((j, k)\) respectively,j or removing one stone from each of \((i, l)\) and \((j, k)\) and moving them to \((i, k)\) and \((j, l)\) respectively. Two ways of piling the stones are equivalent if they can be obtained from one another by a sequence of stone moves. How many different non-equivalent ways can Steve pile the stones on the grid?
[ "Let the number of stones in row \\(i\\) be \\(r_i\\) and let the number of stones in column \\(i\\) be \\(c_i\\). Since there are \\(m\\) stones, we must have \\(\\sum_{i=1}^n r_i=\\sum_{i=1}^n c_i=m\\)\n\nLemma 1: If any \\(2\\) pilings are equivalent, then \\(r_i\\) and \\(c_i\\) are the same in both pilings \\(\\forall i\\).\n\nProof: We suppose the contrary. Note that \\(r_i\\) and \\(c_i\\) remain invariant after each move, therefore, if any of the \\(r_i\\) or \\(c_i\\) are different, they will remain different.\n\nLemma 2: Any \\(2\\) pilings with the same \\(r_i\\) and \\(c_i\\) \\(\\forall i\\) are equivalent.\n\nProof: Suppose piling 1 and piling 2 not the same piling. Call a stone in piling 1 wrong if the stone occupies a position such that there are more stones in that position in piling 1 than piling 2. Similarly define a wrong stone in piling 2. Let a wrong stone be at \\((a, b)\\) in piling 1. Since \\(c_b\\) is the same for both pilings, we must have a wrong stone in piling 2 at column b, say at \\((c, b)\\), such that \\(c\\not = a\\). Similarly, we must have a wrong stone in piling 1 at row c, say at \\((c, d)\\) where \\(d \\not = b\\). Clearly, making the move \\((a,b);(c,d) \\implies (c,b);(a,d)\\) in piling 1 decreases the number of wrong stones in piling 1. Therefore, the number of wrong stones in piling 1 must eventually be \\(0\\) after a sequence of moves, so piling 1 and piling 2 are equivalent.\n\nLemma 3: Given the sequences \\(g_i\\) and \\(h_i\\) such that \\(\\sum_{i=1}^n g_i=\\sum_{i=1}^n h_i=m\\) and \\(g_i, h_i\\geq 0 \\forall i\\), there is always a piling that satisfies \\(r_i=g_i\\) and \\(c_i=h_i\\) \\(\\forall i\\).\n\nProof: We take the lowest \\(i\\), \\(j\\), such that \\(g_i, h_j >0\\) and place a stone at \\((i, j)\\), then we subtract \\(g_i\\) and \\(h_j\\) by \\(1\\) each, until \\(g_i\\) and \\(h_i\\) become \\(0\\) \\(\\forall i\\), which will happen when \\(m\\) stones are placed, because \\(\\sum_{i=1}^n g_i\\) and \\(\\sum_{i=1}^n h_i\\) are both initially \\(m\\) and decrease by \\(1\\) after each stone is placed. Note that in this process \\(r_i+g_i\\) and \\(c_i+h_i\\) remains invariant, thus, the final piling satisfies the conditions above.\n\nBy the above lemmas, the number of ways to pile is simply the number of ways to choose the sequences \\(r_i\\) and \\(c_i\\) such that \\(\\sum_{i=1}^n r_i=\\sum_{i=1}^n c_i=m\\) and \\(r_i, c_i \\geq 0 \\forall i\\). By stars and bars, the number of ways is \\(\\binom{n+m-1}{m}^{2}\\).\n\nSolution by Shaddoll" ]
USAMO-2015-5
https://artofproblemsolving.com/wiki/index.php/2015_USAMO_Problems/Problem_5
Let \(a, b, c, d, e\) be distinct positive integers such that \(a^4 + b^4 = c^4 + d^4 = e^5\). Show that \(ac + bd\) is a composite number.
[ "Note: This solution is definitely not what the folks at MAA intended, but it works!\n\nLook at the statement \\(a^4+b^4=e^5\\). This can be viewed as a special case of Beal's Conjecture (http://en.wikipedia.org/wiki/Beal%27s_conjecture), stating that the equation \\(A^x+B^y=C^z\\) has no solutions over positive integers for \\(gcd(a, b, c) = 1\\) and \\(x, y, z > 2\\). This special case was proven in 2009 by Michael Bennet, Jordan Ellenberg, and Nathan Ng, as \\((x, y, z) = (2, 4, n)\\). This case \\(a^4+b^4=e^5\\) is obviously contained under that special case, so \\(a\\) and \\(b\\) must have a common factor greater than \\(1\\).\n\nCall the greatest common factor of \\(a\\) and \\(b\\) \\(f\\). Then \\(a = f \\cdot a_1\\) for some \\(a_1\\) and likewise \\(b = f \\cdot b_1\\) for some \\(b_1\\). Then consider the quantity \\(ac+bd\\).\n\n\\(ac+bd = f \\cdot a_1 \\cdot c + f \\cdot b_1 \\cdot d = f\\cdot(a_1 \\cdot c + b_1 \\cdot d)\\).\n\nBecause \\(c\\) and \\(d\\) are both positive, \\((a_1 \\cdot c + b_1 \\cdot d) > 1\\), and by definition \\(f > 1\\), so \\(ac+bd\\) is composite.\n\n~BealsConjecture", "A more conventional approach, using proof by contradiction:\n\nWithout loss of generality, assume \\(a>d\\)\n\nSince \\(a^4 +b^4=c^4+d^4\\), It is obvious that \\(b<c\\)\n\nWe construct the equation \\((a^4 +b^4)c^2d^2-(c^4+d^4)a^2b^2=(a^2c^2-b^2d^2)(a^2d^2-b^2c^2)\\) (the right side is the factorization of the left) Which factors into: \\(e^5(cd-ab)(cd+ab)=(ac-bd)(ac+bd)(ad-bc)(ad+bc)\\) (by using \\(a^4 +b^4=c^4+d^4=e^5\\) and factoring)\n\nIf \\(ac-bd\\) or \\(ad-bc\\) equal zero, then \\(a:b\\) equals \\(c:d\\) or \\(d:c\\) respectively, which is impossible because \\(a^4 +b^4=c^4+d^4\\) and because \\(a,b,c,d,e\\) are distinct. Therefore the right side of the equation above is non-zero and the left side must be divisible by \\(ac+bd\\)\n\nIf \\(ac+bd\\) were prime,\n\nWe have \\(ac+bd-(cd+ab)=(a-d)(c-b)>0\\)\n\nWhich means that \\(ac+bd>cd+ab>cd-ab\\) and neither \\(cd+ab\\) nor \\(cd-ab\\) can be multiples of \\(ac+bd\\) meaning \\(e\\) must be a multiple of \\(ac+bd\\) and \\(e=k(ac+bd)\\) for some integer \\(k\\)\n\nBut clearly this is impossible since \\((k(ac+bd))^5>a^4+b^4\\).\n\nTherefore, by contradiction, \\(ac+bd\\) is composite." ]
USAMO-2015-6
https://artofproblemsolving.com/wiki/index.php/2015_USAMO_Problems/Problem_6
Consider \(0<\lambda<1\), and let \(A\) be a multiset of positive integers. Let \(A_n=\{a\in A: a\leq n\}\). Assume that for every \(n\in\mathbb{N}\), the set \(A_n\) contains at most \(n\lambda\)numbers. Show that there are infinitely many \(n\in\mathbb{N}\) for which the sum of the elements in \(A_n\) is at most \(\frac{n(n+1)}{2}\lambda\). (A multiset is a set-like collection of elements in which order is ignored, but repetition of elements is allowed and multiplicity of elements is significant. For example, multisets \(\{1, 2, 3\}\) and \(\{2, 1, 3\}\) are equivalent, but \(\{1, 1, 2, 3\}\) and \(\{1, 2, 3\}\) differ.)
[ "Proposed by mengmeng142857.\n\nWe prove this by contradiction. Suppose that the number of times that an integer \\(a\\) appears in \\(A\\) is \\(k_a\\). Let the sum of the elements in \\(A_n\\) be \\(S_n=\\sum_{i=1}^{n}i\\cdot k_i\\). If there are only finitely many \\(n\\in\\mathbb{N}\\) such that \\(S_n\\leq\\frac{n(n+1)}{2}\\lambda\\), then by the Well Ordering Principle, there must be a largest \\(m\\in\\mathbb{N}\\) such that \\(S_m\\leq\\frac{m(m+1)}{2}\\lambda\\), and for all \\(n>m\\), \\(S_n>\\frac{n(n+1)}{2}\\lambda\\).\n\nNow, observe that for some \\(n>m\\),\n\n\\[\n\\frac{1}{n}\\cdot S_{n}+\\frac{1}{n(n-1)}\\cdot S_{n-1}+\\frac{1}{(n-1)(n-2)}\\cdot S_{n-2}+...+\\frac{1}{2\\cdot1}\\cdot S_{1} =\\sum_{i=1}^{n}k_i \\leq n\\lambda\n\\]\n\nAlso, \\(\\frac{S_{n}}{n}>\\frac{n+1}{2}\\lambda\\), \\(\\frac{S_{n-1}}{n(n-1)}>\\frac{1}{2}\\lambda\\), \\(\\frac{S_{n-2}}{(n-1)(n-2)}>\\frac{1}{2}\\lambda\\), \\(\\hdots\\) \\(\\frac{S_{m+1}}{(m+2)(m+1)}>\\frac{1}{2}\\lambda\\), \\(\\frac{S_{m}}{(m+1)m}\\leq \\frac{1}{2}\\lambda\\).\n\nNow, for any \\(n\\in\\mathbb{N}\\), we let \\(\\frac{S_n}{(n+1)n}-\\frac{1}{2}\\lambda=d_n\\), then\n\n\\[\n\\frac{1}{n}\\cdot S_n+\\sum_{i=1}^{n-1}\\frac{S_i}{i(i+1)} =n\\lambda+(n+1)\\cdot d_n+\\sum_{i=m+1}^{n-1}d_i+\\sum_{i=1}^{m}d_i \\leq n\\lambda\n\\]\n\nLet \\(\\sum_{i=1}^{m}d_i=-c\\) (where \\(c\\) is positive, otherwise we already have a contradiction), then\n\n\\[\nn\\lambda-c< n\\lambda+(n+1)\\cdot d_n+\\sum_{i=m+1}^{n-1}d_i-c =\\sum_{i=1}^{n}k_i\\leq n\\lambda\n\\]\n\nDividing both sides by \\(n\\), \\(\\lambda-c/n<\\frac{\\sum_{i=1}^{n}k_i}{n}\\leq \\lambda\\). Taking the limit yields \\(\\lim_{n\\to\\infty}{\\frac{\\sum_{i=1}^{n}k_i}{n}}=\\lambda\\). Observe that this would imply that\n\n\\[\n\\lim_{n\\to\\infty}{k_n} =\\lim_{n\\to\\infty}{\\left(\\sum_{i=1}^{n}k_i-\\sum_{i=1}^{n-1}k_i\\right)} =\\lim_{n\\to\\infty}{\\left(n\\cdot\\frac{\\sum_{i=1}^{n}k_i}{n}-(n-1)\\cdot\\frac{\\sum_{i=1}^{n-1}k_i}{n-1}\\right)} =\\lambda\n\\]\n\nHowever, as \\(k_n\\) is an integer, it is impossible for it to converge to \\(\\lambda\\). This yields the desired contradiction." ]
USAMO-2016-1
https://artofproblemsolving.com/wiki/index.php/2016_USAMO_Problems/Problem_1
Let \(X_1, X_2, \ldots, X_{100}\) be a sequence of mutually distinct nonempty subsets of a set \(S\). Any two sets \(X_i\) and \(X_{i+1}\) are disjoint and their union is not the whole set \(S\), that is, \(X_i\cap X_{i+1}=\emptyset\) and \(X_i\cup X_{i+1}\neq S\), for all \(i\in\{1, \ldots, 99\}\). Find the smallest possible number of elements in \(S\).
[ "The answer is that \\(|S| \\ge 8\\).\n\nFirst, we provide a inductive construction for \\(S = \\left\\{ 1, \\dots, 8 \\right\\}\\). Actually, for \\(n \\ge 4\\) we will provide a construction for \\(S = \\left\\{ 1, \\dots, n \\right\\}\\) which has \\(2^{n-1} + 1\\) elements in a line. (This is sufficient, since we then get \\(129\\) for \\(n = 8\\).) The idea is to start with the following construction for \\(|S| = 4\\):\n\n\\[\n\\begin{array}{ccccccccc} 34 & 1 & 23 & 4 & 12 & 3 & 14 & 2 & 13 \\end{array}.\n\\]\n\nThen inductively, we do the following procedure to move from \\(n\\) to \\(n+1\\): take the chain for \\(n\\) elements, delete an element, and make two copies of the chain (which now has even length). Glue the two copies together, joined by \\(\\varnothing\\) in between. Then place the element \\(n+1\\) in alternating positions starting with the first (in particular, this hits \\(n+1\\)). For example, the first iteration of this construction gives:\n\n\\[\n\\begin{array}{ccccccccc} 345 & 1 & 235 & 4 & 125 & 3 & 145 & 2 & 5 \\\\ 34 & 15 & 23 & 45 & 12 & 35 & 14 & 25 & \\end{array}\n\\]\n\nNow let's check \\(|S| \\ge 8\\) is sufficient. Consider a chain on a set of size \\(|S| = 7\\). (We need \\(|S| \\ge 7\\) else \\(2^{|S|} < 100\\).) Observe that there are sets of size \\(\\ge 4\\) can only be neighbored by sets of size \\(\\le 2\\), of which there are \\(\\binom 71 + \\binom 72 = 28\\). So there are \\(\\le 30\\) sets of size \\(\\ge 4\\). Also, there are \\(\\binom 73 = 35\\) sets of size \\(3\\). So the total number of sets in a chain can be at most \\(30 + 28 + 35 = 93 < 100\\).", "My proof that \\(|S|\\ge 8\\) is basically the same as the one above. Here is another construction for \\(|S| = 8\\) that I like because it works with remainders and it's pretty intuitive. The basic idea is to assign different subsets to different remainders when divided by particular numbers, and then to use the Chinese Remainder Theorem to show that all of the subsets are distinct. The motivation for this comes from the fact that we want \\(X_i\\) and \\(X_{i+1}\\) to always be disjoint, so remainders are a great way to systematically make that happen, since \\(i\\) and \\(i+1\\) do not have the same remainder modulo any positive integer greater than \\(1.\\) Anyway, here is the construction:\n\nLet \\(S = \\left\\{ {1, 2, ..., 8} \\right\\}.\\) For \\(i = 1, 2, ..., 98,\\) we will choose which elements of the set \\(\\left\\{ {1, 2, 3, 4} \\right\\}\\) belong to \\(X_i\\) based on the remainder of \\(i\\) modulo \\(9,\\) and we will choose which elements of the set \\(\\left\\{ {5, 6, 7, 8} \\right\\}\\) belong to \\(X_i\\) based on the remainder of \\(i\\) modulo \\(11.\\) We do this as follows:\n\n\\[\nX_i\\cap\\left\\{ {1, 2, 3, 4} \\right\\} = \\begin{array}{ll} \\emptyset & i\\equiv 0 \\text{ (mod } 9\\text{)} \\\\ \\left\\{ {1, 2} \\right\\} & i\\equiv 1 \\text{ (mod } 9\\text{)} \\\\ \\left\\{ {3} \\right\\} & i\\equiv 2 \\text{ (mod } 9\\text{)} \\\\ \\left\\{ {1, 4} \\right\\} & i\\equiv 3 \\text{ (mod } 9\\text{)} \\\\ \\left\\{ {2} \\right\\} & i\\equiv 4 \\text{ (mod } 9\\text{)} \\\\ \\left\\{ {3, 4} \\right\\} & i\\equiv 5 \\text{ (mod } 9\\text{)} \\\\ \\left\\{ {1} \\right\\} & i\\equiv 6 \\text{ (mod } 9\\text{)} \\\\ \\left\\{ {2, 3} \\right\\} & i\\equiv 7 \\text{ (mod } 9\\text{)} \\\\ \\left\\{ {4} \\right\\} & i\\equiv 8 \\text{ (mod } 9\\text{)} \\\\ \\end{array}\n\\]\n\n\\[\nX_i\\cap\\left\\{ {5, 6, 7, 8} \\right\\} = \\begin{array}{ll} \\emptyset & i\\equiv 0 \\text{ (mod } 11\\text{)} \\\\ \\left\\{ {5, 6} \\right\\} & i\\equiv 1 \\text{ (mod } 11\\text{)} \\\\ \\left\\{ {7, 8} \\right\\} & i\\equiv 2 \\text{ (mod } 11\\text{)} \\\\ \\left\\{ {5} \\right\\} & i\\equiv 3 \\text{ (mod } 11\\text{)} \\\\ \\left\\{ {6, 8} \\right\\} & i\\equiv 4 \\text{ (mod } 11\\text{)} \\\\ \\left\\{ {5, 7} \\right\\} & i\\equiv 5 \\text{ (mod } 11\\text{)} \\\\ \\left\\{ {6} \\right\\} & i\\equiv 6 \\text{ (mod } 11\\text{)} \\\\ \\left\\{ {5, 8} \\right\\} & i\\equiv 7 \\text{ (mod } 11\\text{)} \\\\ \\left\\{ {6, 7} \\right\\} & i\\equiv 8 \\text{ (mod } 11\\text{)} \\\\ \\left\\{ {8} \\right\\} & i\\equiv 9 \\text{ (mod } 11\\text{)} \\\\ \\left\\{ {7} \\right\\} & i\\equiv 10 \\text{ (mod } 11\\text{)} \\\\ \\end{array}\n\\]\n\nFinally, we specially define \\(X_{99} = \\left\\{ {1, 2, 3} \\right\\}\\) and \\(X_{100} = \\left\\{ {5, 6, 7} \\right\\}.\\)\n\nIt is relatively easy to see that this configuration satisfies all of the desired conditions. We see that \\(X_{98} = \\left\\{ {4, 7} \\right\\},\\) so \\(X_{98}\\) and \\(X_{99}\\) are disjoint, as are \\(X_{99}\\) and \\(X_{100}.\\) The remainder configuration above takes care of the rest, so any two consecutive sets are disjoint. Then, by the Chinese Remainder Theorem, no two integers from \\(1\\) to \\(98\\) have the same combination of residues modulo \\(9\\) and modulo \\(11,\\) so all of the sets \\(X_i\\) are distinct for \\(i = 1, 2, ..., 98.\\) It is also easy to verify that none of these match \\(X_{99}\\) or \\(X_{100},\\) since they all have at most two elements of \\(\\left\\{ {1, 2, 3, 4} \\right\\}\\) and at most two elements of \\(\\left\\{ {5, 6, 7, 8} \\right\\},\\) whereas \\(X_{99}\\) and \\(X_{100}\\) do not satisfy this; hence all of the sets are distinct. Finally, notice that, for any pair of consecutive sets, at least one of them has at most \\(3\\) elements, while the other has at most \\(4.\\) Thus, their union always has at most \\(7\\) elements, so \\(X_i\\cup X_{i+1}\\neq S\\) for all \\(i = 1, 2, ..., 99.\\)\n\nAll of the conditions are satisfied, so this configuration works. We thus conclude that \\(\\text{min}\\left(\\left|S\\right|\\right) = 8.\\)\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2016-2
https://artofproblemsolving.com/wiki/index.php/2016_USAMO_Problems/Problem_2
Prove that for any positive integer \(k,\) \[ \left(k^2\right)!\cdot\prod_{j=0}^{k-1}\frac{j!}{\left(j+k\right)!} \] is an integer.
[ "Define \\(v_p(N)\\) for all rational numbers \\(N\\) and primes \\(p\\), where if \\(N=\\frac{x}{y}\\), then \\(v_p(N)=v_p(x)-v_p(y)\\), and \\(v_p(x)\\) is the greatest power of \\(p\\) that divides \\(x\\) for integer \\(x\\). Note that the expression(that we're trying to prove is an integer) is clearly rational, call it \\(N\\).\n\n\\(v_p(N)=\\sum_{i=1}^\\infty \\left\\lfloor \\frac{k^{2}}{p^{i}} \\right\\rfloor+\\sum_{j=0}^{k-1} \\sum_{i=1}^\\infty \\left\\lfloor \\frac{j}{p^{i}}\\right\\rfloor-\\sum_{j=k}^{2k-1} \\sum_{i=1}^\\infty \\left\\lfloor \\frac{j}{p^{i}} \\right\\rfloor\\), by Legendre. Clearly, \\(\\left\\lfloor{\\frac{x}{p}}\\right\\rfloor={\\frac{x-r(x,p)}{p}}\\), and \\(\\sum_{i=0}^{k-1} r(i,m)\\leq \\sum_{i=k}^{2k-1} r(i,m)\\), where \\(r(i,m)\\) is the remainder function(we take out groups of \\(m\\) which are just permutations of numbers \\(1\\) to \\(m\\) until there are less than \\(m\\) left, then we have \\(m\\) distinct values, which the minimum sum is attained at \\(0\\) to \\(k-1\\)). Thus, \\(v_p(N)=\\sum_{m=p^{i}, i\\in \\mathbb{N}_{+}}-\\frac{k^{2}}{m}+\\left\\lfloor{\\frac{k^{2}}{m}}\\right\\rfloor-\\frac{\\sum_{i=0}^{k-1} r(i,m)-\\sum_{i=k}^{2k-1} r(i,m)}{m} \\geq \\sum_{m=p^{i}, i\\in \\mathbb{N}} \\left\\lceil -\\frac{k^{2}}{m}+\\lfloor{\\frac{k^{2}}{m}}\\rfloor\\right\\rceil \\geq 0\\), as the term in each summand is a sum of floors also and is clearly an integer.", "Consider an \\(k\\times k\\) grid, which is to be filled with the integers \\(1\\) through \\(k^2\\) such that the numbers in each row are in increasing order from left to right, and such that the numbers in each column are in increasing order from bottom to top. In other words, we are creating an \\(k\\times k\\) standard Young tableaux.\n\nThe Hook Length Formula is the source of the controversy, as it is very powerful and trivializes this problem. The Hook Length Formula states that the number of ways to create this standard Young tableaux (call this \\(N\\) for convenience) is:\n\n\\[\nN = \\frac{\\left(k^2\\right)!}{\\prod_{1\\le i, j\\le k}(i+j-1)}.\n\\]\n\nNow, we do some simple rearrangement:\n\n\\[\nN = \\left(k^2\\right)!\\cdot\\prod_{j=1}^{k}\\prod_{i=1}^{k}\\frac{1}{i+j-1} = \\left(k^2\\right)!\\cdot\\prod_{j=1}^{k}\\frac{\\left(j-1\\right)!}{\\left(j+k-1\\right)!}\n\\]\n\n\\[\n= \\left(k^2\\right)!\\cdot\\prod_{j=0}^{k-1}\\frac{j!}{\\left(j+k\\right)!}.\n\\]\n\nThis is exactly the expression given in the problem! Since the expression given in the problem equals the number of distinct \\(k\\times k\\) standard Young tableaux, it must be an integer, so we are done.", "Let us work in the ring of polynomials \\(\\mathbb{Z}[q]\\). Define\n\n\\[\n[n]_q = 1 + q + \\cdots + q^{n-1}, \\qquad [n]_q! = \\prod_{i=1}^n [i]_q.\n\\]\n\nSince\n\n\\[\n[n]_q! = \\prod_{i=1}^n\\frac{1 - q^i}{1 - q} = \\prod_{d=1}^n \\frac{\\Phi_d(q)^{\\lfloor n/d\\rfloor}}{1 - q},\n\\]\n\nwe have\n\n\\[\n\\nu_d\\bigl([n]_q!\\bigr) = \\lfloor n/d\\rfloor.\n\\]\n\nSet\n\n\\[\nP_k(q) = [k^2]_q!\\,\\prod_{j=0}^{k-1}\\frac{[j]_q!}{[j+k]_q!}.\n\\]\n\nThen\n\n\\[\n\\nu_1\\bigl(P_k(q)\\bigr) = 0\n\\]\n\n\\[\n\\nu_d\\bigl(P_k(q)\\bigr) = \\lfloor k^2/d\\rfloor + \\sum_{j=0}^{k-1}\\lfloor j/d\\rfloor - \\sum_{j=0}^{k-1}\\lfloor (j+k)/d\\rfloor.\n\\]\n\nCombine the last two sums:\n\n\\[\n\\sum_{j=0}^{k-1}\\lfloor j/d\\rfloor - \\sum_{j=0}^{k-1}\\lfloor (j+k)/d\\rfloor = -\\sum_{j=0}^{k-1}\\Bigl(\\lfloor (j+k)/d\\rfloor - \\lfloor j/d\\rfloor\\Bigr).\n\\]\n\nHence\n\n\\[\n\\nu_d\\bigl(P_k(q)\\bigr) = \\Big\\lfloor\\frac{k^2}{d}\\Big\\rfloor - \\sum_{j=0}^{k-1}\\Bigl(\\lfloor\\tfrac{j+k}{d}\\rfloor - \\lfloor\\tfrac{j}{d}\\rfloor\\Bigr).\n\\]\n\nBut for each \\(0\\le j<k\\), the difference \\(\\lfloor\\tfrac{j+k}{d}\\rfloor - \\lfloor\\tfrac{j}{d}\\rfloor\\) counts the multiples of \\(d\\) in the interval \\((j,j+k]\\). As \\(j\\) runs from \\(0\\) to \\(k-1\\), these \\(k\\) intervals lie inside \\(\\{1,2,\\dots,k^2\\}\\), so\n\n\\[\n\\sum_{j=0}^{k-1}\\Bigl(\\lfloor\\tfrac{j+k}{d}\\rfloor - \\lfloor\\tfrac{j}{d}\\rfloor\\Bigr) \\;\\le\\; \\Big\\lfloor\\frac{k^2}{d}\\Big\\rfloor.\n\\]\n\nIt follows that \\(\\nu_d(P_k(q))\\ge0\\) , hence \\(P_k(q)\\in\\mathbb{Z}[q]\\). Evaluating at \\(q=1\\) completes the proof.\n\n~Lopkiloinm\n\nNote: This solution is fundamentally different from the first, which works purely in the integers. Here, we work in the integer polynomial ring \\(\\mathbb{Z}[q]\\)—a graded algebra—which allows us to test integrality by tracking the multiplicities of better organized elements like \\(\\Phi_d(q)\\) for all integers \\(1 \\le d \\le k^2\\). In contrast, working in \\(\\mathbb{Z}\\)—a non-graded algebra—requires using \\(p\\)-adic valuations via Legendre's formula, which only considers primes which are less organized. The graded structure of \\(\\mathbb{Z}[q]\\) simplifies the analysis, making it a powerful strategy to lift problems into a graded algebra whenever possible." ]
USAMO-2016-3
https://artofproblemsolving.com/wiki/index.php/2016_USAMO_Problems/Problem_3
Let \(\triangle ABC\) be an acute triangle, and let \(I_B, I_C,\) and \(O\) denote its \(B\)-excenter, \(C\)-excenter, and circumcenter, respectively. Points \(E\) and \(Y\) are selected on \(\overline{AC}\) such that \(\angle ABY = \angle CBY\) and \(\overline{BE}\perp\overline{AC}.\) Similarly, points \(F\) and \(Z\) are selected on \(\overline{AB}\) such that \(\angle ACZ = \angle BCZ\) and \(\overline{CF}\perp\overline{AB}.\) Lines \(I_B F\) and \(I_C E\) meet at \(P.\) Prove that \(\overline{PO}\) and \(\overline{YZ}\) are perpendicular.
[ "This problem can be proved in the following two steps.\n\n1. Let \\(I_A\\) be the \\(A\\)-excenter, then \\(I_A,O,\\) and \\(P\\) are colinear. This can be proved by the Trigonometric Form of Ceva's Theorem for \\(\\triangle I_AI_BI_C.\\)\n\n2. Show that \\(I_AY^2-I_AZ^2=OY^2-OZ^2,\\) which implies \\(\\overline{OI_A}\\perp\\overline{YZ}.\\) This can be proved by multiple applications of the Pythagorean Thm.", "We find point \\(T\\) on line \\(YZ,\\) we prove that \\(TY \\perp OI_A\\) and state that \\(P\\) is the point \\(X(24)\\) from ENCYCLOPEDIA OF TRIANGLE, therefore \\(P \\in OI_A.\\)\n\nLet \\(\\omega\\) be circumcircle of \\(\\triangle ABC\\) centered at \\(O.\\) Let \\(Y_1,\\) and \\(Z_1\\) be crosspoints of \\(\\omega\\) and \\(BY,\\) and \\(CZ,\\) respectively. Let \\(T\\) be crosspoint of \\(YZ\\) and \\(Y_1 Z_1.\\) In accordance the Pascal theorem for pentagon \\(AZ_1BCY_1,\\) \\(AT\\) is tangent to \\(\\omega\\) at \\(A.\\)\n\nLet \\(I_A, I_B, I_C\\) be \\(A, B,\\) and \\(C\\)-excenters of \\(\\triangle ABC.\\) Denote\n\n\\[\na = BC, b = AC, c = AB, 2\\alpha = \\angle CAB, 2\\beta = \\angle ABC, 2\\gamma = \\angle ACB,\n\\]\n\n\\[\n\\psi = 90^\\circ – \\gamma + \\beta, X = AI_A \\cap \\omega, X_1 = BC \\cap AI_A,\n\\]\n\n\\[\nI = BI_B \\cap CI_C, U= YZ \\cap AI_A, W = Y_1Z_1 \\cap AI_A,\n\\]\n\n\\(V\\) is the foot ot perpendicular from \\(O\\) to \\(AI_A.\\)\n\n\\(I\\) is ortocenter of \\(\\triangle I_A I_B I_C\\) and incenter of \\(\\triangle ABC.\\)\n\n\\(\\omega\\) is the Nine–point circle of \\(\\triangle I_A I_B I_C.\\)\n\n\\(Y_1\\) is the midpoint of \\(II_B, Z_1\\) is the midpoint of \\(II_C\\) in accordance with property of Nine–point circle \\(\\implies\\)\n\n\\[\nY_1Z_1 || I_B AI_C || VO, IW = AW \\implies TW \\perp AI.\n\\]\n\n\\[\n\\angle AXC = 180 ^\\circ – 2\\gamma – \\alpha = 90 ^\\circ – \\gamma + \\beta = \\psi.\n\\]\n\n\\[\n\\angle TAI = \\angle VOA = 2\\beta + \\alpha = 90 ^\\circ – \\gamma + \\beta = \\psi.\n\\]\n\n\\[\nI_A X_1 = IX_1 = BX_1 = 2R \\sin \\alpha \\implies\n\\]\n\n\\[\n\\cot \\angle OI_A A = \\frac {VI_A}{VO} = \\frac {R \\sin \\psi + 2R \\sin \\alpha}{R \\cos \\psi} = \\tan \\psi + \\frac{2 \\sin\\alpha}{\\cos \\psi}.\n\\]\n\n\\[\nCX = \\frac {ab}{b+c} \\implies \\frac {AI}{IX}= \\frac {AC}{CX}= \\frac {b+c}{a} \\implies AI = AX \\frac {b+c}{a+b+c},\n\\]\n\n\\[\nAW = \\frac {AI}{2}, UW = AU – AW,\n\\]\n\nIn \\(\\triangle ABC\\) segment \\(YZ\\) cross segment \\(AX \\implies \\frac {AU}{UX} = \\frac {m + nk}{k+1},\\) where \\(n = \\frac {a}{b}, m = \\frac{a}{c}, k=\\frac {b}{c},\\)\n\n\\[\n\\frac {AU}{UX} = \\frac{2a}{b+c} \\implies AU = AX \\cdot \\frac {b+c}{2a +b +c}.\n\\]\n\n\\[\n\\frac {AU – AW}{AW} = \\frac {b+c} {2a + b + c}.\n\\]\n\n\\[\n\\cot \\angle UTW = \\frac {TW}{UW} = \\frac {AW \\cdot \\tan \\psi}{AU – AW} = \\tan \\psi \\cdot \\frac {2a +b+c}{b+c} =\n\\]\n\n\\[\n=\\tan \\psi \\cdot \\frac {2a}{b+c} + \\tan \\psi = \\frac {2 \\sin \\alpha}{\\sin \\psi} \\tan \\psi + \\tan \\psi = \\tan \\psi + \\frac{2 \\sin\\alpha}{\\cos \\psi}\n\\]\n\n\\[\n\\implies \\angle UTW = \\angle OI_AA .\n\\]\n\n\\[\nTW \\perp AI_A \\implies TYZ \\perp OI_A.\n\\]\n\nLet \\(\\triangle II_B I_C\\) be the base triangle with orthocenter \\(I_A,\\) center of Nine-points circle \\(O \\implies OI_A\\) be the Euler line of \\(\\triangle II_B I_C.\\)\n\n\\(\\triangle ABC\\) is orthic triangle of \\(\\triangle II_B I_C,\\)\n\n\\(\\triangle DEF\\) is orthic-of-orthic triangle.\n\n\\(P\\) is perspector of base triangle and orthic-of-orthic triangle.\n\nTherefore \\(P\\) is point \\(X(24)\\) of ENCYCLOPEDIA OF TRIANGLE CENTERS which lies on Euler line of the base triangle. [[1]]\n\nClaim\n\n\\[\n\\frac {AU}{UX} = \\frac {m + nk}{k + 1}.\n\\]\n\nProof\n\n\\[\n\\frac {[AYZ]}{[ABC]} = \\frac {AZ \\cdot AY}{AB \\cdot AC} = \\frac {1}{(n + 1) \\cdot (m+1)},\n\\]\n\n\\[\n\\frac {[BXZ]}{[ABC]} = \\frac {BZ \\cdot BX}{AB \\cdot BC} = \\frac {n}{(n + 1) \\cdot (k+1)},\n\\]\n\n\\[\n\\frac {[CXY]}{[ABC]} = \\frac {CY \\cdot CX}{AC \\cdot BC} = \\frac {mk}{(m + 1) \\cdot (k+1)},\n\\]\n\n\\[\n\\frac {[XYZ]}{[ABC]} = 1 - \\frac {[AYZ]}{[ABC]} – \\frac {[BXZ]}{[ABC]} – \\frac {[CXY]}{[ABC]} = \\frac {m+nk}{(m + 1) \\cdot (k+1)\\cdot (n+1)},\n\\]\n\n\\[\n\\frac {AU}{UX} = \\frac {[AYZ]}{[XYZ]} = \\frac {m + nk}{k + 1}.\n\\]\n\nvladimir.shelomovskii@gmail.com, vvsss" ]
USAMO-2016-4
https://artofproblemsolving.com/wiki/index.php/2016_USAMO_Problems/Problem_4
Find all functions \(f:\mathbb{R}\rightarrow \mathbb{R}\) such that for all real numbers \(x\) and \(y\), \[ (f(x)+xy)\cdot f(x-3y)+(f(y)+xy)\cdot f(3x-y)=(f(x+y))^2. \]
[ "Step 1: Set \\(x = y = 0\\) to obtain \\(f(0) = 0.\\)\n\nStep 2: Set \\(x = 0\\) to obtain \\(f(y)f(-y) = f(y)^2.\\)\n\n\\(\\indent\\) In particular, if \\(f(y) \\ne 0\\) then \\(f(y) = f(-y).\\)\n\n\\(\\indent\\) In addition, replacing \\(y \\to -t\\), it follows that \\(f(t) = 0 \\implies f(-t) = 0\\) for all \\(t \\in \\mathbb{R}.\\)\n\nStep 3: Set \\(x = 3y\\) to obtain \\(\\left[f(y) + 3y^2\\right]f(8y) = f(4y)^2.\\)\n\n\\(\\indent\\) In particular, replacing \\(y \\to t/8\\), it follows that \\(f(t) = 0 \\implies f(t/2) = 0\\) for all \\(t \\in \\mathbb{R}.\\)\n\nStep 4: Set \\(y = -x\\) to obtain \\(f(4x)\\left[f(x) + f(-x) - 2x^2\\right] = 0.\\)\n\n\\(\\indent\\) In particular, if \\(f(x) \\ne 0\\), then \\(f(4x) \\ne 0\\) by the observation from Step 3, because \\(f(4x) = 0 \\implies f(2x) = 0 \\implies f(x) = 0.\\) Hence, the above equation implies that \\(2x^2 = f(x) + f(-x) = 2f(x)\\), where the last step follows from the first observation from Step 2.\n\n\\(\\indent\\) Therefore, either \\(f(x) = 0\\) or \\(f(x) = x^2\\) for each \\(x.\\)\n\n\\(\\indent\\) Looking back on the equation from Step 3, it follows that \\(f(y) + 3y^2 \\ne 0\\) for any nonzero \\(y.\\) Therefore, replacing \\(y \\to t/4\\) in this equation, it follows that \\(f(t) = 0 \\implies f(2t) = 0.\\)\n\nStep 5: If \\(f(a) = f(b) = 0\\), then \\(f(b - a) = 0.\\)\n\n\\(\\indent\\) This follows by choosing \\(x, y\\) such that \\(x - 3y = a\\) and \\(3x - y = b.\\) Then \\(x + y = \\tfrac{b - a}{2}\\), so plugging \\(x, y\\) into the given equation, we deduce that \\(f\\left(\\tfrac{b - a}{2}\\right) = 0.\\) Therefore, by the third observation from Step 4, we obtain \\(f(b - a) = 0\\), as desired.\n\nStep 6: If \\(f \\not\\equiv 0\\), then \\(f(t) = 0 \\implies t = 0.\\)\n\n\\(\\indent\\) Suppose by way of contradiction that there exists an nonzero \\(t\\) with \\(f(t) = 0.\\) Choose \\(x, y\\) such that \\(f(x) \\ne 0\\) and \\(x + y = t.\\) The following three facts are crucial:\n\n\\(\\indent\\) 1. \\(f(y) \\ne 0.\\) This is because \\((x + y) - y = x\\), so by Step 5, \\(f(y) = 0 \\implies f(x) = 0\\), impossible.\n\n\\(\\indent\\) 2. \\(f(x - 3y) \\ne 0.\\) This is because \\((x + y) - (x - 3y) = 4y\\), so by Step 5 and the observation from Step 3, \\(f(x - 3y) = 0 \\implies f(4y) = 0 \\implies f(2y) = 0 \\implies f(y) = 0\\), impossible.\n\n\\(\\indent\\) 3. \\(f(3x - y) \\ne 0.\\) This is because by the second observation from Step 2, \\(f(3x - y) = 0 \\implies f(y - 3x) = 0.\\) Then because \\((x + y) - (y - 3x) = 4x\\), Step 5 together with the observation from Step 3 yield \\(f(3x - y) = 0 \\implies f(4x) = 0 \\implies f(2x) = 0 \\implies f(x) = 0\\), impossible.\n\n\\(\\indent\\) By the second observation from Step 4, these three facts imply that \\(f(y) = y^2\\) and \\(f(x - 3y) = \\left(x - 3y\\right)^2\\) and \\(f(3x - y) = \\left(3x - y\\right)^2.\\) By plugging into the given equation, it follows that\n\n\\[\n\\begin{align*} \\left(x^2 + xy\\right)\\left(x - 3y\\right)^2 + \\left(y^2 + xy\\right)\\left(3x - y\\right)^2 = 0. \\end{align*}\n\\]\n\nBut the above expression miraculously factors into \\(\\left(x + y\\right)^4\\)! This is clearly a contradiction, since \\(t = x + y \\ne 0\\) by assumption. This completes Step 6.\n\nStep 7: By Step 6 and the second observation from Step 4, the only possible solutions are \\(f \\equiv 0\\) and \\(f(x) = x^2\\) for all \\(x \\in \\mathbb{R}.\\) It's easy to check that both of these work, so we're done.", "From steps 1 and 2 of Solution 1 we have that \\(f(0)=0\\), and \\(f(-y) \\cdot f(y)=(f(y))^2\\). Therefore, if \\(f(y) \\ne 0\\), then \\(f(y)=f(-y)\\). Furthermore, setting \\(y=-x\\) gives us \\((f(x)-x^2) \\cdot f(4x) + (f(-x)-x^2) \\cdot f(4x) = f(0) = 0\\). The LHS can be factored as \\((f(x)+f(-x)-2x^2) \\cdot f(4x) = 0\\). In particular, if \\(f(4x) \\ne 0\\), then we have \\(f(x)+f(-x)=2x^2\\). However, since we have from step 2 that \\(f(x)=f(-x)\\), assuming \\(f(x) \\ne 0\\), the equation becomes \\(2f(x)=2x^2\\), so for every \\(x\\), \\(f(x)\\) is equivalent to either \\(0\\) or \\(x^2\\). From step 6 of Solution 1, we can prove that \\(f(x)=0\\), and \\(f(x)=x^2\\) are the only possible solutions.", "Step 1: \\(x=y=0 \\implies f(0)=0\\)\n\nStep 2: \\(x=0 \\implies f(y)f(-y)=f(y)^{2}\\). Now, assume \\(y \\not = 0\\). Then, if \\(f(y)=0\\), we substitute in \\(-y\\) to get \\(f(y)f(-y)=f(-y)^{2}\\), or \\(f(y)=f(-y)=0\\). Otherwise, we divide both sides by \\(f(y)\\) to get \\(f(y)=f(-y)\\). If \\(y=0\\), we obviously have \\(f(0)=f(0)\\). Thus, the function is even. . Step 3: \\(y=-x \\implies 2f(4x)(f(x)-x^{2})=0\\). Thus, \\(\\forall x\\), we have \\(f(4x)=0\\) or \\(f(x)=x^{2}\\).\n\nStep 4: We now assume \\(f(x) \\not = 0\\), \\(x\\not = 0\\). We have \\(f(\\frac{x}{4})=\\frac{x^{2}}{16}\\). Now, setting \\(x=y=\\frac{x}{4}\\), we have \\(f(\\frac{x}{2})=\\frac{x^{2}}{4}\\) or \\(f(\\frac{x}{2})=0\\). The former implies that \\(f(x)=0\\) or \\(x^{2}\\). The latter implies that \\(f(x)=0\\) or \\(f(x)=\\frac{x^{2}}{2}\\). Assume the latter. \\(y=-2x \\implies -\\frac{3y^{2}}{2}f(7y)-{2y^{2}}f(5y)=\\frac{x^{2}}{2}\\). Clearly, this implies that \\(f(x)\\) is negative for some \\(m\\). Now, we have \\(f(\\frac{m}{4})=\\frac{m^{2}}{16} \\implies f(\\frac{m}{2})=0,\\frac{m^{2}}{4} \\implies f(m) \\geq 0\\), which is a contradiction. Thus, \\(\\forall x\\)\\(f(x)=0\\) or \\(f(x)=x^{2}\\).\n\nStep 5: We now assume \\(f(x)=0\\), \\(f(y)=y^{2}\\) for some \\(x,y \\not = 0\\). Let \\(m\\) be sufficiently large integer, let \\(z=|4^{m}x|\\) and take the absolute value of \\(y\\)(since the function is even). Choose \\(c\\) such that \\(3z-c=y\\). Note that we have \\(\\frac{c}{z}\\)~\\(3\\) and \\(\\frac{y}{z}\\)~\\(0\\). Note that \\(f(z)=0\\). Now, \\(x=z, y=c \\implies\\) LHS is positive, as the second term is positive and the first term is nonnegative and thus the right term is equal to \\((z+c)^{4}\\)~\\(256z^{4}\\). Now if \\(f(z-3c)=0\\), the second term of the LHS/RHS clearly ~0 as \\(m \\to \\infty\\). if \\(f(z-3c)=0\\), then we have LHS/RHS ~ \\(0\\), otherwise, we have LHS/RHS~\\(\\frac{8^{2}\\cdot 3z^{4}}{256z^{4}}\\)~\\(\\frac{3}{4}\\), a contradiction, as we're clearly not dividing by \\(0\\), and we should have LHS/RHS=1.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2016-5
https://artofproblemsolving.com/wiki/index.php/2016_USAMO_Problems/Problem_5
An equilateral pentagon \(AMNPQ\) is inscribed in triangle \(ABC\) such that \(M\in\overline{AB},\) \(Q\in\overline{AC},\) and \(N, P\in\overline{BC}.\) Let \(S\) be the intersection of lines \(MN\) and \(PQ.\) Denote by \(\ell\) the angle bisector of \(\angle MSQ.\) Prove that \(\overline{OI}\) is parallel to \(\ell,\) where \(O\) is the circumcenter of triangle \(ABC,\) and \(I\) is the incenter of triangle \(ABC.\)
[ "Let \\(D\\) be the intersection of line \\(AI\\) and the circumcircle of \\(\\Delta ABC\\) (other than \\(A\\)), then \\(OD\\perp BC\\). Let \\(R\\) be the point such that \\(NPQR\\) is a rhombus. It follows that \\(OD\\perp QR\\).\n\nSince \\(AM=AQ\\), \\(AI\\perp MQ\\), or \\(DI\\perp MQ\\). It follows that \\(\\angle ODI=\\angle RQM\\).\n\nSince \\(BO=OD\\), \\(MA=AQ\\), \\(\\angle BOD=\\angle MAQ\\), it follows that \\(\\Delta BOD\\sim\\Delta MAQ\\), so \\(AQ/MQ=OD/BD\\).\n\nIt is given that \\(AQ=NP=RQ\\), and by basic properties of the incenter, \\(ID=BD\\). Therefore, \\(RQ/MQ=OD/ID\\), so \\(\\Delta RQM\\sim\\Delta ODI\\). Since the rotation between the two triangles in 90 degrees, \\(OI\\perp MR\\). However, \\(l\\) is parallel to the bisector of \\(MNR\\), which is perpendicular to \\(MR\\), so we are done.", "Write \\(\\angle{JKL} = K\\) for all \\(J,K,L\\) chosen as distinct vertices of triangle \\(ABC\\). Define \\(a, b, c\\) as sides opposite to angles \\(A, B\\), and \\(C\\), respectively. Place the triangle in the Euclidean plane with \\(A\\) at the origin and \\(C\\) on the positive x-axis. Assume without loss of generality that C is acute.\n\nConsider the sides of the pentagon as vectors and note that\n\n\\[\n\\overrightarrow{AM} + \\overrightarrow{MN} + \\overrightarrow{NP} = \\overrightarrow{AQ} + \\overrightarrow{QP} \\qquad (1)\n\\]\n\nDefine \\(\\delta\\) and \\(\\gamma\\) as the angles made between the positive x-axis and \\(\\overrightarrow{MN}\\) and \\(\\overrightarrow{QP}\\), respectively. Considering the x and y coordinates of the vectors in \\((1)\\), it follows that\n\n\\[\n\\cos \\delta - \\cos \\gamma = 1 - \\cos A - \\cos C \\qquad (2)\n\\]\n\n\\[\n\\sin \\delta - \\sin \\gamma = \\sin C - \\sin A \\qquad (3)\n\\]\n\nSuppose \\(\\sin C - \\sin A = 0\\). Then \\(A = C\\), and the triangle is isosceles. In this case, it is clear by symmetry that \\(\\overline{OI}\\) is vertical. Further, since point \\(S\\) exists, \\(\\delta \\neq \\gamma\\), so \\(\\delta + \\gamma = 180\\) and \\(\\overrightarrow{MN} + \\overrightarrow{QP}\\) must be vertical as well.\n\nFor the remainder of the proof, assume \\(\\sin C \\neq \\sin A\\). Note that\n\n\\[\n\\frac{\\cos y – \\cos x}{\\sin x - \\sin y} = \\frac{ (\\cos y - \\cos x)(\\sin x + \\sin y)}{ (\\sin x - \\sin y)(\\sin x + \\sin y)} = \\frac {\\sin x + \\sin y}{\\cos x + \\cos y}\n\\]\n\nwhenever \\(x, y \\in \\mathbb{R}\\) and \\(\\sin x \\neq \\sin y\\). Note further that the slope of the line defined by the vector formed by summing vectors \\((\\cos x, \\sin x)\\) and \\((\\cos y, \\sin y)\\) is this expression. Since \\(\\ell\\) is parallel to \\(\\overrightarrow{MN} + \\overrightarrow{QP}\\), the slope of \\(\\ell\\) can be formed by dividing expressions in \\((2)\\) and \\((3)\\) and inverting the sign:\n\n\\[\n\\frac{\\cos A + \\cos C – 1}{\\sin C - \\sin A} \\qquad (4)\n\\]\n\nDetermine the coordinates of \\(I\\) by drawing perpendiculars from \\(I\\) to the sides and vertices of the triangle. By exploiting congruence between pairs of right triangles that share a vertex, one can partition \\(a,b,c\\) into \\(p+q, p+r, q+r\\) where \\(p,q,r\\) are bases of these triangles that lie on the sides of triangle \\(ABC\\). From here it is clear that \\(I = \\left(\\frac{b + c – a}{2} , \\frac{b + c – a}{2}\\tan(A / 2)\\right)\\).\n\nTo find the coordinates of \\(O\\), note that \\(\\angle{OJK} + \\angle{OJL} = \\angle{J}\\) and that \\(\\angle {OJK} = \\angle{OKJ}\\) in any acute triangle \\(JKL\\). It easily follows that \\(\\angle{OJK} = 90 – L\\). Note also that the perpendicular from \\(O\\) to \\(\\overline{JK}\\) bisects \\(\\overline{JK}\\). Hence,\n\n\\[\nO = \\left(\\frac{b}{2},\\frac{b}{2} \\cot B\\right) \\qquad (5)\n\\]\n\nif triangle \\(ABC\\) is acute.\n\nIf triangle \\(JKL\\) is obtuse at \\(\\angle J\\), then it can be similarly shown that \\(\\angle{OKL} = \\angle{OLK} = J – 90\\) but that the remaining angles of this form are still \\(90-L\\) and \\(90-K\\). It easily follows that \\((5)\\) holds if \\(\\angle A\\) is obtuse. If \\(\\angle B\\) is obtuse then \\(\\angle OAC = B – 90\\) and the \\(y\\) coordinate of \\(O\\) is \\(-\\frac{b}{2} \\tan{B-90}\\). From this, \\((5)\\) follows in this case as well.\n\nWe can conclude the slope of \\(\\overline {OI}\\) is\n\n\\[\n\\left(\\frac{b \\cot {B} – (b + c – a) \\tan(A/2)}{a – c}\\right) = \\left( \\frac {\\cos B – \\tan{(A/2)}\\sin B}{\\sin A - \\sin C}\\right) + \\tan(A/2) \\quad (6)\n\\]\n\nby the Law of Sines and rearrangement.\n\nSetting \\((6) = (4)\\) is equivalent to\n\n\\[\n1 - \\cos A - \\cos C = \\cos B + \\tan(A/2)(\\sin A - \\sin B - \\sin C)\n\\]\n\nSince \\(\\tan(A/2) = \\frac{\\sin A}{1 + \\cos A}\\), this equation is equivalent to\n\n\\[\n(1 + \\cos A)(\\cos A + \\cos B + \\cos C – 1) = (\\sin A)(\\sin B + \\sin C - \\sin A)\n\\]\n\nThis equation is equivalent to\n\n\\[\n\\cos(A + B) + \\cos(A + C) + \\cos B + \\cos C = 0\n\\]\n\nwhich is evident.", "Let \\(A', B',\\) and \\(C'\\) be the arc midpoints of \\(BC, CA, AB,\\) respectively. Let \\(E\\) be crosspoint of \\(AI\\) and \\(B'C'.\\)\n\nTherefore \\(O\\) is the circumcenter of triangle \\(A'B'C'.\\)\n\nPoints \\(A, I, E,\\) and \\(A'\\) are collinear. \\(\\angle A'EB' = \\frac {\\overset{\\Large\\frown} {AC'}+\\overset{\\Large\\frown} {B'C} +\\overset{\\Large\\frown} {CA'}}{2} = 90^\\circ \\implies AA' \\perp B'C'\\) \\(\\implies I\\) is orthocenter of \\(\\triangle A'B'C' \\implies\\)\n\n\\(IO\\) is the Euler line of \\(\\triangle A'B'C'.\\)\n\nLet \\(G\\) be the centroid of \\(\\triangle A'B'C' \\implies G\\) lies on line \\(IO\\) \\(\\implies \\overline {OG} = \\frac {\\overline{OA'} + \\overline{OB'} + \\overline{OC'}}{3}\\) is paraller to \\(\\overline{OI}.\\) \\(\\overline{OB'} \\perp \\overline{AC} \\implies \\overline{OB'} \\perp \\overline{QA}.\\)\n\nSimilarly \\(\\overline{OC'} \\perp \\overline{AM}, \\overline{OA'} \\perp \\overline{NP},\\) rotation from \\(\\overline{OA'}\\) to \\(\\overline{NP}\\), from \\(\\overline{OB'}\\) to \\(\\overline{QA},\\) and from \\(\\overline{OC'}\\) to \\(\\overline{AM}\\) is in clockwise direction, \\(|\\overline{QA}|=| \\overline{AM}| =|\\overline{NP}|, |\\overline{OA'}| = |\\overline{OB'}| = |\\overline{OC'}| \\implies\\)\n\n\\(\\overline{QA} + \\overline{AM} + \\overline{NP} = \\overline{QM} + \\overline{NP}\\) is perpendicular to \\(|\\overline{OI}.\\)\n\n\\(|\\overline{MN}| = |\\overline{QP}|\\) therefore in accordance with Claim \\(\\overline{MN} + \\overline{QP}\\) is parallel to \\(\\overline{OI}.\\)\n\nThis sum is parallel to \\(\\ell,\\), so we are done.\n\nClaim\n\nLet \\(|\\overline{BA}|= |\\overline{CD}|, \\overline{CB} + |\\overline{BA} + |\\overline{AD} = |\\overline{CD}.\\) Then \\((\\overline{BA} + \\overline{CD}) \\perp (\\overline{CB} + \\overline{AD}).\\)\n\nProof\n\n\\[\n(\\overline{BA} + \\overline{CD}) \\cdot (\\overline{CB} + \\overline{AD}) = (\\overline{BA} + \\overline{CD}) \\cdot (\\overline{CB} + \\overline{CD} – \\overline{CB} – \\overline{BA}) = |\\overline{CD}|^2 – |\\overline{BA}|^2 = 0.\n\\]\n\nvladimir.shelomovskii@gmail.com, vvsss" ]
USAMO-2016-6
https://artofproblemsolving.com/wiki/index.php/2016_USAMO_Problems/Problem_6
Integers \(n\) and \(k\) are given, with \(n\ge k\ge 2.\) You play the following game against an evil wizard. The wizard has \(2n\) cards; for each \(i = 1, ..., n,\) there are two cards labeled \(i.\) Initially, the wizard places all cards face down in a row, in unknown order. You may repeatedly make moves of the following form: you point to any \(k\) of the cards. The wizard then turns those cards face up. If any two of the cards match, the game is over and you win. Otherwise, you must look away, while the wizard arbitrarily permutes the \(k\) chosen cards and turns them back face-down. Then, it is your turn again. We say this game is \(\textit{winnable}\) if there exist some positive integer \(m\) and some strategy that is guaranteed to win in at most \(m\) moves, no matter how the wizard responds. For which values of \(n\) and \(k\) is the game winnable?
[ "## Case I:\n\nWe first prove that the game is winnable whenever \\(n > k\\) by demonstrating a winning strategy in this case.\n\nOn the \\(i\\)th move, choose the \\(k\\) cards in positions \\(i\\) through \\(i+k-1.\\) Assuming that you do not win on any earlier move, repeat this for \\(1\\le i \\le 2n-k+1.\\)\n\nAssume that you did not win on any of the first \\(2n-k+1\\) moves, as described above. Let \\(j\\) be an integer such that \\(1\\le j\\le 2n-k.\\) On the \\(j\\)th move, the wizard revealed the cards in positions \\(j\\) through \\(j+k-1,\\) so you know the labels of all of these cards (just not necessarily in the right order). Then, on the \\((j+1)\\)th move, the wizard revealed the cards in positions \\(j+1\\) through \\(j+k,\\) which means that you get to see all of the cards that were moved to positions \\(j+1\\) through \\(j+k.\\) This means that you can uniquely determine the label on card \\(j,\\) since you knew all of the labels from \\(j\\) through \\(j+k-1,\\) and the card in position \\(j\\) could not have moved anywhere else since your last move.\n\nIt follows that, after the sequence of \\(2n-k+1\\) moves described above, you know the labels on the first \\(2n-k\\) cards. Since \\(n > k,\\) we have \\(2n-k \\ge n+1,\\) so there must be a pair of cards with matching labels in this group of \\(2n-k\\) cards, by the Pigeonhole Principle. On your next move, you can pick a group of \\(k\\) cards that includes that pair of matching cards, and you win.\n\nWe have created a strategy that is guaranteed to win in at most \\(m = 2n-k+2\\) moves. Thus, the game is winnable for all \\(n > k.\\)\n\n## Case II:\n\nWe now prove that the game is not winnable if \\(n=k.\\) We will say that the game is in a state \\(S\\) if your knowledge about the card labels is of the following form:\n\nThere exists a group of \\(n\\) cards for which you know that those \\(n\\) cards have all of the labels \\(1, 2, ..., n\\) (i.e. you know that they have all distinct labels) in some order, but you know nothing about which of those \\(n\\) cards have which labels. (Call this group of cards Group \\(A.\\))\n\nSuppose that the game is in such a state \\(S.\\) We will now show that, regardless of your next move, you cannot guarantee victory or an escape from state \\(S.\\)\n\nClearly, the \\(n\\) cards that are not in Group \\(A\\) must also have all of the labels \\(1, 2, ..., n.\\) (You might know something about which cards have which labels, or you might not.) Call this other collection of cards Group \\(B.\\)\n\nIf, on the next move, you pick all of the cards from Group \\(A\\) or all of the cards from Group \\(B,\\) then you clearly will not get a matching pair. The wizard will then arbitrarily permute those cards. Thus, for those \\(n\\) chosen cards, you know their labels are all distinct, but you know nothing about which cards have which labels. Thus, you are back in state \\(S.\\)\n\nNow, suppose you pick \\(x\\) cards from Group \\(A\\) and \\(n-x\\) cards from Group \\(B,\\) where \\(x\\) is an integer and \\(1\\le x\\le n-1.\\) Then, the cards chosen from Group \\(B\\) will form a set of labels \\(P\\subset Z_n,\\) where \\(Z_n = \\left\\{ {1, 2, ..., n} \\right\\}\\) and \\(|P| = n-x.\\) However, you know nothing about which cards in Group \\(A\\) have which labels. Thus, there is no way for you to prevent the \\(x\\) cards from Group \\(A\\) to form the exact set of labels \\(Q = Z_n\\setminus P.\\) In such a case, there will be no matching cards, so you will not win. Furthermore, the wizard will then arbitrarily permute these \\(n\\) cards, so you will know that they have all \\(n\\) distinct labels, but you will know nothing about which cards have which labels. Therefore, you are again in state \\(S.\\)\n\nWe have covered all cases, so it follows that, once you enter state \\(S,\\) you cannot guarantee escape from state \\(S\\) or victory.\n\nNow, look at the very first move you make. Obviously, you cannot guarantee victory on the first move, as you know nothing about which cards have which labels. Assuming that you do not win on the first move, the \\(n\\) cards you chose have all distinct labels. The wizard then permutes the \\(n\\) cards you chose, so you now know that those \\(n\\) cards have all distinct labels but know nothing about which cards have which labels. Therefore, if you do not win on your first move, then the game enters state \\(S,\\) and we have already proven that you cannot guarantee victory from this point.\n\nWe therefore conclude that the game is not winnable if \\(n=k.\\) We proved earlier that the game is winnable if \\(n>k,\\) so the game is winnable if and only if \\(n>k\\ge 2.\\)", "We claim that the game is winnable if and only if \\(n>k\\). Suppose after the first step, the cards \\(1\\) to \\(n\\) are shuffled around. Notice that we have \\(n\\) cards that we don't know the position of (which are all cards from \\(1\\) to \\(n\\)). Now, suppose we pick \\(p\\) known cards. Note that the \\(p\\) cards are all different(since the known cards are the cards from \\(1\\) to \\(n\\)), and there is still a possibility that the other cards from the unknown cards complement and cause \\(1\\) to \\(n\\). Therefore, we are in the same state as before, and the game is unwinnable.\n\nNow, suppose \\(n>k\\). Denote the ith card counting from the left. We pick cards \\(1\\) to \\(k\\), keeping track of the set of values of the cards. Then, we pick cards \\(2\\) to \\(k+1\\), adding the value of the \\(k+1\\)th card into the set of value of cards. We keep doing this, until we pick cards \\(2n-k\\) to \\(2n-1\\), at which point we know the exact number on the \\(2n\\)th card. Now, we go back to \\(1\\) through \\(k\\), and repeat this process, until we reveal the \\(2n-1\\)th card(unless we win during the process). This process terminates only when there are less or equal to \\(k\\) cards that we don't know the exact numbers on or if we somehow win, clearly, as otherwise we're still revealing new information by picking cards from \\(1\\) through \\(k\\). Note that we now know the exact values on \\(2n-k\\) of the cards. By the Pigeonhole Principle, since \\(k<n\\), \\(2\\) of them are the same, and we pick those \\(2\\) cards and \\(k-2\\) other random cards and we win.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2017-1
https://artofproblemsolving.com/wiki/index.php/2017_USAMO_Problems/Problem_1
Prove that there are infinitely many distinct pairs \((a,b)\) of relatively prime positive integers \(a>1\) and \(b>1\) such that \(a^b+b^a\) is divisible by \(a+b\).
[ "Let \\(n=a+b\\). Since \\(gcd(a,b)=1\\), we know \\(gcd(a,n)=1\\). We can rewrite the condition as\n\n\\[\na^{n-a}+(n-a)^a \\equiv 0 \\mod{n}\n\\]\n\n\\[\na^{n-a}\\equiv-(-a)^a \\mod{n}\n\\]\n\nAssume \\(a\\) is odd. Since we need to prove an infinite number of pairs exist, it suffices to show that infinitely many pairs with \\(a\\) odd exist.\n\nThen we have\n\n\\[\na^{n-a}\\equiv a^a \\mod{n}\n\\]\n\n\\[\n1 \\equiv a^{2a-n} \\mod{n}\n\\]\n\nWe know by Euler's theorem that \\(a^{\\varphi(n)} \\equiv 1 \\mod{n}\\), so if \\(2a-n=\\varphi(n)\\) we will have the required condition.\n\nThis means \\(a=\\frac{n+\\varphi(n)}{2}\\). Let \\(n=2p\\) where \\(p\\) is a prime, \\(p\\equiv 1\\mod{4}\\). Then \\(\\varphi(n) = 2p*\\left(1-\\frac{1}{2}\\right)\\left(1-\\frac{1}{p}\\right) = p-1\\), so\n\n\\[\na = \\frac{2p+p-1}{2} = \\frac{3p-1}{2}\n\\]\n\nNote the condition that \\(p\\equiv 1\\mod{4}\\) guarantees that \\(a\\) is odd, since \\(3p-1 \\equiv 2\\mod{4}\\)\n\nThis makes \\(b = \\frac{p+1}{2}\\). Now we need to show that \\(a\\) and \\(b\\) are relatively prime. We see that\n\n\\[\ngcd\\left(\\frac{3p-1}{2},\\frac{p+1}2\\right)=\\frac{gcd(3p-1,p+1)}{2}\n\\]\n\n\\[\n=\\frac{gcd(p-3,4)}{2}=\\frac22=1\n\\]\n\nBy the Euclidean Algorithm.\n\nTherefore, for all primes \\(p \\equiv 1\\mod{4}\\), the pair \\(\\left(\\frac{3p-1}{2},\\frac{p+1}{2}\\right)\\) satisfies the criteria, so infinitely many such pairs exist.", "Take \\(a=2n-1, b=2n+1, n\\geq 2\\). It is obvious (use the Euclidean Algorithm, if you like), that \\(\\gcd(a,b)=1\\), and that \\(a,b>1\\).\n\nNote that\n\n\\[\na^2 = 4n^2-4n+1 \\equiv 1 \\pmod{4n}\n\\]\n\n\\[\nb^2 = 4n^2+4n+1 \\equiv 1 \\pmod{4n}\n\\]\n\nSo\n\n\\[\na^b+b^a = a(a^2)^n+b(b^2)^{n-1} \\equiv\n\\]\n\n\\[\na\\cdot 1^n + b\\cdot 1^{n-1} \\equiv a+b = 4n \\equiv 0 \\pmod{4n}\n\\]\n\nSince \\(a+b=4n\\), all such pairs work, and we are done.", "Let \\(x\\) be odd where \\(x>1\\). We have \\(x^2-1=(x-1)(x+1),\\) so \\(x^2-1 \\equiv 0 \\pmod{2x+2}.\\) This means that \\(x^{x+2}-x^x \\equiv 0 \\pmod{2x+2},\\) and since x is odd, \\(x^{x+2}+(-x)^x \\equiv 0 \\pmod{2x+2},\\) or \\(x^{x+2}+(x+2)^x \\equiv 0 \\pmod{2x+2},\\) as desired.", "I claim that the ordered pair \\((2^{n} - 1, 2^{n} + 1)\\) satisfies the criteria for all \\(n \\geq 2.\\) Proof: It is easy to see that the order modulo \\(2^{n+1}\\) of \\((2^{n} - 1)^k\\) is \\(2,\\) since \\((2^{n} - 1)^2 = 2^{2n} - 2 \\cdot 2^{n} + 1 \\equiv 1 \\mod 2^{n+1}.\\), and we can assert and prove similarly for the order modulo \\(2^{n+1}\\) of \\((2^{n} + 1)^k.\\) Thus, the remainders modulo \\(2^{n+1}\\) that the sequences of powers of \\(2^{n} - 1\\) and \\(2^{n} + 1\\) generate are \\(1\\) for even powers and \\(2^{n} - 1\\) and \\(2^{n} + 1\\) for odd powers, respectively. Since \\(2^{n} + 1\\) and \\(2^{n} - 1\\) are both odd for \\(n \\geq 2,\\)\n\n\\[\n(2^n + 1)^{2^n - 1} + (2^n - 1)^{2^n + 1} \\equiv 0 \\mod 2^{n+1}.\n\\]\n\nSince there are infinitely many powers of \\(2\\) and since all ordered pairs \\((2^n - 1, 2^n+1)\\) contain relatively prime integers, we are done. \\(\\boxed{}\\)\n\n-fidgetboss_4000", "Note that\n\n\\[\na^b+b^a=a^b-a^a+a^a+b^a.\n\\]\n\nTo get rid of the \\(a^a+b^a\\) part \\(\\pmod{a+b},\\) we can use the sum of powers factorization. However, \\(a\\) must be odd for us to do this. If we assume that \\(a\\) is odd,\n\n\\[\na^b-a^a+a^a+b^a\\equiv a^b-a^a+(a+b)(\\text{an integer})\\equiv a^b-a^a \\equiv a^a\\left(a^{b-a}-1\\right)\\pmod{a+b}.\n\\]\n\nBecause \\(a\\) and \\(b\\) are relatively prime, \\(a+b\\) cannot divide \\(a^a.\\) Thus, we have to show that there exists an integer \\(b\\) such that for odd \\(a,\\)\n\n\\[\na^{b-a}\\equiv 1 \\pmod{a+b}.\n\\]\n\nSuppose that \\(a=2n-1.\\) To keep the powers small, we try \\(b=2n+k\\) for small values of \\(k.\\) We can find that \\(b=2n\\) does not work. \\(b=2n+1\\) works though, as \\(a+b=4n\\) and\n\n\\[\n(2n-1)^{2n+1-2n+1}\\equiv (2n-1)^2\\equiv 4n^2-4n+1\\equiv 4n(n-1)+1 \\equiv 1 \\pmod{4n}.\n\\]\n\nBecause \\(a\\) is odd, \\(b=a+2\\) is relatively prime to \\(a.\\) Thus,\n\n\\[\n(a,b)=(2n-1,2n+1)\n\\]\n\nis a solution for positive \\(n\\ge 2.\\) There are infinitely many possible values for \\(n,\\) so the proof is complete. \\(\\blacksquare\\)\n\n~BS2012" ]
USAMO-2017-2
https://artofproblemsolving.com/wiki/index.php/2017_USAMO_Problems/Problem_2
Let \(m_1, m_2, \ldots, m_n\) be a collection of \(n\) positive integers, not necessarily distinct. For any sequence of integers \(A = (a_1, \ldots, a_n)\) and any permutation \(w = w_1, \ldots, w_n\) of \(m_1, \ldots, m_n\), define an \(A\)-inversion of \(w\) to be a pair of entries \(w_i, w_j\) with \(i < j\) for which one of the following conditions holds: \[ a_i \ge w_i > w_j, \] \[ w_j > a_i \ge w_i, \] or \[ w_i > w_j > a_i. \] Show that, for any two sequences of integers \(A = (a_1, \ldots, a_n)\) and \(B = (b_1, \ldots, b_n)\), and for any positive integer \(k\), the number of permutations of \(m_1, \ldots, m_n\) having exactly \(k\) \(A\)-inversions is equal to the number of permutations of \(m_1, \ldots, m_n\) having exactly \(k\) \(B\)-inversions.
[]
USAMO-2017-3
https://artofproblemsolving.com/wiki/index.php/2017_USAMO_Problems/Problem_3
Let \(ABC\) be a scalene triangle with circumcircle \(\Omega\) and incenter \(I.\) Ray \(AI\) meets \(BC\) at \(D\) and \(\Omega\) again at \(M;\) the circle with diameter \(DM\) cuts \(\Omega\) again at \(K.\) Lines \(MK\) and \(BC\) meet at \(S,\) and \(N\) is the midpoint of \(IS.\) The circumcircles of \(\triangle KID\) and \(\triangle MAN\) intersect at points \(L_1\) and \(L.\) Prove that \(\Omega\) passes through the midpoint of either \(IL_1\) or \(IL.\)
[ "Let \\(X\\) be the point on circle \\(\\Omega\\) opposite \\(M\\). This means \\(\\angle MAX = 90^\\circ, BC \\perp XM.\\)\n\n\\(\\angle XKM = \\angle DKM = 90^\\circ \\implies\\) the points \\(X, D,\\) and \\(K\\) are collinear.\n\nLet \\(D' = BC \\cap XM \\implies DD' \\perp XM \\implies\\)\n\n\\(S\\) is the orthocenter of \\(\\triangle DMX \\implies\\) the points \\(X, A,\\) and \\(S\\) are collinear.\n\nLet \\(\\omega\\) be the circle centered at \\(S\\) with radius \\(R = \\sqrt {SK \\cdot SM}.\\)\n\nWe denote \\(I_\\omega\\) inversion with respect to \\(\\omega.\\)\n\nNote that the circle \\(\\Omega\\) has diameter \\(MX\\) and contain points \\(A, B, C,\\) and \\(K.\\)\n\n\\(I_\\omega (K) = M \\implies\\) circle \\(\\Omega \\perp \\omega \\implies C = I_\\omega (B), X = I_\\omega (A).\\)\n\n\\(I_\\omega (K) = M \\implies\\) circle \\(KMD \\perp \\omega \\implies D' = I_\\omega (D) \\in KMD \\implies\\) \\(\\angle DD'M = 90^\\circ \\implies\\) the points \\(X, D',\\) and \\(M\\) are collinear.\n\nLet \\(F \\in AM, MF = MI.\\) It is well known that \\(MB = MI = MC \\implies\\)\n\n\\(\\Theta = BICF\\) is circle centered at \\(M.\\) \\(C = I_\\omega (B) \\implies \\Theta \\perp \\omega.\\)\n\nLet \\(I' = I_\\omega (I ) \\implies I' \\in \\Theta \\implies \\angle II'M = 90^\\circ.\\) \\(I' = I_\\omega (I ), X = I_\\omega (A ) \\implies AII'X\\) is cyclic.\n\n\\(\\angle XI'I = \\angle XAI = 90^\\circ \\implies\\) the points \\(X, I' ,\\) and \\(F\\) are collinear.\n\n\\(I'IDD'\\) is cyclic \\(\\implies \\angle I'D'M = \\angle I'D'C + 90^\\circ = \\angle I'ID + 90^\\circ,\\) \\(\\angle XFM = \\angle I'FI = 90^\\circ – \\angle I'IF = 90^\\circ – \\angle I'ID \\implies\\)\n\n\\(\\angle XFM + \\angle I'D'M = 180^\\circ \\implies I'D'MF\\) is cyclic.\n\nTherefore point \\(F\\) lies on \\(I_\\omega (IDK).\\)\n\n\\(FA \\perp SX, SI' \\perp FX \\implies I\\) is orthocenter of \\(\\triangle FSX.\\)\n\n\\(N\\) is midpoint \\(SI, M\\) is midpoint \\(FI, I\\) is orthocenter of \\(\\triangle FSX, A\\) is root of height \\(FA \\implies AMN\\) is the nine-point circle of \\(\\triangle FSX \\implies I' \\in AMN.\\)\n\nLet \\(N' = I_\\omega (N) \\implies R^2 = SN \\cdot SN' = SI \\cdot SI' \\implies\\)\n\n\\[\n\\frac {SN'}{SI'} = \\frac {SI}{SN} =2 \\implies\n\\]\n\n\\(\\angle XN'I' = \\angle XSI' = 90^\\circ – \\angle AXI' = \\angle IFX \\implies N'XIF\\) is cyclic.\n\nTherefore point \\(F\\) lies on \\(I_\\omega (AMN) \\implies I_\\omega(F) = L \\implies\\)\n\nThe points \\(F, L,\\) and \\(S\\) are collinear, \\(AXFL\\) is cyclic.\n\nPoint \\(I\\) is orthocenter \\(\\triangle FSX \\implies XI \\perp SF, \\angle ILS = \\angle SI'F = 90^\\circ\\) \\(\\implies\\) The points \\(X, I, E,\\) and \\(L\\) are collinear.\n\n\\(AXFL\\) is circle \\(\\implies AI \\cdot IF = IL \\cdot XI\\implies\\)\n\n\\(AI \\cdot \\frac {IF}{2} = \\frac {IL}{2} \\cdot IX \\implies AI \\cdot IM = EI \\cdot IX \\implies AEMX\\) is cyclic.\n\n\\[\nE \\in \\Omega.\n\\]\n\nvladimir.shelomovskii@gmail.com, vvsss\n\n## Contact\n\nContact v_Enhance at https://www.facebook.com/v.Enhance.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2017-4
https://artofproblemsolving.com/wiki/index.php/2017_USAMO_Problems/Problem_4
Let \(P_1, \ldots, P_{2n}\) be \(2n\) distinct points on the unit circle \(x^2 + y^2 = 1\) other than \((1,0)\). Each point is colored either red or blue, with exactly \(n\) of them red and exactly \(n\) of them blue. Let \(R_1, \ldots, R_n\) be any ordering of the red points. Let \(B_1\) be the nearest blue point to \(R_1\) traveling counterclockwise around the circle starting from \(R_1\). Then let \(B_2\) be the nearest of the remaining blue points to \(R_2\) traveling counterclockwise around the circle from \(R_2\), and so on, until we have labeled all the blue points \(B_1, \ldots, B_n\). Show that the number of counterclockwise arcs of the form \(R_i \rightarrow B_i\) that contain the point \((1,0)\) is independent of the way we chose the ordering \(R_1, \ldots, R_n\) of the red points.
[ "I define a sequence to be, starting at \\((1,0)\\) and tracing the circle counterclockwise, and writing the color of the points in that order - either R or B. For example, possible sequences include \\(RB\\), \\(RBBR\\), \\(BBRRRB\\), \\(BRBRRBBR\\), etc. Note that choosing an \\(R_1\\) is equivalent to choosing an \\(R\\) in a sequence, and \\(B_1\\) is defined as the \\(B\\) closest to \\(R_1\\) when moving rightwards. If no \\(B\\)s exist to the right of \\(R_1\\), start from the far left. For example, if I have the above example \\(RBBR\\), and I define the 2nd \\(R\\) to be \\(R_1\\), then the first \\(B\\) will be \\(B_1\\). Because no \\(R\\) or \\(B\\) can be named twice, I can simply remove \\(R_1\\) and \\(B_1\\) from my sequence when I choose them. I define this to be a move. Hence, a possible move sequence of \\(BBRRRB\\) is: \\(BBR_1RRB_1\\implies B_2BRR_2\\implies B_3R_3\\)\n\nNote that, if, in a move, \\(B_n\\) appears to the left of \\(R_n\\), then \\(\\stackrel{\\frown}{R_nB_n}\\) intersects \\((1,0)\\)\n\nNow, I define a commencing \\(B\\) to be a \\(B\\) which appears to the left of all \\(R\\)s, and a terminating \\(R\\) to be a \\(R\\) which appears to the right of all \\(B\\)s. Let the amount of commencing \\(B\\)s be \\(j\\), and the amount of terminating \\(R\\)s be \\(k\\), I claim that the number of arcs which cross \\((1,0)\\) is constant, and it is equal to \\(\\text{max}(j,k)\\). I will show this with induction.\n\nBase case is when \\(n=1\\). In this case, there are only two possible sequences - \\(RB\\) and \\(BR\\). In the first case, \\(\\stackrel{\\frown}{R_1B_1}\\) does not cross \\((1, 0)\\), but both \\(j\\) and \\(k\\) are \\(0\\), so \\(\\text{max}(j,k)=0\\). In the second example, \\(j=1\\), \\(k=1\\), so \\(\\text{max}(j,k)=1\\). \\(\\stackrel{\\frown}{R_1B_1}\\) crosses \\((1,0)\\) since \\(B_1\\) appears to the left of \\(R_1\\), so there is one arc which intersects. Hence, the base case is proved.\n\nFor the inductive step, suppose that for a positive number \\(n\\), the number of arcs which cross \\((1,0)\\) is constant, and given by \\(\\text{max}(j, k)\\) for any configuration. Now, I will show it for \\(n+1\\).\n\nSuppose I first choose \\(R_1\\) such that \\(B_1\\) is to the right of \\(R_1\\) in the sequence. This implies that \\(\\stackrel{\\frown}{R_1B_1}\\) does not cross \\((1,0)\\). But, neither \\(R_1\\) nor \\(B_1\\) is a commencing \\(B\\) or terminating \\(R\\). These numbers remain constant, and now after this move we have a sequence of length \\(2n\\). Hence, by assumption, the total amount of arcs is \\(0+\\text{max}(j,k)=\\text{max}(j,k)\\).\n\n- Here is a counter-case. \\(BRR_{1}B_{1}RR\\) : j = 1, k = 2 => \\(BRRR\\) : j = 1, k = 3. These numbers may not remain constant.\n\nThus, this solution probably doesn't work.\n\nNow suppose that \\(R_1\\) appears to the right of \\(B_1\\), but \\(B_1\\) is not a commencing \\(B\\). This implies that there are no commencing \\(B\\)s in the series, because there are no \\(B\\)s to the left of \\(B_1\\), so \\(j=0\\). Note that this arc does intersect \\((1,0)\\), and \\(R_1\\) must be a terminating \\(R\\). \\(R_1\\) must be a terminating \\(R\\) because there are no \\(B\\)s to the right of \\(R_1\\), or else that \\(B\\) would be \\(B_1\\). The \\(2n\\) length sequence that remains has \\(0\\) commencing \\(B\\)s and \\(k-1\\) terminating \\(R\\)s. Hence, by assumption, the total amount of arcs is \\(1+\\text{max}(0,k-1)=1+k-1=k=\\text{max}(j,k)\\).\n\nFinally, suppose that \\(R_1\\) appears to the right of \\(B_1\\), and \\(B_1\\) is a commencing \\(B\\). We know that this arc will cross \\((1,0)\\). Analogous to the previous case, \\(R_1\\) is a terminating \\(R\\), so the \\(2n\\) length sequence which remains has \\(j-1\\) commencing \\(B\\)s and \\(k-1\\) terminating \\(R\\)s. Hence, by assumption, the total amount of arcs is \\(1+\\text{max}(j-1,k-1)=1+\\text{max}(j,k)-1=\\text{max}(j,k)\\).\n\nThere are no more possible cases, hence the induction is complete, and the number of arcs which intersect \\((1,0)\\) is indeed a constant which is given by \\(\\text{max}(j,k)\\).\n\n-william122", "Lemma: If we switch the ordering of two consecutive \\(R_k\\), \\(R_{k+1}\\), the number of arcs crossing \\((1,0)\\) stays invariant.\n\nProof: There are two situations. If the two arcs don't cross this is simple because the actual arcs stay the same, and only the number order of the arcs change. Otherwise, if \\(B_k\\) is to the left of \\(R_{k+1}\\), the two arcs will have some overlap. The order of the points counterclockwise must be RRBB or some rotation of it. Notice how in both the old and the new ordering, the arc between the second R and the first B is counted twice and the arcs between the same colors are only counted once. Thus no matter where \\((1,0)\\) is the number of arcs containing it will stay invariant.\n\nLemma 2: We can swap the place of any two points \\(R_i\\) and \\(R_j\\) while keeping the number of crossings invariant.\n\nProof: By Lemma 1 we can use consecutive arc moves to swap \\(R_i\\) with \\(R_{i+1}\\), then the new \\(R_{i+1}\\) with \\(R_{i+2}\\) and so on until it swaps with \\(R_j\\). Now, we can do the reverse swaps to move the new \\(R_{j-1}\\) to \\(R_{j-2}\\), and so on into the original position of \\(R_i\\). Since the other points were moved once in one direction and once in the opposite, all the other points are kept in place.\n\nNow, let \\(I\\) be the sequence where \\(R_1\\) through \\(R_n\\) is numbered counterclockwise. We show that any ordering has the same number of crossings as \\(I\\). We can swap the current place of \\(R_1\\) with the desired position in \\(I\\). Now we can ignore \\(R_1\\) and swap the currrent place of \\(R_2\\) with the place of \\(R_2\\) in \\(I\\), and so on until all the points are in the desired position. Thus, all orderings of \\(R_1\\) to \\(R_n\\) have the same number of crossings as \\(I\\).\n\n-tigershark22" ]
USAMO-2017-5
https://artofproblemsolving.com/wiki/index.php/2017_USAMO_Problems/Problem_5
Let \(\mathbf{Z}\) denote the set of all integers. Find all real numbers \(c > 0\) such that there exists a labeling of the lattice points \(( x, y ) \in \mathbf{Z}^2\) with positive integers for which: only finitely many distinct labels occur, and for each label \(i\), the distance between any two points labeled \(i\) is at least \(c^i\).
[ "See page 11 of this PDF: https://web.evanchen.cc/exams/USAMO-2017-notes.pdf", "For \\(c\\le 1,\\) we can label every lattice point \\(1.\\) For \\(c\\le \\sqrt[4]{2},\\) we can make a \"checkerboard\" labeling, i.e. label \\((x, y)\\) with \\(1\\) if \\(x+y\\) is even and \\(2\\) if \\(x+y\\) is odd. One can easily verify that these labelings satisfy the required conditions. Therefore, a labeling as desired exists for all \\(0 < c\\le \\sqrt[4]{2}.\\)\n\nAn iterated version of the checkerboard labeling can actually work for all values \\(c < \\sqrt{2}.\\) For convenience, define the original lattice grid to be the set of all lattice points in the coordinate plane. Define a modified lattice grid of size \\(x\\) to be a structure similar to the lattice points on the coordinate plane, but with the minimum separation between any two points equaling \\(x\\) (as opposed to \\(1\\)).\n\nOn the first step, assign a label \\(1\\) to half of the points in a checkerboard arrangement. One can see that the points that have not yet been labeled form a modified lattice grid of size \\(\\sqrt{2}\\) (this lattice grid is also rotated by \\(45^{\\circ}\\) from the original lattice grid). At this point, for the second step, assign a label \\(2\\) to half of the points, again in a checkerboard arrangement. At this point, the points that have not yet been labeled form a modified lattice grid of size \\(2\\) (and again, it is rotated \\(45^{\\circ}\\) from the modified lattice grid after the first step). One then continues in this fashion. For the \\(N^{\\text{th}}\\) step, the points we are labeling are separated by at least \\(\\sqrt{2}\\times\\left(\\sqrt{2}\\right)^{N-1} = \\left(\\sqrt{2}\\right)^N > c^N,\\) so we know that our labeling at each step is acceptable.\n\nAfter the \\(N^{\\text{th}}\\) step (where \\(N\\) is a natural number), the points that have not yet been labeled form a modified lattice grid with size \\(\\left(\\sqrt{2}\\right)^N.\\) Since \\(c < \\sqrt{2},\\) we will eventually have \\(\\left(\\sqrt{2}\\right)^N > c^{N+1}\\) for some sufficiently large \\(N.\\) At this point, we can label all remaining points in the original lattice grid \\(N+1,\\) and this produces a labeling of all of the lattice points in the plane that satisfies all of the conditions. Therefore, a labeling as desired exists for all \\(c < \\sqrt{2}.\\)\n\nWe now prove that no labeling as desired exists for any \\(c\\ge 2.\\) To do this, we will prove that labeling a \\(2^k\\)-by-\\(2^k\\) square grid of lattice points requires at least \\(k+3\\) distinct labels for all natural numbers \\(k\\); hence for a sufficiently large section of the lattice plane the number of distinct labels required grows arbitrarily large, so the entire lattice plane cannot be labeled with finitely many distinct labels. We will prove this using induction.\n\nFor the base case, \\(k=1,\\) we have four points in a square of side length \\(1.\\) The maximum distance between any two of these points is \\(\\sqrt{2} < c^1\\) for all \\(c\\ge 2,\\) so all four points must have different labels. This completes the base case.\n\nNow, for the inductive step, suppose that labeling a \\(2^k\\)-by-\\(2^k\\) square grid of lattice points requires at least \\(k+3\\) distinct labels for some natural number \\(k.\\) We will now prove that labeling a \\(2^{k+1}\\)-by-\\(2^{k+1}\\) square grid of lattice points requires at least \\(k+4\\) distinct labels.\n\nTake a \\(2^{k+1}\\)-by-\\(2^{k+1}\\) square grid of lattice points. Divide this grid into four quadrants, \\(A, B, C,\\) and \\(D.\\) By the inductive hypothesis, \\(A\\) requires at least \\(k+3\\) distinct labels. At least one of these labels must be \\(k+3\\) or greater; take one such label and call it \\(L.\\)\n\nThe largest distance between any two points in the entire grid is \\(\\sqrt{2}\\left(2^{k+1} - 1\\right) < c^{k+3}\\) for all \\(c\\ge 2.\\) Therefore, the label \\(L\\) cannot be used anywhere else in the grid. However, \\(B, C,\\) and \\(D\\) each require at least \\(k+3\\) distinct labels as well by the inductive hypothesis. Thus, they must use at least one label that is not used in \\(A.\\) It follows that the entire grid requires at least \\(k+4\\) distinct labels. This completes the inductive step, and thus we conclude that no labeling as desired exists for any \\(c\\ge 2.\\)\n\nI have heard from others that the actual boundary is \\(\\sqrt{2}.\\) This makes intuitive sense, since the iterated checkerboard labeling outlined above just breaks down at this value (you will be able to get closer and closer to labeling all of the lattice points, but you can never get there, since you will never have \\(\\left(\\sqrt{2}\\right)^N > c^{N+1}\\)). The inductive argument above seems fairly loose, so I think that it can be sharpened to bring the upper bound down to \\(\\sqrt{2},\\) but I am not sure yet how exactly to do so. I think the way to do it is to somehow force \\(2\\) new labels (instead of just \\(1\\)) each time you double the side length of the square grid.", "We claim that for all \\(c<\\sqrt{2}\\), there exists such a configuration.\n\nLemma 1a: If there exists a configuration for \\(c=c_1\\), then for any \\(c_2\\le{c_1}\\), there also exists a configuration for \\(c=c_2\\). Note that any labeling that can be done for \\(c=c_1\\) exists for \\(c=c_2\\). Thus, we may copy the configuration from \\(c=c_1\\) to \\(c=c_2\\).\n\nLemma 1b: If there does not exist a configuration for \\(c=c_1\\), then for any \\(c_2\\ge{c_1}\\), there does not exist a configuration for \\(c=c_2\\). Note that any labeling which may be done for \\(c=c_2\\) should also be available for \\(c=c_1\\). However, we already stated that \\(c=c_1\\) has no configuration, and so \\(c=c_2\\) doesn't either.\n\nWe now prove that \\(c=\\sqrt{2}\\) does not work. WLOG, label the origin of our lattice plan with label \\(1\\). WLOG, we can cover every point that we can with label \\(1\\) because not labeling any such possible point does not affect the distance between points that aren't allowed to be labelled.\n\nName the state of the lattice grid before using label \\(1\\) state \\(S\\). After covering every point possible, we are left with only lattice points that have coordinates that sum to an odd integer. Notice, however, that if we discard the covered points and rotate all the new points by \\(\\frac{\\pi}{4}\\) radians counterclockwise, we end up with the same grid as state \\(S\\), only all the distances are multiplied by scalar \\(\\sqrt{2}\\). However, before placing label \\(2\\)s, note that we are in the same state \\(S\\) as before, because our minimum distance is also multiplied by scalar \\(\\sqrt{2}\\). Thus, we will always be in an infinite loop, stuck in state \\(S\\), and thus it is impossible to complete the labeling.\n\nIt suffices to show that any \\(c=2^{k}\\) works given \\(k<0.5\\). Note that the two smallest distances between any two points in the starting grid are \\(1\\) and \\(\\sqrt{2}\\). When we are currently in state \\(S\\) using label \\(I\\), it is obvious that if \\(2ki-i\\le{1}\\), we can break out of state \\(S\\) (Our effective minimum length would be at most 1, and we can cover every remaining point). Because \\(k<0.5\\), there does exist an \\(I\\) where this occurs. Thus, any \\(c<\\sqrt{2}\\) works, hence proved. \\(\\blacksquare{}\\) ~SigmaPiE" ]
USAMO-2017-6
https://artofproblemsolving.com/wiki/index.php/2017_USAMO_Problems/Problem_6
Find the minimum possible value of \[ \frac{a}{b^3+4}+\frac{b}{c^3+4}+\frac{c}{d^3+4}+\frac{d}{a^3+4}, \] given that \(a,b,c,d,\) are nonnegative real numbers such that \(a+b+c+d=4\).
[]
USAMO-2018-1
https://artofproblemsolving.com/wiki/index.php/2018_USAMO_Problems/Problem_1
Let \(a,b,c\) be positive real numbers such that \(a+b+c=4\sqrt[3]{abc}\). Prove that \[ 2(ab+bc+ca)+4\min(a^2,b^2,c^2)\ge a^2+b^2+c^2. \]
[ "WLOG let \\(a \\leq b \\leq c\\). Add \\(2(ab+bc+ca)\\) to both sides of the inequality and factor to get:\n\n\\[\n4(a(a+b+c)+bc) \\geq (a+b+c)^2\n\\]\n\n\\[\n\\frac{4a\\sqrt[3]{abc}+bc}{2} \\geq 2\\sqrt[3]{a^2b^2c^2}\n\\]\n\nThe last inequality is true by AM-GM. Since all these steps are reversible, the proof is complete.\n\n- It should actually be 4(a)(a+b+c) + 4bc which results in a wrong inequality by AM-GM\n\nHence, the solution is wrong.", "https://wiki-images.artofproblemsolving.com//6/69/IMG_8946.jpg\n\n-srisainandan6", "Similarly to Solution 2, we will prove homogeneity but we will use that to solve the problem differently. Let \\(f(a,b,c)=a+b+c-4\\sqrt[3]{abc}\\). Note that \\(f(a,b,c)=f(ka,kb,kc)\\), thus proving homogeneity.\n\nWLOG, we can scale down all variables such that the lowest one is \\(1\\). WLOG, let this be \\(a=1\\). We now have \\(1+b+c=4\\sqrt[3]{bc}\\), and we want to prove \\(2bc+2b+2c+4\\ge 1+b^2+c^2.\\) Adding \\(2bc\\) to both sides and subtracting \\(2b+2c\\) gives us \\(4bc+4\\ge 1+ (b+c)(b+c-2)\\), or \\(4bc+3\\ge (b+c)(b+c-2)\\). Let \\(\\sqrt[3]{bc}=x\\). Now, we have\n\n\\[\n4x^3+3 \\ge (4x-1)(4x-3)\n\\]\n\n\\[\n4x^3 - 16x^2 + 16x \\ge 0\n\\]\n\n\\[\n4x^2 - 16 + 16 \\ge 0\n\\]\n\n\\[\n4(x-2)^2 \\ge 0\n\\]\n\nBy the trivial inequality, this is always true. Since all these steps are reversible, the proof is complete. ~SigmaPiE", "WLOG, let \\(a \\le b\\) and \\(a \\le c\\) and add \\(2(ab+bc+ca)\\) to both sides to make the left side a square.\n\n\\[\n4(ab+bc+ca)+4a^2 \\ge (a+b+c)^2\n\\]\n\n\\[\n4bc+4a(a+b+c) \\ge (a+b+c)^2\n\\]\n\nNow we perform the substitution for \\(a+b+c\\).\n\n\\[\n4bc+4a(4(abc)^{1/3}) \\ge 16(abc)^{2/3}\n\\]\n\nMultiply both sides by \\(a\\) because we want all the terms to have a factor of \\((abc)^{1/3}\\).\n\n\\[\n4abc+16a^2(abc)^{1/3} \\ge 16a(abc)^{2/3}\n\\]\n\nDivide both sides by \\(4(abc)^{1/3}\\) .\n\n\\[\n(abc)^{2/3}+4a^2 \\ge 4a(abc)^{1/3}\n\\]\n\n\\[\n((abc)^{1/3}-2a)^2 \\ge 0\n\\]\n\nThis is true by the trivial inequality. Lastly, all the steps are reversible so the given inequality has been proved.\n\n-Themathcanadian" ]
USAMO-2018-2
https://artofproblemsolving.com/wiki/index.php/2018_USAMO_Problems/Problem_2
Find all functions \(f:(0,\infty) \to (0,\infty)\) such that \[ f\left(x+\frac{1}{y}\right)+f\left(y+\frac{1}{z}\right) + f\left(z+\frac{1}{x}\right) = 1 \] for all \(x,y,z >0\) with \(xyz =1.\)
[ "Obviously, the output of \\(f\\) lies in the interval \\((0,1)\\). Define \\(g:(0,1)\\to(0,1)\\) as \\(g(x)=f\\left(\\frac1x-1\\right)\\). Then for any \\(a,b,c\\in(0,1)\\) such that \\(a+b+c=1\\), we have \\(g(a)=f\\left(\\frac1a-1\\right)=f\\left(\\frac{1-a}a\\right)=f\\left(\\frac{b+c}a\\right)\\). We can transform \\(g(b)\\) and \\(g(c)\\) similarly:\n\n\\[\ng(a)+g(b)+g(c)=f\\left(\\frac ca+\\frac ba\\right)+f\\left(\\frac ab+\\frac cb\\right)+f\\left(\\frac bc+\\frac ac\\right)\n\\]\n\nLet \\(x=\\frac ca\\), \\(y=\\frac ab\\), \\(z=\\frac bc\\). We can see that the above expression is equal to \\(1\\). That is, for any \\(a,b,c\\in(0,1)\\) such that \\(a+b+c=1\\), \\(g(a)+g(b)+g(c)=1\\).\n\n(To motivate this, one can start by writing \\(x=\\frac ab\\), \\(y=\\frac bc\\), \\(z=\\frac ca\\), and normalizing such that \\(a+b+c=1\\).)\n\nFor convenience, we define \\(h:\\left(-\\frac13,\\frac23\\right)\\to\\left(-\\frac13,\\frac23\\right)\\) as \\(h(x)=g\\left(x+\\frac13\\right)-\\frac13\\), so that for any \\(a,b,c\\in\\left(-\\frac13,\\frac23\\right)\\) such that \\(a+b+c=0\\), we have\n\n\\[\nh(a)+h(b)+h(c)=g\\left(a+\\frac13\\right)-\\frac13+g\\left(b+\\frac13\\right)-\\frac13+g\\left(c+\\frac13\\right)-\\frac13=1-1=0.\n\\]\n\nObviously, \\(h(0)=0\\). If \\(|a|<\\frac13\\), then \\(h(a)+h(-a)+h(0)=0\\) and thus \\(h(-a)=-h(a)\\). Furthermore, if \\(a,b\\) are in the domain and \\(|a+b|<\\frac13\\), then \\(h(a)+h(b)+h(-(a+b))=0\\) and thus \\(h(a+b)=h(a)+h(b)\\).\n\nAt this point, we should realize that \\(h\\) should be of the form \\(h(x)=kx\\). We first prove this for some rational numbers. If \\(n\\) is a positive integer and \\(x\\) is a real number such that \\(|nx|<\\frac13\\), then we can repeatedly apply \\(h(a+b)=h(a)+h(b)\\) to obtain \\(h(nx)=nh(x)\\). Let \\(k=6h\\left(\\frac16\\right)\\), then for any rational number \\(r=\\frac pq\\in\\left(0,\\frac13\\right)\\) where \\(p,q\\) are positive integers, we have \\(h(r)=6p*h\\left(\\frac1{6q}\\right)=\\frac{6p}q*h\\left(\\frac16\\right)=kr\\).\n\nNext, we prove it for all real numbers in the interval \\(\\left(0,\\frac13\\right)\\). For the sake of contradiction, assume that there is some \\(x\\in\\left(0,\\frac13\\right)\\) such that \\(h(x)\\ne kx\\). Let \\(E=h(x)-kx\\), then obviously \\(0<|E|<1\\). The idea is to \"amplify\" this error until it becomes so big as to contradict the bounds on the output of \\(h\\). Let \\(N=\\left\\lceil\\frac1{|E|}\\right\\rceil\\), so that \\(N\\ge2\\) and \\(|NE|\\ge1\\). Pick any rational \\(r\\in\\left(\\frac{N-1}Nx,x\\right)\\), so that\n\n\\[\n0<x-r<N(x-r)<x<\\frac13.\n\\]\n\nAll numbers and sums are safely inside the bounds of \\(\\left(-\\frac13,\\frac13\\right)\\). Thus\n\n\\[\nh(N(x-r))=Nh(x-r)=N(h(x)+h(-r))=N(h(x)-h(r))=kN(x-r)+NE,\n\\]\n\nbut picking any rational number \\(s\\in\\left(N(x-r),\\frac13\\right)\\) gives us \\(|kN(x-r)|<|ks|\\), and since \\(ks=h(s)\\in\\left(-\\frac13,\\frac23\\right)\\), we have \\(kN(x-r)\\in\\left(-\\frac13,\\frac23\\right)\\) as well, but since \\(NE\\ge1\\), this means that \\(h(N(x-r))=kN(x-r)+NE\\notin\\left(-\\frac13,\\frac23\\right)\\), giving us the desired contradiction.\n\nWe now know that \\(h(x)=kx\\) for all \\(0<x<\\frac13\\). Since \\(h(-x)=-h(x)\\) for \\(|x|<\\frac13\\), we obtain \\(h(x)=kx\\) for all \\(|x|<\\frac13\\). For \\(x\\in\\left(\\frac13,\\frac23\\right)\\), we have \\(h(x)+h\\left(-\\frac x2\\right)+h\\left(-\\frac x2\\right)=0\\), and thus \\(h(x)=kx\\) as well. So \\(h(x)=kx\\) for all \\(x\\) in the domain. Since \\(h(x)\\) is bounded by \\(-\\frac13\\) and \\(\\frac23\\), we have \\(-\\frac12\\le k\\le1\\). It remains to work backwards to find \\(f(x)\\).\n\n\\[\n\\begin{align*} h(x) &= kx \\\\ g(x) &= kx+\\frac{1-k}3 \\\\ f(x) &= \\frac k{1+x}+\\frac{1-k}3\\quad\\left(-\\frac12\\le k\\le1\\right). \\end{align*}\n\\]\n\n- wzs26843545602" ]
USAMO-2018-3
https://artofproblemsolving.com/wiki/index.php/2018_USAMO_Problems/Problem_3
For a given integer \(n\ge 2,\) let \(\{a_1,a_2,…,a_m\}\) be the set of positive integers less than \(n\) that are relatively prime to \(n.\) Prove that if every prime that divides \(m\) also divides \(n,\) then \(a_1^k+a_2^k + \dots + a_m^k\) is divisible by \(m\) for every positive integer \(k.\)
[ "https://maa.org/sites/default/files/pdf/AMC/usamo/2018/2018USAMO.pdf The integer m in the statement of the problem is ϕ(n), where ϕ is the Euler totient function. Throughout our proof we write p s || m, if s is the greatest power of p that divides m. We begin with the following lemma: Lemma 1. If p is a prime and p s divides n for some positive integer s, then 1k + 2k + · · · + n k is divisible by p s-1 for any integer k ≥ 1. 2018 USAMO – Solutions 4 Proof. Let {a1, a2, . . . , am} be a complete reduced residue set modulo p s and m = p s-1 (p - 1). First we prove by induction on s that for any positive integer k, a k 1 + a k 2 + · · · + a k m is divisible by p s-1 . The base case s = 1 is true. Suppose the statement holds for some value of s. Consider the statement for s + 1. Note that {a1, . . . , am, ps + a1, . . . , ps + am, . . . , ps (p - 1) + a1, . . . , ps (p - 1) + am} is a complete reduced residue set modulo p s+1. Therefore, the desired sum of k-th powers is equal to a k 1 + · · · + a k m + · · · + (p s (p - 1) + a1) k + · · · + (p s (p - 1) + am) k ≡ p(a k 1 + · · · + a k m) ≡ 0 (mod p s ), where we have used the induction hypothesis for the second congruence. This gives the induction step. Now we are ready to prove the lemma. Because numbers from 1 to n can be split into blocks of consecutive numbers of length p s , it is enough to show that 1k + 2k +· · · + (p s ) k is divisible by p s-1 for any positive integer k. We use induction on s. The statement is true for s = 1. Assume the statement is true for s - 1. The sum 1 k + 2k + · · · + (p s ) k = a k 1 + a k 2 + · · · + a k m + p k � 1 k + 2k + · · · + (p s-1 ) k � is divisible by p s-1 , because p s-1 | a k 1 + · · · + a k m and by the induction hypothesis p s-2 | 1 k + · · · + (p s-1 ) k . Now we proceed to prove a second lemma, from which the statement of the problem will immediately follow: Lemma 2. Suppose p is a prime dividing n. Let {a1, . . . , am} be a complete reduced residue set mod n, and define s by p s || m. Then p s divides a k 1 + · · · + a k m for any integer k ≥ 1. Proof. We fix p, and use induction on the number of prime factors of n (counted by multiplicity) that are different from p. If there are no prime factors other than p, then n = p s+1 , m = p s (p - 1), and we proved in Lemma 1 that a k 1 +· · ·+a k m is divisible by p s . Now suppose the statement is true for n. We show that it is true for nq, where q is a prime not equal to p. Case 1. q divides n. We have p s || ϕ(n) and p s || ϕ(nq), because ϕ(nq) = qϕ(n). If {a1, a2, . . . , am} is a complete reduced residue set modulo n, then {a1, . . . , am, n + a1, . . . , n + am, . . . , n(q - 1) + a1, . . . , n(q - 1) + am} is a complete reduced residue set modulo nq. The new sum of k-th powers is equal to a k 1 + · · · + a k m + · · · + (n(q - 1) + a1) k + · · · + (n(q - 1) + am) k = mnk � 1 k + · · · + (q - 1)k � + 2018 USAMO – Solutions 5 � k 1 � n k-1 � 1 k-1 + · · · + (q - 1)k-1 � (a1 + · · · + am) + · · · + q(a k 1 + · · · + a k m). This sum is divisible by p s because p s || m and p s | a j 1 + a j 2 + · · · + a j m for any positive integer j. Case 2. q doesn't divide n. Suppose p b || q - 1, where b ≥ 0. Note that ϕ(nq) = ϕ(n)(q - 1), so p s || ϕ(n) and p s+b || ϕ(nq). Let {a1, . . . , am} be a complete reduced residue set modulo n. The complete reduced residue set modulo nq consists of the mq numbers {a1, . . . , am, n + a1, . . . , n + am, . . . , n(q - 1) + a1, . . . , n(q - 1) + am} with the m elements {qa1, qa2, . . . , qam} removed. The new sum of k-th powers is equal to a k 1 + · · · + a k m + · · · + (n(q - 1) + a1) k + · · · + (n(q - 1) + am) k - q k (a k 1 + · · · + a k m) = mnk � 1 k + · · · + (q - 1)k � + � k 1 � n k-1 � 1 k-1 + · · · + (q - 1)k-1 � (a1 + · · · + am) + · · · · · · + � k k - 1 � n (1 + · · · + (q - 1)) (a k-1 1 + · · · + a k-1 m ) + q(a k 1 + · · · + a k m) - q k (a k 1 + · · · + a k m). Each term � k j � n k-j � 1 k-j + · · · + (q - 1)k-j � (a j 1 + · · · + a j m), for 0 ≤ j ≤ k - 1, is divisible by p s+b because p | n k-j , p s | a j 1 + · · · + a j m, and p b-1 | 1 k-j + · · · + (q - 1)k-j by Lemma 1. Also (q k - q)(a k 1 + · · · + a k m) is divisible by p s+b because p b | q - 1 | q k - q and p s | a k 1 + · · · + a k m. Thus p s+b divides our sum and our proof is complete. Remark. In fact, one can also show the converse statement: if {a1, a2, . . . , am} is as defined in the problem and a k 1 + a k 2 + · · · + a k m is divisible by m for every positive integer k, then every prime that divides m also divides n. These problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2018-4
https://artofproblemsolving.com/wiki/index.php/2018_USAMO_Problems/Problem_4
Let \(p\) be a prime, and let \(a_1, \dots, a_p\) be integers. Show that there exists an integer \(k\) such that the numbers \[ a_1 + k, a_2 + 2k, \dots, a_p + pk \] produce at least \(\tfrac{1}{2} p\) distinct remainders upon division by \(p\).
[ "\\(\\textbf{Lemma: }\\) For fixed \\(i\\neq j,\\) where \\(i, j\\in\\{1, 2, ..., p\\},\\) the statement \\(a_i + ik\\equiv a_j + jk\\text{ (mod } p\\text{)}\\) holds for exactly one \\(k\\in {1, 2, ..., p}.\\)\n\n\\(\\textbf{Proof: }\\) Notice that the left side minus the right side is congruent to \\((a_i - a_j) + (i - j)k\\) modulo \\(p.\\) For this difference to equal \\(0,\\) there is a unique solution for \\(k\\) modulo \\(p\\) given by \\(k\\equiv (a_j - a_i)(i - j)^{-1}\\text{ (mod } p\\text{)},\\) where we have used the fact that every nonzero residue modulo \\(p\\) has a unique multiplicative inverse. Therefore, there is exactly one \\(k\\in {1, 2, ..., p}\\) that satisfies \\(a_i + ik\\equiv a_j + jk\\text{ (mod } p\\text{)}\\) for any fixed \\(i\\neq j. \\textbf{ End Lemma}\\)\n\nSuppose that you have \\(p\\) graphs \\(G_1, G_2, ..., G_p,\\) and graph \\(G_k\\) consists of the vertices \\((i, k)\\) for all \\(1\\le i\\le p.\\) Within any graph \\(G_k,\\) vertices \\((i_1, k)\\) and \\((i_2, k)\\) are connected by an edge if and only if \\(a_{i_1} + i_1k\\equiv a_{i_2} + i_2k\\text{ (mod } p\\text{)}.\\) Notice that the number of disconnected components of any graph \\(G_k\\) equals the number of distinct remainders when divided by \\(p\\) given by the numbers \\(a_1 + k, a_2 + 2k, ..., a_p + pk.\\)\n\nThese \\(p\\) graphs together have exactly one edge for every unordered pair of elements of \\(\\{1, 2, ..., p\\},\\) so they have a total of exactly \\(\\frac{p(p-1)}{2}\\) edges. Therefore, there exists at least one graph \\(G_k\\) that has strictly fewer than \\(\\frac{p}{2}\\) edges, meaning that it has more than \\(\\frac{p}{2}\\) disconnected components. Therefore, the collection of numbers \\(\\{a_i + ik: 1\\le i\\le p\\}\\) for this particular value of \\(k\\) has at least \\(\\frac{p}{2}\\) distinct remainders modulo \\(p.\\) This completes the proof.\n\n\\[\n\\textbf{Note: This is the same problem as 2018 USAJMO Problem 5.}\n\\]", "We consider the Lemma and it's proof as in the above solution.\n\nFor every \\(k\\), let \\(x_k\\) be the number of distinct residues that \\(a_i + ik\\) takes on. Further for each residue \\(r_i\\), let \\(n_i\\) be the number of times it is achieved by an \\(a_j + jk\\) for \\(i \\in [1, x_k]\\). Note that \\(\\sum_{i = 1}^{x_k} n_i = p.\\)\n\nThe number of pairs \\((a_i, a_j)\\) s.t \\((a_i + ik) - (a_j + jk) \\equiv 0\\) for a given \\(k\\) is,\n\n\\[\n\\begin{align*} \\sum_{i = 1}^{x_k} \\binom{n_i}{2} &\\ge x_k \\binom{(\\sum_{i = 1}^{x_k} n_i)/x_k}{2} \\\\ &= \\frac{1}{2}p(\\frac{p}{x_k} - 1). \\end{align*}\n\\]\n\nThe first equality is given by a well known theorem, which can be proven by C-S or Jensen's.\n\nSumming over all \\(k\\), the number of times a pair \\((a_i, a_j)\\) has \\((a_i + ik) - (a_j + jk) \\equiv 0\\) is,\n\n\\[\n\\begin{align*} \\sum_{k = 1}^{p} \\frac{1}{2}p(\\frac{p}{x_k} - 1). \\end{align*}\n\\]\n\nAlternatively, every pair \\((a_i, a_j)\\) has a unique \\(k\\) s.t \\((a_i + ik) - (a_j + jk) \\equiv 0\\) by the Lemma so we double-count as,\n\n\\[\n\\begin{align*} \\binom{p}{2} &\\ge \\sum_{k = 1}^{p} \\frac{1}{2}p(\\frac{p}{x_k} - 1). \\\\ 2p - 1 &\\ge \\sum_{k = 1}^{p} \\frac{p}{x_k}. \\end{align*}\n\\]\n\nBy AM-HM inequality,\n\n\\[\n\\begin{align*} \\sum_{k = 1}^{p} \\frac{p}{x_k} &\\ge \\frac{p^2}{\\sum_{k = 1}^{p} \\frac{x_k}{p}} \\\\ 2p-1 &\\ge \\frac{p^3}{\\sum_{k = 1}^{p} x_k} \\\\ \\sum_{k = 1}^{p}x_k &\\ge \\frac{p^3}{2p-1} \\ge \\frac{p^2}{2}. \\end{align*}\n\\]\n\nSo it is clear that \\(\\exists k\\) s.t \\(x_k \\ge \\frac{p}{2}\\).\n\n~Aaryabhatta1" ]
USAMO-2018-5
https://artofproblemsolving.com/wiki/index.php/2018_USAMO_Problems/Problem_5
In convex cyclic quadrilateral \(ABCD,\) we know that lines \(AC\) and \(BD\) intersect at \(E,\) lines \(AB\) and \(CD\) intersect at \(F,\) and lines \(BC\) and \(DA\) intersect at \(G.\) Suppose that the circumcircle of \(\triangle ABE\) intersects line \(CB\) at \(B\) and \(P\), and the circumcircle of \(\triangle ADE\) intersects line \(CD\) at \(D\) and \(Q\), where \(C,B,P,G\) and \(C,Q,D,F\) are collinear in that order. Prove that if lines \(FP\) and \(GQ\) intersect at \(M\), then \(\angle MAC = 90^{\circ}.\)
[ "\\[\n[asy] /* Geogebra to Asymptote conversion, documentation at artofproblemsolving.com/Wiki go to User:Azjps/geogebra */ import graph; size(13cm); real labelscalefactor = 0.5; /* changes label-to-point distance */ pen dps = linewidth(0.7) + fontsize(10); defaultpen(dps); /* default pen style */ pen dotstyle = black; /* point style */ real xmin = -14.573333333333343, xmax = 3.56, ymin = -4.74, ymax = 8.473333333333338; /* image dimensions */ pen sexdts = rgb(0.1803921568627451,0.49019607843137253,0.19607843137254902); pen cczzff = rgb(0.8,0.6,1); pen ccwwff = rgb(0.8,0.4,1); pen wvvxds = rgb(0.396078431372549,0.3411764705882353,0.8235294117647058); pen rvwvcq = rgb(0.08235294117647059,0.396078431372549,0.7529411764705882); pen wrwrwr = rgb(0.3803921568627451,0.3803921568627451,0.3803921568627451); /* draw figures */ draw(circle((-4.80390015600624,-0.5952574102964114), 2.6896620042551294), linewidth(1) + sexdts); draw((-5.58,1.98)--(-4.06,-3.18), linewidth(1) + cczzff); draw((-3.8846455730896308,1.9324397054436309)--(-7.42,-1.22), linewidth(1) + ccwwff); draw(circle((-7.256640463424001,0.8150682664688008), 2.0416143581437347), linewidth(1) + sexdts); draw(circle((-4.74356842508035,1.555353052957629), 0.9380526686465809), linewidth(1) + sexdts); draw((-13.275816730447621,2.195893092761112)--(-3.912501589632712,1.120300499610314), linewidth(1) + wvvxds); draw((-3.775365275873233,5.118495172394378)--(-8.947688716232484,-0.3288482488643849), linewidth(1) + wvvxds); draw((-5.58,1.98)--(-3.775365275873233,5.118495172394378), linewidth(1) + wvvxds); draw((-5.58,1.98)--(-7.42,-1.22), linewidth(1) + wvvxds); draw((-3.8846455730896308,1.9324397054436309)--(-4.06,-3.18), linewidth(1) + rvwvcq); draw((-3.8846455730896308,1.9324397054436309)--(-3.775365275873233,5.118495172394378), linewidth(1) + rvwvcq); draw((-13.275816730447621,2.195893092761112)--(-4.06,-3.18), linewidth(1) + rvwvcq); draw((-13.275816730447621,2.195893092761112)--(-3.775365275873233,5.118495172394378), linewidth(1) + rvwvcq); draw((-8.947688716232484,-0.3288482488643849)--(-0.2874232022466262,3.539053630345969), linewidth(1) + rvwvcq); draw((-7.42,-1.22)--(-0.2874232022466262,3.539053630345969), linewidth(1) + rvwvcq); draw((-0.2874232022466262,3.539053630345969)--(-4.06,-3.18), linewidth(1) + rvwvcq); draw(circle((-6.783982903277441,1.6253383695771872), 2.915556239332651), linewidth(1) + rvwvcq); draw((-5.58,1.98)--(-3.8846455730896308,1.9324397054436309), linewidth(1) + ccwwff); draw((-5.216225985226909,0.7450829498492438)--(-3.912501589632712,1.120300499610314), linewidth(1) + ccwwff); draw((-8.947688716232484,-0.3288482488643849)--(-5.58,1.98), linewidth(1) + ccwwff); draw((-8.947688716232484,-0.3288482488643849)--(-5.216225985226909,0.7450829498492438), linewidth(1) + ccwwff); draw((-5.58,1.98)--(-3.912501589632712,1.120300499610314), linewidth(1) + ccwwff); /* dots and labels */ dot((-5.58,1.98),dotstyle); label(\"$A$\", (-5.52,2.113333333333337), N * labelscalefactor); dot((-7.42,-1.22),dotstyle); label(\"$B$\", (-7.36,-1.0866666666666638), SW * labelscalefactor); dot((-4.06,-3.18),dotstyle); label(\"$C$\", (-4,-3.046666666666664), NE * labelscalefactor); dot((-3.8846455730896308,1.9324397054436309),dotstyle); label(\"$D$\", (-3.8266666666666733,2.06), NE * labelscalefactor); dot((-5.216225985226909,0.7450829498492438),linewidth(4pt) + dotstyle); label(\"$E$\", (-5.16,0.8466666666666699), NE * labelscalefactor); dot((-3.775365275873233,5.118495172394378),linewidth(4pt) + dotstyle); label(\"$F$\", (-3.72,5.22), NE * labelscalefactor); dot((-13.275816730447621,2.195893092761112),linewidth(4pt) + dotstyle); label(\"$G$\", (-13.226666666666675,2.3), NE * labelscalefactor); dot((-8.947688716232484,-0.3288482488643849),linewidth(4pt) + dotstyle); label(\"$P$\", (-8.893333333333342,-0.22), NE * labelscalefactor); dot((-3.912501589632712,1.120300499610314),linewidth(4pt) + dotstyle); label(\"$Q$\", (-3.88,1.18), NE * labelscalefactor); dot((-0.2874232022466262,3.539053630345969),linewidth(4pt) + dotstyle); label(\"$X$\", (-0.24,3.6466666666666705), NE * labelscalefactor); dot((-7.211833579631486,1.4993048370077748),linewidth(4pt) + dotstyle); label(\"$M$\", (-7.16,1.60666666666667), NE * labelscalefactor); clip((xmin,ymin)--(xmin,ymax)--(xmax,ymax)--(xmax,ymin)--cycle); /* end of picture */ [/asy]\n\\]\n\n\\[\n\\begin{align*} &\\mathrel{\\phantom{=}}\\angle DEQ+\\angle AED+\\angle AEP\\\\ &=\\angle DAQ+\\angle AQD+\\angle AEP\\\\ &=180-\\angle ADC+\\angle AEP\\\\ &=180-\\angle ADC+\\angle ABP\\\\ &=\\angle ABP+\\angle ABC\\\\ &=180 \\end{align*}\n\\]\n\nso \\(P,E,Q\\) are collinear. Furthermore, note that \\(DQBP\\) is cyclic because:\n\n\\[\n\\angle EDQ = \\angle BAE = BPE.\n\\]\n\nNotice that since \\(A\\) is the intersection of \\((EDQ)\\) and \\((BPE)\\), it is the Miquel point of \\(DQBP\\).\n\nNow define \\(X\\) as the intersection of \\(BQ\\) and \\(DP\\). From Pappus's theorem on \\(BFPDGQ\\) that \\(A,M,X\\) are collinear. It's a well known property of Miquel points that \\(\\angle EAX = 90\\), so it follows that \\(MA \\perp AE\\), as desired. \\(\\blacksquare\\) ~AopsUser101" ]
USAMO-2018-6
https://artofproblemsolving.com/wiki/index.php/2018_USAMO_Problems/Problem_6
Let \(a_n\) be the number of permutations \((x_1, x_2, \dots, x_n)\) of the numbers \((1,2,\dots, n)\) such that the \(n\) ratios \(\frac{x_k}{k}\) for \(1\le k\le n\) are all distinct. Prove that \(a_n\) is odd for all \(n\ge 1.\)
[ "Write out the mapping of each \\(k\\) to \\(x_k\\) as follows:\n\n\\[\n1\\ \\ 2\\ \\ 3\\ \\ \\dots \\ \\ n\n\\]\n\n\\[\nx_1\\ x_2\\ x_3\\ \\dots \\ x_n\n\\]\n\nNow, consider what happens when the two rows are swapped (and the top-bottom pairs are reordered so that the top reads (1,2,3,...,n)). This will result in a new permutation if and only if \\(x_k = j\\) does not imply \\(x_j = k\\) for all \\(k,j\\) in \\((1,2,3,\\dots,n)\\). I will denote this new permutation as \\((y_1,y_2,y_3,\\dots,y_n)\\). Notice that \\((y_1,y_2,y_3,\\dots,y_n)\\) is a valid permutation if and only if \\((x_1,x_2,x_3,\\dots,x_n)\\) is valid, because each fraction \\(\\frac{y_k}{k}\\) is the reciprocal of \\(\\frac{x_k}{k}\\). This means that we need only consider the parity of number of cases in which \\((y_1,y_2,y_3,\\dots, y_n) = (x_1,x_2,x_3,\\dots, x_n)\\), as there will be an even number of cases where \\((y_1,y_2,y_3,\\dots, y_n) \\ne (x_1,x_2,x_3,\\dots, x_n)\\). Let the number of valid permutations where \\((y_1,y_2,y_3,\\dots, y_n) = (x_1,x_2,x_3,\\dots, x_n)\\) be \\(b_n\\).\n\nNotice that the only permutations that have the property \\(x_{x_k}=k\\) (which is an equivalent statement to the one given above) are those that are formed by taking pairs of elements and swapping their positions having the maximum number of pairs possible and having no two pairs both contain the same element. Some more necessary conditions will be outlined below after we split \\(n\\) into two cases.\n\nCase 1: \\(n\\) is odd. If \\(n\\) is odd then there must be exactly one \\(k\\) such that \\(x_k = k\\). This yields \\(b_n=n\\cdot b_{n-1}\\)* which has the same parity as \\(b_{n-1}\\), so we need only consider the parity of \\(b_n\\) for odd \\(n\\). (*Note: This is wrong, as \\(b_n\\) is only the number of valid involutions for \\(1,2,...,n\\), not \\(1,2,...,n+1\\) with one number removed.)\n\nCase 2: \\(n\\) is even. If \\(n\\) is even then can be no \\(k\\) so that \\(x_k=k\\), or else there must be at least two distinct numbers \\(k\\) and \\(j\\) so that \\(x_k=k\\) and \\(x_j=j\\) which violates the given conditions. Denote a pair of numbers that are swapped to arrive at the final permutation as the pair \\((a,b)\\). Then if a permutation yields an invalid arrangement there must be two pairs \\((a,b)\\) and \\((c,d)\\) such that \\(\\frac{a^2}{b^2} = \\frac{c^2}{d^2}\\). But notice that the two pairs \\((a,c)\\) and \\((b,d)\\) will also result in an invalid arrangement. So, there must be an even number of invalid arrangements (Note: This proof doesn't work, as no specific pairing of invalid arrangements is constructed. Indeed, when three pairs all with the same fraction are present, this doesn't work.), meaning the parity we desire is just the number of ways to separate \\(n\\) objects into \\(\\frac n2\\) pairs! However, this is quite simply just \\((n-1)(n-3)(n-5)\\cdots(3)(1)\\), which is clearly the product of odd numbers. So we conclude that there are an odd number of valid permutations \\((x_1,x_2,x_3,\\dots,x_n)\\). \\(\\blacksquare\\)" ]
USAMO-2019-1
https://artofproblemsolving.com/wiki/index.php/2019_USAMO_Problems/Problem_1
Let \(\mathbb{N}\) be the set of positive integers. A function \(f:\mathbb{N}\to\mathbb{N}\) satisfies the equation \[ \underbrace{f(f(\ldots f}_{f(n)\text{ times}}(n)\ldots))=\frac{n^2}{f(f(n))} \] for all positive integers \(n\). Given this information, determine all possible values of \(f(1000)\).
[ "Let \\(f^r(x)\\) denote the result when \\(f\\) is applied to \\(f^{r-1}(x)\\), where \\(f^1(x)=f(x)\\). \\(\\hfill \\break \\hfill \\break\\) If \\(f(p)=f(q)\\), then \\(f^2(p)=f^2(q)\\) and \\(f^{f(p)}(p)=f^{f(q)}(q)\\)\n\n\\[\n\\implies p^2=f^2(p)\\cdot f^{f(p)}(p)=f^2(q)\\cdot f^{f(q)}(q)=q^2\n\\]\n\n\\[\n\\implies p=\\pm q\n\\]\n\n\\(\\implies p=q\\) since \\(p,q>0\\).\n\nTherefore, \\(f\\) is injective. It follows that \\(f^r\\) is also injective.\n\nLemma 1: If \\(f^r(b)=a\\) and \\(f(a)=a\\), then \\(b=a\\).\n\nProof:\n\n\\(f^r(b)=a=f^r(a)\\) which implies \\(b=a\\) by injectivity of \\(f^r\\).\n\nLemma 2: If \\(f^2(m)=f^{f(m)}(m)=m\\), and \\(m\\) is odd, then \\(f(m)=m\\).\n\nProof:\n\nLet \\(f(m)=k\\). Since \\(f^2(m)=m\\), \\(f(k)=m\\). So, \\(f^2(k)=k\\). \\(\\newline f^2(k)\\cdot f^{f(k)}(k)=k^2\\).\n\nSince \\(k\\neq0\\), \\(f^{f(k)}(k)=k\\)\n\n\\[\n\\implies f^m(k)=k\n\\]\n\n\\[\n\\implies f^{gcd(m, 2)}(k)=k\n\\]\n\n\\[\n\\implies f(k)=k\n\\]\n\nThis proves Lemma 2.\n\nI claim that \\(f(m)=m\\) for all odd \\(m\\).\n\nOtherwise, let \\(m\\) be the least counterexample.\n\nSince \\(f^2(m)\\cdot f^{f(m)}(m)=m^2\\), either\n\n\\((1) f^2(m)=k<m\\), contradicted by Lemma 1 since \\(k\\) is odd and \\(f^2(k)=k\\).\n\n\\((2) f^{f(m)}(m)=k<m\\), also contradicted by Lemma 1 by similar logic.\n\n\\((3) f^2(m)=m\\) and \\(f^{f(m)}(m)=m\\), which implies that \\(f(m)=m\\) by Lemma 2. This proves the claim.\n\nBy injectivity, \\(f(1000)\\) is not odd. I will prove that \\(f(1000)\\) can be any even number, \\(x\\). Let \\(f(1000)=x, f(x)=1000\\), and \\(f(k)=k\\) for all other \\(k\\). If \\(n\\) is equal to neither \\(1000\\) nor \\(x\\), then \\(f^2(n)\\cdot f^{f(n)}(n)=n\\cdot n=n^2\\). This satisfies the given property.\n\nIf \\(n\\) is equal to \\(1000\\) or \\(x\\), then \\(f^2(n)\\cdot f^{f(n)}(n)=n\\cdot n=n^2\\) since \\(f(n)\\) is even and \\(f^2(n)=n\\). This satisfies the given property.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2019-2
https://artofproblemsolving.com/wiki/index.php/2019_USAMO_Problems/Problem_2
Let \(ABCD\) be a cyclic quadrilateral satisfying \(AD^2 + BC^2 = AB^2\). The diagonals of \(ABCD\) intersect at \(E\). Let \(P\) be a point on side \(\overline{AB}\) satisfying \(\angle APD = \angle BPC\). Show that line \(PE\) bisects \(\overline{CD}\).
[ "Let \\(PE \\cap DC = M\\). Also, let \\(N\\) be the midpoint of \\(AB\\). Note that only one point \\(P\\) satisfies the given angle condition. With this in mind, construct \\(P'\\) with the following properties:\n\n(1) \\(AP' \\cdot AB = AD^2\\)\n\n(2) \\(BP' \\cdot AB = CD^2\\)\n\nClaim: \\(P = P'\\)\n\nProof: The conditions imply the similarities \\(ADP \\sim ABD\\) and \\(BCP \\sim BAC\\) whence \\(\\measuredangle APD = \\measuredangle BDA = \\measuredangle BCA = \\measuredangle CPB\\) as desired. \\(\\square\\)\n\nClaim: \\(PE\\) is a symmedian in \\(AEB\\)\n\nProof: We have\n\n\\[\n\\begin{align*} AP \\cdot AB = AD^2 \\iff AB^2 \\cdot AP &= AD^2 \\cdot AB \\\\ \\iff \\left( \\frac{AB}{AD} \\right)^2 &= \\frac{AB}{AP} \\\\ \\iff \\left( \\frac{AB}{AD} \\right)^2 - 1 &= \\frac{AB}{AP} - 1 \\\\ \\iff \\frac{AB^2 - AD^2}{AD^2} &= \\frac{BP}{AP} \\\\ \\iff \\left(\\frac{BC}{AD} \\right)^2 &= \\left(\\frac{BE}{AE} \\right)^2 = \\frac{BP}{AP} \\end{align*}\n\\]\n\nas desired. \\(\\square\\)\n\nSince \\(P\\) is the isogonal conjugate of \\(N\\), \\(\\measuredangle PEA = \\measuredangle MEC = \\measuredangle BEN\\). However \\(\\measuredangle MEC = \\measuredangle BEN\\) implies that \\(M\\) is the midpoint of \\(CD\\) from similar triangles, so we are done. \\(\\square\\)", "Let \\(\\omega\\) be the circle centered at \\(A\\) with radius \\(AD.\\)\n\nLet \\(\\Omega\\) be the circle centered at \\(B\\) with radius \\(BC.\\)\n\nWe denote \\(I_\\omega\\) and \\(I_\\Omega\\) inversion with respect to \\(\\omega\\) and \\(\\Omega,\\) respectively.\n\n\\[\nB'= I_\\omega (B), C'= I_\\omega (C), D = I_\\omega (D) \\implies\n\\]\n\n\\[\nAB' \\cdot AB = AD^2, \\angle ACB = \\angle AB'C'.\n\\]\n\n\\[\nA'= I_\\Omega (A), D'= I_\\Omega (D), C = I_\\Omega (C) \\implies\n\\]\n\n\\[\nBA' \\cdot AB = BC^2, \\angle BDA = \\angle BA'D'.\n\\]\n\nLet \\(\\theta\\) be the circle \\(ABCD.\\)\n\n\\(I_\\omega (\\theta) = B'C'D,\\) straight line, therefore\n\n\\[\n\\angle AB'C' = \\angle AB'D' = \\angle ACB.\n\\]\n\n\\(I_\\Omega (\\theta) = A'D'C,\\) straight line, therefore\n\n\\[\n\\angle BA'D' = \\angle BA'C = \\angle BDA.\n\\]\n\n\\(ABCD\\) is cyclic \\(\\implies \\angle BA'C = \\angle AB'D.\\)\n\n\\[\nAB' + BA' = \\frac {AD^2 + BC^2 }{AB} = AB \\implies\n\\]\n\npoints \\(A'\\) and \\(B'\\) are coincide.\n\nDenote \\(A' = B' = Q \\in AB.\\)\n\nSuppose, we move point \\(Q\\) from \\(A\\) to \\(B.\\) Then \\(\\angle AQD\\) decreases monotonically, \\(\\angle BQC\\) increases monotonically. So, there is only one point where\n\n\\[\n\\angle AQD = \\angle BQC \\implies P = Q.\n\\]\n\n\\[\nB = I_\\omega (P), D' = I_\\omega (D'), C' = I_\\omega (C), A = I_\\omega (\\infty) \\implies\n\\]\n\n\\(\\hspace{19mm} I_\\omega (CD'P) = AC'D'B\\) is cyclic.\n\n\\[\n\\angle ACD = \\angle ABD = \\angle CC'D \\implies C' D' || CD \\implies\n\\]\n\n\\(\\hspace{19mm} C'D'CD\\) is trapezoid.\n\nIt is known that the intersection of the diagonals, intersection point of the lines containing the lateral sides of the trapezoid and the midpoints of two parallel sides are collinear.\n\nvladimir.shelomovskii@gmail.com, vvsss" ]
USAMO-2019-3
https://artofproblemsolving.com/wiki/index.php/2019_USAMO_Problems/Problem_3
Let \(K\) be the set of all positive integers that do not contain the digit \(7\) in their base-\(10\) representation. Find all polynomials \(f\) with nonnegative integer coefficients such that \(f(n)\in K\) whenever \(n\in K\).
[ "I claim the only such polynomials are of the form \\(f(n)=k\\) where \\(k\\in K\\), or \\(f(n)=an+b\\) where \\(a\\) is a power of 10, \\(b\\in K\\), and \\(b<a\\). Obviously, these polynomials satisfy the conditions. We now prove that no other polynomial works. That is, we prove that for any other polynomial \\(f\\) with nonnegative coefficients, there is some \\(n\\in K\\) such that \\(f(n)\\notin K\\).\n\nWe first prove the result for monomials, polynomials in which only one coefficient is nonzero. This is obvious for constant polynomials \\(f(n)=k\\notin K\\). The next simplest case is \\(f(n)=an\\) with \\(a\\) not a power of 10, and hence \\(\\lg a\\) is irrational. By Dirichlet's approximation theorem, the set of multiples of \\(\\lg a\\) is dense \\(\\bmod\\ 1\\), and thus contains an element with fractional part in the interval \\([\\lg 7,\\lg 8)\\). In other words, there is a power of \\(a\\) whose decimal expansion starts with a 7. Let \\(a^x\\) be the smallest power of \\(a\\) that is not in \\(K\\). Obviously, \\(x>0\\), so letting \\(n=a^{x-1}\\) completes the proof of this part.\n\nWe have now proven that for any \\(a\\) that is not a power of 10, there is some \\(n\\in K\\) such that \\(an\\notin K\\). We proceed to the case where \\(f(n)=an^x\\) for \\(x>1\\). This splits into 2 cases. If \\(ax\\) is not a power of 10, then pick \\(m\\in K\\) such that \\(axm\\notin K\\). For any \\(y\\), we have\n\n\\[\nf(10^y+m)=a10^{yx}+axm*10^{y(x-1)}+...+am^x\n\\]\n\nIf we choose \\(y\\) to be large enough, then the terms in the expression above will not interfere with each other, and the resulting number will contain a 7 in the decimal expansion, and thus not be in \\(K\\).\n\nOn the other hand, if \\(ax\\) is a power of 10, then \\(a\\) and \\(x\\) are both powers of 10, and \\(x\\ge10\\). Obviously, \\(\\frac12ax(x-1)\\) is not a power of 10. By the previous step, which establishes the result for \\(x=2\\), we can pick \\(m\\in K\\) such that \\(\\frac12ax(x-1)m^2\\notin K\\). Then, for any \\(y\\),\n\n\\[\nf(10^y+m)=a10^{yx}+axm*10^{y(x-1)}+\\frac12ax(x-1)m^2*10^{y(x-2)}+...+am^x\n\\]\n\nSimilarly, picking a sufficient large \\(y\\) settles this case.\n\nNow, we extend our proof to general polynomials. If a polynomial \\(f(n)=a_0+a_1n+a_2n^2+...+a_xn^x\\) satisfies the conditions of the problem, then for any \\(m,y>0\\):\n\n\\[\nf(m*10^y)=a_xm^x*10^{yx}+...+a_1m*10^y+a_0\n\\]\n\nSimilarly, choosing \\(y\\) to be sufficiently large results in the terms not interfering with each other. If \\(f\\) contains any monomials that do not satisfy the conditions of the problem, then picking suitable \\(m\\) and sufficiently large \\(y\\) causes \\(f(m*10^y)\\) to not be in \\(K\\). Thus, \\(f\\) is a linear polynomial of the form \\(ax+b\\) where \\(a\\) is 0 or a power of 10, and \\(b\\in K\\). It suffices to rule out those polynomials where \\(a>0\\) and \\(b>a\\). If this is the case, then since the digit of \\(b\\) corresponding to \\(a\\) is not 7, there must be a single-digit number \\(n\\) such that the digit of \\(f(n)\\) corresponding to \\(a\\) is 7. Therefore, we are done.\n\n-wzs26843545602\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2019-4
https://artofproblemsolving.com/wiki/index.php/2019_USAMO_Problems/Problem_4
Let \(n\) be a nonnegative integer. Determine the number of ways that one can choose \((n+1)^2\) sets \(S_{i,j}\subseteq\{1,2,\ldots,2n\}\), for integers \(i,j\) with \(0\leq i,j\leq n\), such that: \(\bullet\) for all \(0\leq i,j\leq n\), the set \(S_{i,j}\) has \(i+j\) elements; and \(\bullet\) \(S_{i,j}\subseteq S_{k,l}\) whenever \(0\leq i\leq k\leq n\) and \(0\leq j\leq l\leq n\).
[ "Note that there are \\((2n)!\\) ways to choose \\(S_{1, 0}, S_{2, 0}... S_{n, 0}, S_{n, 1}, S_{n, 2}... S_{n, n}\\), because there are \\(2n\\) ways to choose which number \\(S_{1, 0}\\) is, \\(2n-1\\) ways to choose which number to append to make \\(S_{2, 0}\\), \\(2n-2\\) ways to choose which number to append to make \\(S_{3, 0}\\), etc. After that, note that \\(S_{n-1, 1}\\) contains the \\(n-1\\) in \\(S_{n-1, 0}\\) and 1 other element chosen from the 2 elements in \\(S_{n, 1}\\) not in \\(S_{n-1, 0}\\) so there are 2 ways for \\(S_{n-1, 1}\\). By the same logic there are 2 ways for \\(S_{n-1, 2}\\) as well so \\(2^n\\) total ways for all \\(S_{n-1, j}\\), so doing the same thing \\(n-1\\) more times yields a final answer of \\((2n)!\\cdot 2^{\\left(n^2\\right)}\\).\n\n-Stormersyle", "There are \\(\\frac{(2n)!}{2^n}\\) ways to choose \\(S_{0,0}, S_{1,1} .... S_{n,n}\\). Since, there are \\(\\binom{2n}{2}\\) ways to choose \\(S_{1,1}\\), and after that, to generate \\(S_{2,2}\\), you take \\(S_{1,1}\\) and add 2 new elements, getting you \\(\\binom{2n-2}{2}\\) ways to generate \\(S_{2,2}\\). And you can keep going down the line, and you get that there are \\(\\frac{(2n)!}{2^n}\\) ways to pick \\(S_{0,0}, S_{1,1} ... S_{n,n}\\) Then we can fill out the rest of the gird. First, let's prove a lemma.\n\n## Lemma\n\nClaim: If we know what \\(S_{a,b}\\) is and what \\(S_{a+1,b+1}\\) is, then there are 2 choices for both \\(S_{a,b+1}\\) and \\(S_{a+1,b}\\).\n\nProof: Note \\(a \\leq a+1\\) and \\(b \\leq b+1\\), so \\(S_{a,b}\\subseteq S_{a+1,b+1}\\). Let \\(A\\) be a set that contains all the elements in \\(S_{a+1,b+1}\\) that are not in \\(S_{a,b}\\). \\(A = S_{a+1,b+1} - S_{a,b}\\). We know \\(S_{a+1,b+1}\\) contains total \\(a+b+2\\) elements. And \\(S_{a,b}\\) contains total \\(a+b\\) elements. That means \\(A\\) contains only 2 elements since \\((a+b+2)-(a+b) = 2\\). Let's call these 2 elements \\(m, n\\). \\(A = \\{m, n\\}\\). \\(S_{a,b+1}\\) contains 1 elements more than \\(S_{a,b}\\) and 1 elements less than \\(S_{a+1,b+1}\\). That 1 elements has to select from \\(A\\). It's easy to see \\(S_{a,b+1} = S_{a,b}\\cup \\{m\\}\\) or \\(S_{a,b+1} = S_{a,b}\\cup \\{n\\}\\), so there are 2 choice for \\(S_{a,b+1}\\). Same thing applies to \\(S_{a+1,b}\\).\n\n## Filling in the rest of the grid\n\nWe used our proved lemma, and we can fill in \\(S_{0,1}, S_{1,2} ... S_{n-1,n}\\) then we can fill in the next diagonal, until all \\(S_{i,j}\\) are filled, where \\(i \\leq j\\). But, we haven't finished everything! Fortunately, filling out the rest of the diagonals in a similar fashion is pretty simple. And, it's easy to see that we have made \\(n(n+1)\\) decisions, each with 2 choices, when filling out the rest of the grid, so there are \\(2^{n(n+1)}\\) ways to finish off.\n\n## Finishing off\n\nTo finish off, we have \\(\\frac{(2n)!}{2^n} \\cdot 2^{n(n+1)}\\) ways to fill in the grid, which gets us \\(\\boxed{(2n)! \\cdot 2^{n^2}}\\)\n\n-Alexlikemath", "Let \\(C_{j}\\) represent the set of sets of the form \\(S_{ij}\\) for \\(1 \\le i \\le n\\), \\(a_{ij}\\) denote \\(S_{(i+1)j} \\backslash S_{ij}\\), and \\(b_{ij} = S_{i(j+1)} \\backslash S_{ij}\\). Begin by considering \\(C_0\\) and \\(S_{00} = \\emptyset\\). Then given \\(S_{i0}\\) we can create \\(S_{(i+1)0}\\) by adding one element (\\(a_{i0}\\)). Using this, the number of ways to form the sequence of \\(S_{00}, S_{10}, \\dots, S_{n0}\\) are \\((2n)(2n-1) \\cdots (n+1)\\) where we successively add one of the remaining elements of \\([2n]\\) to get consecutive terms in the sequence.\n\nNow consider when we are given \\(C_{j}\\) and we need to find \\(C_{j+1}\\). So far, there have been \\(n + j\\) chosen distinct elements (via \\(S_{nj}\\)). After finding \\(C_{j+1}\\) we will have \\(n + j + 1\\) distinct elements and so in this process we only add one unique element to sets among \\(C_{j+1}\\). There are \\(2n - (n+j) = n-j\\) ways to chose such a new element called \\(x\\).\n\nNow notice that \\(b_{0j}, a_{0(j+1)}, a_{1(j+1)}, \\dots, a_{(n-1)(j+1)}\\) is a permutation of \\(x, a_{0j}, \\dots, a_{(n-1)j}\\) by noting \\(b_{nj} = x\\) and,\n\n\\[\nS_{0j} + b_{0j} + a_{0(j+1)} + a_{1(j+1)} + \\cdots + a_{(n-1)(j+1)} = S_{n(j+1)} = S_{0j} + a_{0j}+ a_{1j} + \\cdots + a_{(n-1)j} + b_{nj}\n\\]\n\nFurthermore,\n\n\\[\n\\begin{align*} S_{0j} + b_{0j} + a_{0(j+1)} &= S_{1(j+1)} = S_{0j} + a_{0j} + b_{1j} \\\\ S_{1j} + b_{1j} + a_{1(j+1)} &= S_{2(j+1)} = S_{1j} + a_{1j} + b_{2j} \\\\ &\\cdots \\\\ S_{(n-1)j} + b_{(n-1)j} + a_{(n-1)(j+1)} &= S_{n(j+1)} = S_{(n-1)j} + a_{(n-1)j} + b_{nj} . \\end{align*}\n\\]\n\nTherefore \\(b_{ij}, a_{i(j+1)}\\) is a permutation of \\(a_{ij}, b_{(i+1)j}\\) for \\(i < n\\). Now let \\(k\\) be the first \\(i\\) such that \\(b_{ij} = x\\). By definition, \\(b_{(k+1)j}, \\dots = x\\). Then the number of ways to order \\(x, a_{0j}, \\dots, a_{(n-1)j}\\) is \\(2^{k-1}\\) as there are 2 permutations for each pair before \\(\\{b_{ij}, a_{i(j+1)}\\} = \\{a_{ij}, b_{(i+1)j}\\}\\) and each pair after is determined by \\(b_{ij}= b_{(i+1)j} = k, a_{i(j+1)} = a_{ij}\\). For \\(k = 0\\), the permutation is completely determined so there is one way.\n\nOverall the number of ways to add the \\(j+1\\)'th row is,\n\n\\[\n(n-j)(1 + \\sum_{k = 1}^n 2^{k-1}) = 2^n(n-j).\n\\]\n\nIn total, there are \\((2n)(2n-1)\\cdots (n+1)\\) ways to find \\(C_0\\) and for each \\(C_{j}\\) there are \\(2^n(n-j+1)\\) ways for \\(1 \\le j \\le n\\). So the answer is,\n\n\\[\n\\frac{2n!}{n!}(2^nn)(2^{n}(n-1)) \\cdots (2^{n}1) = \\boxed{2^{n^2} \\cdot (2n)!}.\n\\]\n\n~Aaryabhatta1\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2019-5
https://artofproblemsolving.com/wiki/index.php/2019_USAMO_Problems/Problem_5
Two rational numbers \(\frac{m}{n}\) and \(\frac{n}{m}\) are written on a blackboard, where \(m\) and \(n\) are relatively prime positive integers. At any point, Evan may pick two of the numbers \(x\) and \(y\) written on the board and write either their arithmetic mean \(\frac{x+y}{2}\) or their harmonic mean \(\frac{2xy}{x+y}\) on the board as well. Find all pairs \((m,n)\) such that Evan can write \(1\) on the board in finitely many steps. Proposed by Yannick Yao
[ "We claim that all odd \\(m, n\\) work if \\(m+n\\) is a positive power of 2.\n\nProof: We first prove that \\(m+n=2^k\\) works. By weighted averages we have that \\(\\frac{n(\\frac{m}{n})+(2^k-n)\\frac{n}{m}}{2^k}=\\frac{m+n}{2^k}=1\\) can be written, so the solution set does indeed work. We will now prove these are the only solutions.\n\nAssume that \\(m+n\\ne 2^k\\), so then \\(m+n\\equiv 0\\pmod{p}\\) for some odd prime \\(p\\). Then \\(m\\equiv -n\\pmod{p}\\), so \\(\\frac{m}{n}\\equiv \\frac{n}{m}\\equiv -1\\pmod{p}\\). We see that the arithmetic mean is \\(\\frac{-1+(-1)}{2}\\equiv -1\\pmod{p}\\) and the harmonic mean is \\(\\frac{2(-1)(-1)}{-1+(-1)}\\equiv -1\\pmod{p}\\), so if 1 can be written then \\(1\\equiv -1\\pmod{p}\\) and \\(2\\equiv 0\\pmod{p}\\) which is obviously impossible, and we are done.\n\n-Stormersyle\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2019-6
https://artofproblemsolving.com/wiki/index.php/2019_USAMO_Problems/Problem_6
Find all polynomials \(P\) with real coefficients such that \[ \frac{P(x)}{yz}+\frac{P(y)}{zx}+\frac{P(z)}{xy}=P(x-y)+P(y-z)+P(z-x) \] holds for all nonzero real numbers \(x,y,z\) satisfying \(2xyz=x+y+z\).
[ "If \\(P(x)=c\\) for a constant \\(c,\\) then \\(\\dfrac{c(x+y+z)}{xyz}=3c\\). We have \\(2c=3c.\\) Therefore \\(c=0.\\)\n\nNow consider the case of non-constant polynomials. First we have\n\n\\[\nxP(x)+yP(y)+zP(z)=xyz(P(x-y)+P(y-z)+P(z-x))\n\\]\n\nfor all nonzero real numbers \\(x,y,z\\) satisfying \\(2xyz=x+y+z\\). Both sides of the equality are polynomials (of \\(x,y,z\\)). They have the same values on the 2-dimensional surface \\(2xyz=x+y+z\\), except for some 1-dimensional curves in it. By continuity, the equality holds for all points on the surface, including those with \\(z=0.\\) Let \\(z=0,\\) we have \\(y=-x\\) and \\(x(P(x)-P(-x))=0.\\) Therefore \\(P\\) is an even function.\n\n(Here is a sketch of an elementary proof. Let \\(z=\\dfrac{x+y}{2xy-1}.\\) We have\n\n\\[\nxP(x)+yP(y)+\\dfrac{x+y}{2xy-1}P(\\dfrac{x+y}{2xy-1})=xy\\dfrac{x+y}{2xy-1}(P(x-y)+P(y-\\dfrac{x+y}{2xy-1})+P(\\dfrac{x+y}{2xy-1}-x)).\n\\]\n\nThis is an equality of rational expressions. By multiplying \\((2xy-1)^N\\) on both sides for a sufficiently large \\(N\\), they become polynomials, say \\(A(x,y)=B(x,y)\\) for all real \\(x, y\\) with \\(x\\ne 0, y\\ne 0, x+y\\ne 0\\) and \\(2xy-1\\ne 0.\\) For a fixed \\(x,\\) we have two polynomials (of \\(y\\)) having same values for infinitely many \\(y\\). They must be identical. Let \\(y=0,\\) we have \\(x^{N+1}(P(x)-P(-x))=0.\\) )\n\nNotice that if \\(P(x)\\) is a solution, then is \\(cP(x)\\) for any constant \\(c.\\) For simplicity, we assume the leading coefficient of \\(P\\) is \\(1\\):\n\n\\[\nP(x)=x^n+a_{n-2}x^{n-2}+\\cdots +a_2x^2+a_0,\n\\]\n\nwhere \\(n\\) is a positive even number.\n\nLet \\(y=\\dfrac{1}{x}\\), \\(z=x+\\dfrac{1}{x}.\\) we have\n\n\\[\nxP(x)+\\dfrac{1}{x}P\\left (\\dfrac{1}{x}\\right )+\\left ( x+\\dfrac{1}{x}\\right ) P\\left ( x+\\dfrac{1}{x}\\right ) =\\left (x+\\dfrac{1}{x}\\right )\\left ( P\\left (x-\\dfrac{1}{x}\\right )+P(-x)+P\\left (\\dfrac{1}{x}\\right )\\right ).\n\\]\n\nSimplify using \\(P(x)=P(-x),\\)\n\n\\[\n\\left (x+\\dfrac{1}{x}\\right ) \\left (P\\left (x+\\dfrac{1}{x}\\right )-P\\left (x-\\dfrac{1}{x}\\right )\\right )=\\dfrac{1}{x}P(x)+xP\\left (\\dfrac{1}{x}\\right ).\n\\]\n\nExpand and combine like terms, both sides are of the form\n\n\\[\nc_{n-1}x^{n-1}+c_{n-3}x^{n-3}+\\cdots+c_1x+c_{-1}x^{-1}+\\cdots+c_{-n+1}x^{-n+1}.\n\\]\n\nThey have the same values for infinitely many \\(x.\\) They must be identical. We just compare their leading terms. On the left hand side it is \\(2nx^{n-1}\\). There are two cases for the right hand sides: If \\(n>2\\), it is \\(x^{n-1}\\); If \\(n=2\\), it is \\((1+a_0)x.\\) It does not work for \\(n>2.\\) When \\(n=2,\\) we have \\(4=1+a_0.\\) therefore \\(a_0=3.\\)\n\nThe solution: \\(P(x)=c(x^2+3)\\) for any constant \\(c.\\)\n\n-JZ" ]
USAMO-2020-1
https://artofproblemsolving.com/wiki/index.php/2020_USAMO_Problems/Problem_1
Let \(ABC\) be a fixed acute triangle inscribed in a circle \(\omega\) with center \(O\). A variable point \(X\) is chosen on minor arc \(AB\) of \(\omega\), and segments \(CX\) and \(AB\) meet at \(D\). Denote by \(O_1\) and \(O_2\) the circumcenters of triangles \(ADX\) and \(BDX\), respectively. Determine all points \(X\) for which the area of triangle \(OO_1O_2\) is minimized.
[ "Let \\(E\\) be midpoint \\(AD.\\) Let \\(F\\) be midpoint \\(BD \\implies\\)\n\n\\[\nEF = ED + FD = \\frac {AD}{2} + \\frac {BD}{2} = \\frac {AB}{2}.\n\\]\n\n\\(E\\) and \\(F\\) are the bases of perpendiculars dropped from \\(O_1\\) and \\(O_2,\\) respectively.\n\nTherefore \\(O_1O_2 \\ge EF = \\frac {AB}{2}.\\)\n\n\\[\nCX \\perp O_1O_2, AX \\perp O_1O \\implies \\angle O O_1O_2 = \\angle AXC\n\\]\n\n\\(\\angle AXC = \\angle ABC (AXBC\\) is cyclic) \\(\\implies \\angle O O_1O_2 = \\angle ABC.\\)\n\nSimilarly \\(\\angle BAC = \\angle O O_2 O_1 \\implies \\triangle ABC \\sim \\triangle O_2 O_1O.\\)\n\nThe area of \\(\\triangle OO_1O_2\\) is minimized if \\(CX \\perp AB\\) because\n\n\\[\n\\frac {[OO_1O_2]} {[ABC]} = \\left(\\frac {O_1 O_2} {AB}\\right)^2 \\ge \\left(\\frac {EF} {AB}\\right)^2 = \\frac {1}{4}.\n\\]\n\nvladimir.shelomovskii@gmail.com, vvsss" ]
USAMO-2020-2
https://artofproblemsolving.com/wiki/index.php/2020_USAMO_Problems/Problem_2
An empty \(2020 \times 2020 \times 2020\) cube is given, and a \(2020 \times 2020\) grid of square unit cells is drawn on each of its six faces. A beam is a \(1 \times 1 \times 2020\) rectangular prism. Several beams are placed inside the cube subject to the following conditions: - The two \(1 \times 1\) faces of each beam coincide with unit cells lying on opposite faces of the cube. (Hence, there are \(3 \cdot 2020^2\) possible positions for a beam.) - No two beams have intersecting interiors. - The interiors of each of the four \(1 \times 2020\) faces of each beam touch either a face of the cube or the interior of the face of another beam. What is the smallest positive number of beams that can be placed to satisfy these conditions?
[ "Take one vertex of the cube as origin and establish 3D coordinates along the cube's edges.\n\nDefine a beam as \\(x-dir\\) if its long edge is parallel to x-axis. Similarly for \\(y-dir\\) and \\(z-dir\\).\n\nDefine a beam's location as (direction, (\\(1 \\times 1\\) face's location in 2D coordinate).\n\nFor example, (y, 2, 4) indicates the beam with vertex (1, 0, 3) and (2, 2020, 4)\n\nApparently \\(x\\) beam needs the other \\(x\\) or \\(y\\) beams to touch its \\(xy\\) faces, \\(x\\) or \\(z\\) beams to touch its \\(xz\\) faces. Similarly for \\(y\\) and \\(z\\) beam.\n\nIf there are only 1-dir or 2-dirs beams, it is easy to approve that \\(2020 \\times 2020\\) is the minimal number beams.\n\n(for example, if only use \\(x-dir\\) and \\(y-dir\\) beams, all the \\(x-dir\\) beams's xz faces can only be touched by \\(x-dir\\) beams. In the other word, \\(2020 x-dir\\) beams will be stacked to meet xz faces touch requirements in that xz layer)\n\nConsider cases with all \\(3-dirs\\) beams. WLOG there is a \\(x-dir\\) beam and it needs \\(x-dir\\) or \\(y-dir\\) beams to touch its \\(xy\\) faces (unless it touches the cube surface).\n\nAnd this \\(y-dir\\) beam also needs a \\(x-dir\\) or \\(y-dir\\) to touch it's \\(xy\\) faces. And so on until one which touches cube surface. So from \\(xy\\) face perspective, it needs \\(2020\\) beams.\n\nSimilarly from \\(xz\\) and \\(yz\\) face perspective, it also needs \\(2020\\) and \\(2020\\) beams.\n\nConsider one beam has four \\(1 \\times 2020\\) faces and it can be counted twice. So there should be at least \\(2020 \\times 3 \\div 2=3030\\) beams.\n\nHere is one solution with 3030 beams.\n\n\\[\n(x, 1, 1),\\ (y, 1, 2),\\ (z, 2, 2),\n\\]\n\n\\[\n\\cdots ,\n\\]\n\n\\[\n(x, (2n+1), (2n+1)),\\ (y, (2n+1), (2n+2)),\\ (z, (2n+2), (2n+2)),\n\\]\n\n\\[\n\\cdots ,\n\\]\n\n\\[\n(x, (2019, 2019)),\\ (y, 2019, 2020),\\ (z, 2020, 2020)\n\\]\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2020-3
https://artofproblemsolving.com/wiki/index.php/2020_USAMO_Problems/Problem_3
Let \(p\) be an odd prime. An integer \(x\) is called a quadratic non-residue if \(p\) does not divide \(x - t^2\) for any integer \(t\). Denote by \(A\) the set of all integers \(a\) such that \(1 \le a < p\), and both \(a\) and \(4 - a\) are quadratic non-residues. Calculate the remainder when the product of the elements of \(A\) is divided by \(p\).
[ "We will be using finite field theory and the Frobenius endomorphism freely. Also we will just give a sketch and leave out details like it's out of fashion.\n\n[b]Case 1[/b]: \\(p\\equiv 3\\pmod{4}\\). Work in \\(\\mathbb{F}_{p^2} = \\mathbb{F}_p[i]\\). Suppose \\(a\\in A\\). Then, \\(a=-x^2\\) for some \\(x\\in\\mathbb{F}_p\\) and \\(4-a=y^2\\) for some \\(y\\in\\mathbb{F}_p\\), so we have\n\n\\[\nx^2+y^2 = -4.\n\\]\n\nThis factors as\n\n\\[\n(x+iy)(x-iy) = (2i)^2,\n\\]\n\nso we have \\(x+iy=2it\\) and \\(x-iy=2i/t\\) for some \\(t\\in\\mathbb{F}_p[i]\\), so\n\n\\[\nx=i(t+1/t)\\quad\\text{and}\\quad y=t-1/t.\n\\]\n\nFor both of these to be in \\(\\mathbb{F}_p\\), we need \\(x^p=x\\) and \\(y^p=y\\), which when we work out with frobenius, tells us that \\(t^{p+1}+1=0\\).\n\nThus, if \\(a\\in A\\), we must have \\(a=(t+1/t)^2\\) for some \\(t\\in\\mathbb{F}_p[i]\\) such that \\(t^{p+1}+1=0\\). Actually it is easy to see that if \\(t\\) satisfies \\(t^{p+1}+1=0\\), then both \\(a\\) and \\(4-a\\) are QNRs (just reverse the frobenius calculation above). Thus,\n\n\\[\nA = \\{(t+1/t)^2 : t\\in\\mathbb{F}_p[i], t^{p+1}+1=0\\}.\n\\]\n\nNote \\((t+1/t)^2 = (t'+1/t')^2\\) if and only if \\(t'\\in\\{t,-t,1/t,-1/t\\}\\), and also \\(t^{p+1}+1=0\\iff (t')^{p+1}+1=0\\). Note that\n\n\\[\nt^{p+1}+1\\mid t^{p^2}-t\n\\]\n\nas \\(2(p+1)\\mid p^2-1\\), so \\(t^{p+1}+1\\) has distinct roots in \\(\\mathbb{F}_p[i]\\). We see that\n\n\\[\nt^{p+1}+1=(t^{\\frac{p+1}{2}}+i)(t^{\\frac{p+1}{2}}-i),\n\\]\n\nand \\(s\\) a root of \\(t^{\\frac{p+1}{2}}-i\\) if and only if \\(1/s\\) is a root. Thus, it is easy to see that\n\n\\[\nA = \\{(t+1/t)^2 : t\\in\\mathbb{F}_p[i], t^{\\frac{p+1}{2}}-i=0\\}.\n\\]\n\nIn this parameterization, only \\(t\\) and \\(-t\\) produce the same value (and these never coincide as \\(t\\ne 0\\)), so we actually have\n\n\\[\nT:=\\prod_{a\\in A}=(-1)^{\\frac{p+1}{4}}\\prod_{\\substack{t\\in\\mathbb{F}_p[i]\\\\t^{\\frac{p+1}{2}}-i=0}}(t+1/t).\n\\]\n\nThis becomes\n\n\\[\nT=(-1)^{\\frac{p+1}{4}}\\prod_{\\substack{t\\in\\mathbb{F}_p[i]\\\\t^{\\frac{p+1}{2}}-i=0}}\\frac{(t+i)(t-i)}{t}.\n\\]\n\nLetting \\(f(x)=x^{\\frac{p+1}{2}}-i\\), we see that this product becomes\n\n\\[\nT=(-1)^{\\frac{p+1}{4}}\\frac{f(-i)f(i)}{f(0)}.\n\\]\n\nIt is tedious but straightforward to see that this always gives \\(2\\).\n\n[b]Case 2[/b]: \\(p\\equiv 1\\pmod{4}\\). Let \\(\\ell\\) be some QNR mod \\(p\\) (sadly we don't have any free QNRs like \\(-1\\) is for \\(3\\pmod{4}\\) primes). Let \\(\\omega\\) be a formal square root of \\(\\ell\\), and work in \\(\\mathbb{F}_{p^2} = \\mathbb{F}_p[\\omega]\\). Then, if \\(a\\in A\\), we have\n\n\\[\n\\omega^2 a = x^2\\quad\\text{and}\\quad\\omega^2(4-a)=y^2,\n\\]\n\nso \\(x^2+y^2=4\\omega^2\\), so factoring and doing the same as before, we get\n\n\\[\nx=\\omega(t+1/t)\\quad\\text{and}\\quad y=\\frac{\\omega}{i}(t-1/t)\n\\]\n\nfor some \\(t\\in\\mathbb{F}_p[\\omega]\\). These are in \\(\\mathbb{F}_p\\) if and only if \\(t^{p-1}+1=0\\) (using frobenius). We see \\(a=(t+1/t)^2\\), so\n\n\\[\nA = \\{(t+1/t)^2 : t\\in\\mathbb{F}_p[\\omega], t^{p-1}+1=0\\}.\n\\]\n\nDoing a similar analysis as before, the desired product becomes\n\n\\[\nT:=\\prod_{a\\in A}=(-1)^{\\frac{p-1}{4}}\\prod_{\\substack{t\\in\\mathbb{F}_p[\\omega]\\\\t^{\\frac{p-1}{2}}-i=0}}\\frac{(t+i)(t-i)}{t},\n\\]\n\nand letting \\(f(x)=x^{\\frac{p-1}{2}}-i\\), we get\n\n\\[\nT=(-1)^{\\frac{p-1}{4}}\\frac{f(-i)f(i)}{f(0)},\n\\]\n\nwhich again is always \\(2\\).\n\nSo the answer is \\(\\boxed{2}\\). These problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2020-4
https://artofproblemsolving.com/wiki/index.php/2020_USAMO_Problems/Problem_4
Suppose that \((a_1, b_1), (a_2, b_2), \ldots , (a_{100}, b_{100})\) are distinct ordered pairs of nonnegative integers. Let \(N\) denote the number of pairs of integers \((i, j)\) satisfying \(1 \le i < j \le 100\) and \(|a_ib_j - a_j b_i|=1\). Determine the largest possible value of \(N\) over all possible choices of the \(100\) ordered pairs.
[ "Let's start off with just \\((a_1, b_1), (a_2, b_2)\\) and suppose that it satisfies the given condition. We could use \\((1, 1), (1, 2)\\) for example. We should maximize the number of conditions that the third pair satisfies. We find out that the third pair should equal \\((a_1+a_2, b_1+b_2)\\):\n\nWe know this must be true:\n\n\\[\n|a_1b_2-a_2b_1| = 1\n\\]\n\nSo\n\n\\[\na_1b_2-a_2b_1 = 1\n\\]\n\nWe require the maximum conditions for \\((a_3, b_3)\\)\n\n\\[\n|a_3b_2-a_2b_3| = 1\n\\]\n\n\\[\n|a_3b_1-a_1b_3| = 1\n\\]\n\nThen one case can be:\n\n\\[\na_3b_2-a_2b_3 = 1\n\\]\n\n\\[\na_3b_1-a_1b_3 = -1\n\\]\n\nWe try to do some stuff such as solving for \\(a_3\\) with manipulations:\n\n\\[\na_3b_2a_1-a_2b_3a_1 = a_1\n\\]\n\n\\[\na_3b_1a_2-a_1b_3a_2 = -a_2\n\\]\n\n\\[\na_3(a_1b_2-a_2b_1) = a_1+a_2\n\\]\n\n\\[\na_3 = a_1+a_2\n\\]\n\n\\[\na_3b_2b_1-a_2b_3b_1 = b_1\n\\]\n\n\\[\na_3b_1b_2-a_1b_3b_2 = -b_2\n\\]\n\n\\[\nb_3(a_1b_2-a_2b_1) = b_1+b_2\n\\]\n\n\\[\nb_3 = b_1+b_2\n\\]\n\nWe showed that 3 pairs are a complete graph; however, 4 pairs are not a complete graph. We will now show that:\n\n\\[\na_4 = a_1+2a_2\n\\]\n\n\\[\nb_4 = b_1+2b_2\n\\]\n\n\\[\n|a_1b_1+2a_2b_1-a_1b_1-2a_1b_2| = 1\n\\]\n\n\\[\n2|a_2b_1-a_1b_2| = 1\n\\]\n\nThis is clearly impossible because \\(1\\) is not even and also \\(|a_2b_1-a_1b_2| = 1\\). The answer is as follows:\n\n\\[\n0+1+2+\\ldots+2\n\\]\n\n\\(a_1\\) has \\(0\\) subtractions that follow condition while \\(a_2\\) has \\(1\\) and then the rest has \\(2\\). There are \\(n\\) terms, so our answer be \\(2n-3\\) and in case of \\(n=100\\) that means\n\n\\[\n\\boxed{N=197}.\n\\]\n\n~Lopkiloinm", "We claim the answer is \\(197\\).\n\nStudy the points \\((0, 0), (a_i, b_i), (a_j, b_j)\\). If we let these be the vertices of a triangle, applying shoelace theorem gives us an area of \\(\\frac{1}{2}|0\\times{b_i}+{a_i}\\times{b_j}+{b_i}\\times{0}-0\\times{a_i}-{b_i}\\times{a_j}-{b_j}\\times{0} = \\frac{1}{2}|a_ib_j - a_j b_i| = \\frac{1}{2}\\). Therefore, the triangle formed by the points \\((0, 0), (a_i, b_i), (a_j, b_j)\\) must have an area of \\(\\frac{1}{2}\\).\n\nTwo cases follow. Case 1: Both \\((a_i, b_i), (a_j, b_j)\\) have exactly one coordinate equal to \\(0\\). Here, one point must be on the \\(x\\) axis and the other on the \\(y\\) axis in order for the triangle to have a positive area. For the area of the triangle to be \\(\\frac{1}{2}\\), it follows that the points must be \\((1, 0), (0, 1)\\) in some order.\n\nCase 2: At least one of \\((a_i, b_i), (a_j, b_j)\\) does not have exactly one coordinate equal to \\(0\\). Define \\(S[l]\\) to be a list of lines such that each line in the list has some two lattice points that, with \\((0, 0)\\), form a triangle with area \\(\\frac{1}{2}\\). Note that for any such line that passes through such two lattice points, we may trivially generate infinite lattice points on the line that have nonnegative coordinates.\n\nNote that lines \\(y=1\\) and \\(x=1\\) are included in \\(S[l]\\), because the points \\((1, 1), (2, 1)\\) serve as examples for \\(y=1\\) and \\((1, 1), (1, 2)\\) serve as examples for \\(x=1\\). For the optimal construction, include the points \\((1, 0)\\) and all the points \\((0, 1), (0, 2), (0, 3), ... , (0, 99)\\), in that order. In this case, every adjacent pair of points would count (\\(98\\)), as well as picking \\((0, 1)\\) and a nonadjacent point (\\(99\\)), so this would be \\(98+99=197\\).\n\nTo prove that this is the maximum, consider the case where some \\(n\\) number of points were neither on \\(x=1\\) nor on \\(y=1\\). In this case, we would be removing \\(n\\) adjacent pairs and \\(n\\) options to choose from after choosing \\((0, 0)\\), resulting in a net loss of \\(2n\\). By having \\(n\\) points on some other combination of lines in \\(S[l]\\), we would trivially have a maximum gain of \\(n-1\\) pairs of points on the lines such that there are no lattice points between those pairs. Because these points are not on \\(x=1\\) or \\(y=1\\), the altitude from a given point to the line formed by \\((0, 0)\\) and \\((0, 1)\\) and \\((1, 0)\\) is not \\(1\\), and so the area of the triangle cannot be \\(\\frac{1}{2}\\). Thus, by not having all points on lines \\(x=1\\) and \\(y=1\\), we cannot exceed the maximum of \\(197\\). Thus, \\(\\boxed{197}\\) is our answer.\n\n~SigmaPiE" ]
USAMO-2020-5
https://artofproblemsolving.com/wiki/index.php/2020_USAMO_Problems/Problem_5
A finite set \(S\) of points in the coordinate plane is called overdetermined if \(|S| \ge 2\) and there exists a nonzero polynomial \(P(t)\), with real coefficients and of degree at most \(|S| - 2\), satisfying \(P(x) = y\) for every point \((x, y) \in S\). For each integer \(n \ge 2\), find the largest integer \(k\) (in terms of \(n\)) such that there exists a set of \(n\) distinct points that is not overdetermined, but has \(k\) overdetermined subsets.
[ "The answer is \\(2^{n-1}-n\\). To construct this, have \\(n-1\\) points on a horizontal line and \\(1\\) point not on it. Then, any subset that does not include the last point is overdetermined.\\\\\n\nFor the bound, the main idea of the problem is the following claim.\\\\\n\nClaim: If \\(S\\) and \\(T\\) are overdetermined subsets with \\(|S|=|T|=k\\) but \\(S\\) and \\(T\\) only differ by one element (one is swapped out for another), then \\(S\\cup T\\) is overdetermined.\\\\\n\nConsider the minimal polynomial of \\(S\\). It has degree at most \\(k-2\\) since \\(S\\) is overdetermined, and similarly with the minimal polynomial of \\(T\\). However, there is a unique polynomial of degree at most \\(k-2\\) passing through \\(S\\cap T\\) since it has \\(k-1\\) points. Thus, this unique polynomial also passes through \\(S\\) and \\(T\\). Thus, this polynomial passes through all points of \\(S\\cup T\\), as desired.\\\\\n\nWe now use this claim to inductively bound the number of overdetermined subsets with \\(k\\) elements. Let \\(m_k\\) denote the number of overdetermined subsets of size \\(k\\). Clearly, \\(m_n=0\\).\\\\\n\nClaim 2: We have\n\n\\[\nm_k\\leq \\frac{m_{k+1}(k+1)+{n\\choose k+1}-m_{k+1}}{n-k}.\n\\]\n\nProof: Each overdetermined subset of size \\(k+1\\) can have up to \\(k+1\\) overdetermined subsets of size \\(k\\), but by the previous claim each non-overdetermined subset of size \\(k+1\\), which there are \\({n\\choose k+1}-m_{k+1}\\) of, can have at most one. Each subset of size \\(k\\) feeds into \\(n-k\\) subsets of size \\(k+1\\). Note that this is increasing in \\(m_{k+1}\\).\\\\\n\nClaim 3: We have\n\n\\[\nm_k\\leq {n-1\\choose k}.\n\\]\n\nProof: Use induction. Note that the expression in Claim 2 is nondecreasing in \\(m_{k+1}\\), and verify that\n\n\\[\n\\frac{{n-1\\choose k+1}(k+1)+{n\\choose k+1}-{n-1\\choose k+1}}{n-k}={n-1\\choose k}.\n\\]\n\nThus, we have that the number of overdetermined subsets is at most\n\n\\[\n\\sum_{k=2}^n {n-1\\choose k}=2^{n-1}-n,\n\\]\n\nas desired.\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2020-6
https://artofproblemsolving.com/wiki/index.php/2020_USAMO_Problems/Problem_6
Let \(n \ge 2\) be an integer. Let \(x_1 \ge x_2 \ge \cdots \ge x_n\) and \(y_1 \ge y_2 \ge \cdots \ge y_n\) be \(2n\) real numbers such that \begin{align*} 0 &= x_1 + x_2 + \cdots + x_n = y_1 + y_2 + \cdots + y_n \\ \text{and }1 &= x_1^2+x_2^2+\cdots+x_n^2=y_1^2+y_2^2+\cdots+y_n^2. \end{align*} Prove that \[ \sum_{i=1}^n(x_iy_i-x_iy_{n+1-i})\ge\frac{2}{\sqrt{n-1}}. \]
[]
USAMO-2021-1
https://artofproblemsolving.com/wiki/index.php/2021_USAMO_Problems/Problem_1
Rectangles \(BCC_1B_2,\) \(CAA_1C_2,\) and \(ABB_1A_2\) are erected outside an acute triangle \(ABC.\) Suppose that \[ \angle BC_1C+\angle CA_1A+\angle AB_1B=180^{\circ}. \] Prove that lines \(B_1C_2,\) \(C_1A_2,\) and \(A_1B_2\) are concurrent.
[ "Let \\(D\\) be the second point of intersection of the circles \\(AB_1B\\) and \\(AA_1C.\\) Then:\n\n\\[\n\\begin{align*} \\angle ADB &= 180^\\circ – \\angle AB_1B,&\\angle ADC &= 180^\\circ – \\angle AA_1C\\\\ \\angle BDC &= 360^\\circ – \\angle ADB – \\angle ADC\\\\ &= 360^\\circ – (180^\\circ – \\angle AB_1B) – (180^\\circ – \\angle AA_1C)\\\\ &= \\angle AB_1B + \\angle AA_1C\\\\ \\angle BDC + \\angle BC_1C &= 180^\\circ \\end{align*}\n\\]\n\nTherefore, \\(BDCC_1B_2\\) is cyclic with diameters \\(BC_1\\) and \\(CB_2\\), and thus \\(\\angle CDB_2 = 90^\\circ.\\) Similarly, \\(\\angle CDA_1 = 90^\\circ\\), meaning points \\(A_1\\), \\(D\\), and \\(B_2\\) are collinear.\n\nSimilarly, the points \\(A_2, D, C_1\\) and \\(C_2, D, B_1\\) are collinear.\n\n(After USAMO 2021 Solution Notes – Evan Chen)\n\nvladimir.shelomovskii@gmail.com, vvsss" ]
USAMO-2021-2
https://artofproblemsolving.com/wiki/index.php/2021_USAMO_Problems/Problem_2
The Planar National Park is an undirected 3-regular planar graph (i.e. all vertices have degree 3). That is: every trail has its two endpoints at two different junctions whereas each junction is the endpoint of exactly three trails. Trails only intersect at junctions (in particular, trails only meet at endpoints). Finally, no trails begin and end at the same two junctions. (An example of one possible layout of the park is shown to the left below, in which there are six junctions and nine trails.) <image> A visitor walks through the park as follows: she begins at a vertex and starts walking along an edge. When she reaches the other endpoint, she turns left. On the next vertex, she turns right, and so on, alternating left and right turns at each vertex. She does this until she gets back to the vertex where she started. What is the largest possible number of times she could have entered any vertex during her walk, over all possible layouts of the park?
[]
USAMO-2021-3
https://artofproblemsolving.com/wiki/index.php/2021_USAMO_Problems/Problem_3
Let \(n \geq 2\) be an integer. An \(n \times n\) board is initially empty. Each minute, you may perform one of three moves: - If there is an L-shaped tromino region of three cells without stones on the board (see figure; rotations not allowed), you may place a stone in each of those cells. - If all cells in a column have a stone, you may remove all stones from that column. - If all cells in a row have a stone, you may remove all stones from that row. \[ [asy] unitsize(20); draw((0,0)--(4,0)--(4,4)--(0,4)--(0,0)); fill((0.2,3.8)--(1.8,3.8)--(1.8, 1.8)--(3.8, 1.8)--(3.8, 0.2)--(0.2, 0.2)--cycle, grey); draw((0.2,3.8)--(1.8,3.8)--(1.8, 1.8)--(3.8, 1.8)--(3.8, 0.2)--(0.2, 0.2)--(0.2, 3.8), linewidth(2)); draw((0,2)--(4,2)); draw((2,4)--(2,0)); [/asy] \] For which \(n\) is it possible that, after some non-zero number of moves, the board has no stones?
[ "Label the bottom right cell as \\(\\omega^0\\) where \\(\\omega = e^{i2\\pi/3}\\). Then label each cell with \\(\\omega^d\\) where \\(d\\) is the Manhattan distance of the current cell from the bottom right cell.\n\nNo matter where an L-shaped tromino is placed, the sum of the labels in the cells is,\n\n\\[\n\\omega^0 + \\omega^1 + \\omega^2 = \\frac{\\omega^3 - 1}{\\omega - 1} = \\frac{0}{\\omega - 1} = 0\n\\]\n\nIf \\(n = 3k + 1\\). Then the sum of all the labels in the cells in a row or column is,\n\n\\[\nk(\\omega^0 + \\omega^1 + \\omega^2)+\\omega^0 = 1\n\\]\n\nIf \\(n = 3k + 2\\). Then the sum of all the labels in cells in a row or column is,\n\n\\[\nk(\\omega^0 + \\omega^1 + \\omega^2) + \\omega^0 + \\omega^1 = 1 + \\omega\n\\]\n\nLet \\(S\\) be the sum of all the labels of cells containing a stone. In either above case, placing an L-shaped Tromino leaves \\(S\\) invariant but removing a row or column of stone causes \\(\\textrm{Re}(S)\\) to decrease (by \\(1\\) or \\(1/2\\) in each case). However the empty grid has \\(S=0\\), so this can't be reached.\n\nHence \\(n=3k\\). We now provide a sequence of moves which empties the grid. Divide the grid into \\(3\\times 3\\) grids and do the following to every sub-grid. (When I refer to placing an L-shaped tromino on a cell, I am refering to the position of the bottom left stone)\n\n- Add an L-shaped tromino to the bottom left cell and middle cell\n- Remove the middle row\n- Add an L-shaped tromino to the middle left cell\n- Remove the left and middle columns\n\nand we're done, the grid is empty" ]
USAMO-2021-4
https://artofproblemsolving.com/wiki/index.php/2021_USAMO_Problems/Problem_4
A finite set \(S\) of positive integers has the property that, for each \(s\in S\), and each positive integer divisor \(d\) of \(s\), there exists a unique element \(t \in S\) satisfying \(\gcd(s,t)=d\) (the elements \(s\) and \(t\) could be equal). Given this information, find all possible values for the number of elements of \(S\).
[]
USAMO-2021-5
https://artofproblemsolving.com/wiki/index.php/2021_USAMO_Problems/Problem_5
Let \(n \geq 4\) be an integer. Find all positive real solutions to the following system of \(2n\) equations: \begin{align*} a_{1} &=\frac{1}{a_{2 n}}+\frac{1}{a_{2}}, & a_{2}&=a_{1}+a_{3}, \\ a_{3}&=\frac{1}{a_{2}}+\frac{1}{a_{4}}, & a_{4}&=a_{3}+a_{5}, \\ a_{5}&=\frac{1}{a_{4}}+\frac{1}{a_{6}}, & a_{6}&=a_{5}+a_{7} \\ &\vdots & &\vdots \\ a_{2 n-1}&=\frac{1}{a_{2 n-2}}+\frac{1}{a_{2 n}}, & a_{2 n}&=a_{2 n-1}+a_{1} \end{align*}
[]
USAMO-2021-6
https://artofproblemsolving.com/wiki/index.php/2021_USAMO_Problems/Problem_6
Let \(ABCDEF\) be a convex hexagon satisfying \(\overline{AB} \parallel \overline{DE}\), \(\overline{BC} \parallel \overline{EF}\), \(\overline{CD} \parallel \overline{FA}\), and \[ AB \cdot DE = BC \cdot EF = CD \cdot FA. \] Let \(X\), \(Y\), and \(Z\) be the midpoints of \(\overline{AD}\), \(\overline{BE}\), and \(\overline{CF}\). Prove that the circumcenter of \(\triangle ACE\), the circumcenter of \(\triangle BDF\), and the orthocenter of \(\triangle XYZ\) are collinear.
[ "Let \\(M_1\\), \\(M_2\\), and \\(M_3\\) be the midpoints of \\(CE\\), \\(AE\\), \\(AC\\) and \\(N_1\\), \\(N_2\\), and \\(N_3\\) be the midpoints of \\(DF\\), \\(BF\\), and \\(BD\\). Also, let \\(H\\) be the orthocenter of \\(XYZ\\). Note that we can use parallel sides to see that \\(X\\), \\(Z\\), and \\(M_3\\) are collinear. Thus we have\n\n\\[\n\\text{Pow}(M_3,(XYZ)) = M_3Z \\cdot M_3X = \\frac 14 AB \\cdot DE\n\\]\n\nby midlines. Applying this argument cyclically, and noting the condition \\(AB \\cdot DE = BC \\cdot EF = CD \\cdot FA\\), \\(M_1\\), \\(M_2\\), \\(M_3\\), \\(N_1\\), \\(N_2\\), \\(N_3\\) all lie on a circle concentric with \\((XYZ)\\).\n\nNext, realize that basic orthocenter properties imply that the circumcenter \\(O_1\\) of \\((ACE)\\) is the orthocenter of \\(\\triangle M_1M_2M_3\\), and likewise the circumcenter \\(O_2\\) of \\((BDF)\\) is the orthocenter of \\(\\triangle N_1N_2N_3\\).\n\nThe rest is just complex numbers; toss on the complex plane so that the circumcenter of \\(\\triangle XYZ\\) is the origin. Then we have\n\n\\[\no_1 = m_1+m_2+m_3 = (c+e)/2+(a+e)/2+(a+c)/2=a+c+e\n\\]\n\n\\[\no_2 = n_1+n_2+n_3 = (b+d)/2+(d+f)/2+(b+f)/2=b+d+f\n\\]\n\n\\[\nh = x+y+z = (a+d)/2+(b+e)/2+(c+f)/2=(a+b+c+d+e+f)/2.\n\\]\n\nNote that from the above we have \\(h=\\frac{o_1+o_2}{2}\\), so \\(H\\) is the midpoint of segment \\(O_1O_2\\). In particular, \\(H\\), \\(O_1\\), and \\(O_2\\) are collinear, as required.\n\n~ Leo.Euler", "We construct two equal triangles, prove that triangle \\(XYZ\\) is the same as medial triangle of both this triangles. We use property of medial triangle and prove that circumcenters of constructed triangles coincide with given circumcenters.\n\nDenote \\(A' = C + E – D, B' = D + F – E, C' = A+ E – F,\\) \\(D' = F+ B – A, E' = A + C – B, F' = B+ D – C.\\) Then \\(A' – D' = C + E – D – ( F+ B – A) = (A + C + E ) – (B+ D + F).\\)\n\nDenote \\(D' – A' = 2\\vec V.\\)\n\nSimilarly we get \\(B' – E' = F' – C' = D' – A' \\implies\\) \\(\\triangle A'C'E' = \\triangle D'F'B'.\\)\n\nThe translation vector maps \\(\\triangle A'C'E'\\) into \\(\\triangle D'F'B'\\) is \\(2\\vec {V.}\\) \\(X = \\frac {A+D}{2} = \\frac { (A+ E – F) + (D + F – E)}{2} = \\frac {C' + B'}{2} = \\frac {E' + F'}{2},\\)\n\nso \\(X\\) is midpoint of \\(AD, B'C',\\) and \\(E'F'.\\) Symilarly \\(Y\\) is the midpoint of \\(BE, A'F',\\) and \\(C'D', Z\\) is the midpoint of \\(CF, A'B',\\) and \\(D'E'.\\) \\(Z + V = \\frac {A' + B'}{2}+ \\frac {D' – A'}{2} = \\frac {B' + D'}{2} = Z'\\) is the midpoint of \\(B'D'.\\)\n\nSimilarly \\(X' = X + V\\) is the midpoint of \\(B'F',Y'= Y + V\\) is the midpoint of \\(D'F'.\\)\n\nTherefore \\(\\triangle X'Y'Z'\\) is the medial triangle of \\(\\triangle B'D'F'.\\)\n\n\\(\\triangle XYZ\\) is \\(\\triangle X'Y'Z'\\) translated on \\(– \\vec {V}.\\)\n\nIt is known (see diagram) that circumcenter of triangle coincide with orthocenter of the medial triangle. Therefore orthocenter \\(H\\) of \\(\\triangle XYZ\\) is circumcenter of \\(\\triangle B'D'F'\\) translated on \\(– \\vec {V}.\\)\n\nIt is the midpoint of segment \\(OO'\\) connected circumcenters of \\(\\triangle B'D'F'\\) and \\(\\triangle A'C'E'.\\)\n\nAccording to the definition of points \\(A', C', E',\\) quadrangles \\(ABCE', CDEA',\\) and \\(AFEC'\\) are parallelograms. Hence\n\n\\[\nAC' = FE, AE' = BC, CE' = AB, CA' = DE, EA' = CD, EC' = AF \\implies\n\\]\n\n\\[\nAC' \\cdot AE' = CE' \\cdot CA' = EA' \\cdot EC' = AB \\cdot DE \\implies\n\\]\n\nPower of points A,C, and E with respect circumcircle \\(\\triangle A'C'E'\\) is equal, hence distances between these points and circumcenter of \\(\\triangle A'C'E'\\) are the same. Therefore circumcenter \\(\\triangle ACE\\) coincide with circumcenter \\(\\triangle A'C'E'.\\)\n\nSimilarly circumcenter of \\(\\triangle BDF\\) coincide with circumcenter of \\(\\triangle B'D'F'.\\)\n\nvladimir.shelomovskii@gmail.com, vvsss\n\nThese problems are copyrighted © by the Mathematical Association of America, as part of the American Mathematics Competitions." ]
USAMO-2022-1
https://artofproblemsolving.com/wiki/index.php/2022_USAMO_Problems/Problem_1
Let \(a\) and \(b\) be positive integers. The cells of an \((a+b+1)\times (a+b+1)\) grid are colored amber and bronze such that there are at least \(a^2+ab-b\) amber cells and at least \(b^2+ab-a\) bronze cells. Prove that it is possible to choose \(a\) amber cells and \(b\) bronze cells such that no two of the \(a+b\) chosen cells lie in the same row or column.
[ "We proceed with induction. Base case: \\(a=b=1\\). We fill in a 3x3 grid that has at least one bronze and one amber cell. We choose one bronze and one amber cell. Note that if the given bronze and amber cells are not in the same row or column, we are done. However, if they are, consider a cell that is not in the same row or column as either cell. Regardless of the color of the cell, we can match it with an opposing color cell, hence proved.\n\nInduction case: Given the assertion for \\((a, b)\\) is true, then \\((a+1, b)\\) is true as well. Label each cell in the \\((a+b+1)(a+b+1)\\) grid as either red, blue, or white. There must be exactly \\(a^2+ab-b\\) red cells, \\(b^2+ab-a\\) blue cells, and \\((a+b+1)(a+b+1)-(a^2+ab-b)-(b^2+ab-a) = 3(a+b) + 1\\) white cells. We show that we can achieve a combination regardless of the amber/bronze of the white cells.\n\nCase 1: There are three or more white cells present in either the bottommost row or rightmost column. Let's ignore the bottom most row and rightmost column, and focus on the \\((a+b+1)(a+b+1)\\) grid of squares on the topleft. By the induction assumption, we know that this grid must have \\(a\\) amber and \\(b\\) bronze not in the same row or column pairwise. [cont]" ]
USAMO-2022-2
https://artofproblemsolving.com/wiki/index.php/2022_USAMO_Problems/Problem_2
Let \(b\geq2\) and \(w\geq2\) be fixed integers, and \(n=b+w\). Given are \(2b\) identical black rods and \(2w\) identical white rods, each of side length \(1\). We assemble a regular \(2n\)-gon using these rods so that parallel sides are the same color. Then, a convex \(2b\)-gon \(B\) is formed by translating the black rods, and a convex \(2w\)-gon \(W\) is formed by translating the white rods. An example of one way of doing the assembly when \(b=3\) and \(w=2\) is shown below, as well as the resulting polygons \(B\) and \(W\). \[ [asy] size(10cm); real w = 2*Sin(18); real h = 0.10 * w; real d = 0.33 * h; picture wht; picture blk; draw(wht, (0,0)--(w,0)--(w+d,h)--(-d,h)--cycle); fill(blk, (0,0)--(w,0)--(w+d,h)--(-d,h)--cycle, black); // draw(unitcircle, blue+dotted); // Original polygon add(shift(dir(108))*blk); add(shift(dir(72))*rotate(324)*blk); add(shift(dir(36))*rotate(288)*wht); add(shift(dir(0))*rotate(252)*blk); add(shift(dir(324))*rotate(216)*wht); add(shift(dir(288))*rotate(180)*blk); add(shift(dir(252))*rotate(144)*blk); add(shift(dir(216))*rotate(108)*wht); add(shift(dir(180))*rotate(72)*blk); add(shift(dir(144))*rotate(36)*wht); // White shifted real Wk = 1.2; pair W1 = (1.8,0.1); pair W2 = W1 + w*dir(36); pair W3 = W2 + w*dir(108); pair W4 = W3 + w*dir(216); path Wgon = W1--W2--W3--W4--cycle; draw(Wgon); pair WO = (W1+W3)/2; transform Wt = shift(WO)*scale(Wk)*shift(-WO); draw(Wt * Wgon); label("$W$", WO); /* draw(W1--Wt*W1); draw(W2--Wt*W2); draw(W3--Wt*W3); draw(W4--Wt*W4); */ // Black shifted real Bk = 1.10; pair B1 = (1.5,-0.1); pair B2 = B1 + w*dir(0); pair B3 = B2 + w*dir(324); pair B4 = B3 + w*dir(252); pair B5 = B4 + w*dir(180); pair B6 = B5 + w*dir(144); path Bgon = B1--B2--B3--B4--B5--B6--cycle; pair BO = (B1+B4)/2; transform Bt = shift(BO)*scale(Bk)*shift(-BO); fill(Bt * Bgon, black); fill(Bgon, white); label("$B$", BO); [/asy] \] Prove that the difference of the areas of \(B\) and \(W\) depends only on the numbers \(b\) and \(w\), and not on how the \(2n\)-gon was assembled.
[ "First notice that the black rods and the white rods form polygons iff in the original \\(2n\\)-gon, if a side is a color \\(x\\), then the side that is parallel to that side in the original \\(2n\\)-gon is also the color \\(x\\).\n\nWe can prove that the difference in areas is only affected by the values of \\(b\\) and \\(w\\) by showing that for any valid arrangement of \\(2b\\) rods and \\(2w\\) rods, we may switch any two adjacent black and white rods(and their \"parallel pairs\"), and end up with the same area difference.\n\n\\[\nPeppa-usamo-2022-2.png\n\\]\n\nIn the figure above (click to expand), after the switch, we can see that after removing the mutually congruent parts, we are left with two parallelograms from each color. Let \\(x\\), \\(\\alpha{}\\), \\(y\\), and \\(\\beta{}\\) be defined as shown. Notice that if we angle chase, the sides of the other parallelogram are the same, but if the angles of the original \\(2n\\)-gon all have measure \\(2k\\), the angles of the new parallelograms are \\(\\alpha{}+180-2k\\) and \\(\\beta{}+180-2k\\), as shown. We must prove that the differences between the areas are the same.\n\nUsing area formulas, the change in the difference of areas is \\(x\\sin{\\alpha{}}-y\\sin{\\beta{}}+y\\sin{(2k-\\beta{})}-x\\sin{(2k-\\alpha{})}\\), which is equal to \\(x(\\sin{k}(2\\cos^2{k})-2\\sin{k}\\cos{k}\\cos{\\alpha{}})-y(\\sin{k}(2\\cos^2{k}))+y(\\sin{k}\\cos{k}\\cos{\\beta{}})\\), or \\(2\\cos{k}(x\\sin{(\\alpha{}-k)}+y\\sin{(k-\\beta{}}))\\). Since \\(\\cos{k}\\) is not \\(0\\) because \\(n\\geq{}3\\), we are left with proving that \\(x\\sin{(\\alpha{}-k)}+y\\sin{(k-\\beta{}})=0\\).\n\nNow we rotate the polygon so that the vertex between the two sides that we switched is at the point \\((0,0)\\), the angle bisector of that vertex is \\(y=0\\), and the black side is in the positive \\(y\\)-direction. Now think of all the sides as vectors, all pointing in the clockwise direction of the \\(2n\\)-gon.\n\nNotice the part labeled \\(a\\) in the black polygons. We have that the vector labeled \\(x\\) is really just the sum of all of the vectors in the part labeled \\(a\\) - or all the vectors in the \\(2n\\)-gon that are in the positive \\(y\\)-direction excluding the one that was interchanged. Also notice that the angle of this vector \\(x\\) has a signed angle of \\(\\alpha{}-k\\) with \\(y=0\\) and has length \\(x\\) - meaning that the vertical displacement of the vector \\(x\\) from \\(y=0\\) is equal to \\(x\\sin{(\\alpha{}-k)}\\)! Similarly, we get that the vertical displacement of the vector \\(y\\) is equivalent to \\(y\\sin{(k-\\beta{}})\\).\n\nAdding these two together, we get that \\(x\\sin{(\\alpha{}-k)}+y\\sin{(k-\\beta{}})\\) is simply the vertical displacement of the sum of the vectors \\(x\\) and \\(y\\). Since the sum of the vectors \\(x\\) and \\(y\\) is equivalent to the sum of the vectors in the positive half of the polygon minus the sum of the black vector that would be switched with the white vector(the leftmost vector in the positive half of the polygon) and the rightmost vector in the positive half(which is the parallel pair of the white vector that would be interchanged later), and we know that this sum happens to have a vertical displacement of \\(0\\), along with the fact that the positive half of the polygon summed together also has a vertical displacement of \\(0\\), we get that the total vertical displacement is \\(0\\), meaning that \\(x\\sin{(\\alpha{}-k)}+y\\sin{(k-\\beta{}})=0\\), and we are done.\n\n~by @peppapig_", "Pick a pair of parallel sides of the regular \\(2n\\)-gon and flip their color. Let those two lines be the horizontal. The area of the polygon formed by rods of the original color decreases by a parallelogram with height equal to the sum of the vertical heights of the rods of that color divided by \\(2\\) and base length \\(1\\). Similarly, the area of the polygon formed by rods of the new color increases by the sum of the vertical heights of the rods of the new color divided by \\(2\\). The sum of the heights of all the rod is equal to twice the height of the \\(2n\\)-gon, so the difference between the areas of \\(B\\) and \\(W\\) changes by the height of the \\(2n\\)-gon, which is fixed when \\(n\\) is fixed.\n\nIf we pair two pairs different-colored pairs of parallel sides of the \\(2n\\)-gon and flip the colors of both pairs, then \\(b\\) and \\(w\\) and the difference in the areas of \\(B\\) and \\(W\\) will remain unchanged. Thus, the difference in the areas of \\(B\\) and \\(W\\) depends only on \\(b\\) and \\(w\\)." ]
USAMO-2022-3
https://artofproblemsolving.com/wiki/index.php/2022_USAMO_Problems/Problem_3
Let \(\mathbb{R}_{>0}\) be the set of all positive real numbers. Find all functions \(f:\mathbb{R}_{>0} \to \mathbb{R}_{>0}\) such that for all \(x,y\in \mathbb{R}_{>0}\) we have \[ f(x) = f(f(f(x)) + y) + f(xf(y)) f(x+y). \]
[ "[WIP]" ]
USAMO-2022-4
https://artofproblemsolving.com/wiki/index.php/2022_USAMO_Problems/Problem_4
Find all pairs of primes \((p, q)\) for which \(p-q\) and \(pq-q\) are both perfect squares.
[ "Since \\(q(p-1)\\) is a perfect square and \\(q\\) is prime, we should have \\(p - 1 = qb^2\\) for some positive integer \\(b\\). Let \\(a^2 = p - q\\). Therefore, \\(q = p - a^2\\), and substituting that into the \\(p - 1 = qb^2\\) and solving for \\(p\\) gives\n\n\\[\np = \\frac{a^2b^2 - 1}{b^2 - 1} = \\frac{(ab - 1)(ab + 1)}{b^2 - 1}.\n\\]\n\nNotice that we also have\n\n\\[\np = \\frac{a^2b^2 - 1}{b^2 - 1} = a^2 + \\frac{a^2 - 1}{b^2 - 1}\n\\]\n\nand so \\(b^2 - 1 | a^2 - 1\\). We run through the cases\n\n- \\(a = 1\\): Then \\(p - q = 1\\) so \\((p, q) = (3, 2)\\), which works.\n- \\(a = b\\): This means \\(p = a^2 + 1\\), so \\(q = 1\\), a contradiction.\n- \\(a > b\\): This means that \\(b^2 - 1 < ab - 1\\). Since \\(b^2 - 1\\) can be split up into two factors \\(F_1, F_2\\) such that \\(F_1 | ab - 1\\) and \\(F_2 | ab + 1\\), we get\n\n\\[\np = \\frac{ab - 1}{F_1} \\cdot \\frac{ab + 1}{F_2}\n\\]\n\nand each factor is greater than \\(1\\), contradicting the primality of \\(p\\).\n\nThus, the only solution is \\((p, q) = (3, 2)\\).", "Let \\(p-q = a^2\\), \\(pq - q = b^2\\), where \\(a, b\\) are positive integers. \\(b^2 - a^2 = pq - q - (p-q) = pq -p\\). So,\n\n\\[\nb^2 - a^2 = p(q-1) \\tag{1}\n\\]\n\n\\(\\bullet\\) For \\(q=2\\), \\(p = b^2 - a^2 = (b-a)(b+a)\\). Then \\(b-a=1\\) and \\(b+a=p\\). \\(a=\\dfrac{p-1}{2}\\) and \\(p-q = a^2\\). Thus, \\(p - 2 = \\left( \\dfrac{p-1}{2} \\right)^2 \\implies p^2 - 6p + 9 = 0\\) and we find \\(p=3\\). Hence \\((p,q) = (3,2)\\).\n\n\\(\\bullet\\) For \\(q=4k+3\\), (\\(k\\geq 0\\) integer), by \\((1)\\), \\(p(4k+2) = b^2 - a^2\\). Let's examine in \\(\\mod 4\\), \\(b^2 - a^2 \\equiv 2 \\pmod{4}\\). But we know that \\(b^2 - a^2 \\equiv 0, 1 \\text{ or } 3 \\pmod{4}\\). This is a contradiction and no solution for \\(q = 4k + 3\\).\n\n\\(\\bullet\\) For \\(q=4k+1\\), (\\(k > 0\\) integer), by \\((1)\\), \\(p(4k) = b^2 - a^2\\). Let \\(k=m\\cdot n\\), where \\(m\\geq n \\geq 1\\) and \\(m, n\\) are integers. Since \\(p>q\\), we see \\(p>4k\\). Thus, by \\((1)\\), \\((b-a)(b+a) = 4p\\cdot m \\cdot n\\). \\(b-a\\) and \\(b+a\\) are same parity and \\(4p\\cdot m \\cdot n\\) is even integer. So, \\(b-a\\) and \\(b+a\\) are both even integers. Therefore,\n\n\\(\\left\\{ \\begin{array}{rcr} b+a = & 2pn \\\\ b-a = & 2m \\end{array} \\right.\\) or \\(\\left\\{ \\begin{array}{rcr} b+a = & 2pm \\\\ b-a = & 2n \\end{array} \\right.\\) Therefore, \\(a=pn - m\\) or \\(a = pm - n\\). For each case, \\(p-q = p - 4mn - 1 < a\\). But \\(p-q = a^2\\), this gives a contradiction. No solution for \\(q = 4k + 1\\).\n\nWe conclude that the only solution is \\((p,q) = (3,2)\\).\n\n(Lokman GÖKÇE)", "We claim that the only solution is \\((p,q) = (3,2)\\). First, this clearly works, since \\(3-2=1^2\\) and \\(3\\cdot2-2=2^2\\).\n\nStart as in Solution 2 - assign positive integers \\(m, n,\\) such that \\(p-q=m^2\\) and \\(pq-q=n^2.\\) Then, \\(pq-p=n^2-m^2,\\) or \\(p(q-1)=(n-m)(n+m).\\)\n\nFirst, \\(p=q\\) is impossible, since \\(p^2-p = p(p-1)\\) is not a perfect square. Now notice that \\(m<p\\) and \\(n<\\max(p,q-1).\\) But \\(p-q>0,\\) so \\(p\\geq q-1.\\) Now, \\(m, n < p.\\)\n\nTherefore, \\(n-m\\) can't ever be a factor of \\(p,\\) and \\(n+m<2p,\\) so \\(n+m=p.\\) Then, \\(n-m=q-1,\\) so \\(p-q=2m-1.\\) But \\(p-q=m^2,\\) so \\(m^2=2m-1,\\) or \\(m=1.\\) Therefore, one of \\(p\\) and \\(q\\) must be \\(2\\) (and \\(1\\) is not a prime).\n\nThus, we conclude that our claim is true.\n\n(grape1984)" ]
USAMO-2022-5
https://artofproblemsolving.com/wiki/index.php/2022_USAMO_Problems/Problem_5
A function \(f: \mathbb{R}\to \mathbb{R}\) is \(\textit{essentially increasing}\) if \(f(s)\leq f(t)\) holds whenever \(s\leq t\) are real numbers such that \(f(s)\neq 0\) and \(f(t)\neq 0\). Find the smallest integer \(k\) such that for any 2022 real numbers \(x_1,x_2,\ldots , x_{2022},\) there exist \(k\) essentially increasing functions \(f_1,\ldots, f_k\) such that \[ f_1(n) + f_2(n) + \cdots + f_k(n) = x_n\qquad \text{for every } n= 1,2,\ldots 2022. \]
[ "Coming soon." ]
USAMO-2022-6
https://artofproblemsolving.com/wiki/index.php/2022_USAMO_Problems/Problem_6
There are \(2022\) users on a social network called Mathbook, and some of them are Mathbook-friends. (On Mathbook, friendship is always mutual and permanent.) Starting now, Mathbook will only allow a new friendship to be formed between two users if they have at least two friends in common. What is the minimum number of friendships that must already exist so that every user could eventually become friends with every other user?
[ "To answer this question, we need to consider how friendships on Mathbook are formed. If two users have at least two friends in common, then they will be able to become friends with each other. This means that the minimum number of friendships that must already exist in order for every user to eventually become friends with every other user is the smallest number of friendships that guarantees that each pair of users has at least two friends in common.\n\nOne way to guarantee that each pair of users has at least two friends in common is to create a complete graph, where each user is friends with every other user. In this case, each pair of users has exactly \\(2022 - 2 = 2020\\) friends in common. However, this is not the minimum number of friendships required, since some pairs of users may have more than two friends in common without forming a complete graph.\n\nTo find the minimum number of friendships required, we can use the fact that a friendship between two users implies that they share all of their mutual friends. In other words, if two users are friends, then they must have at least two friends in common. This means that if we want to guarantee that each pair of users has at least two friends in common, we can do so by ensuring that each user has at least two friends.\n\nTherefore, the minimum number of friendships that must already exist in order for every user to eventually become friends with every other user is \\(2 \\cdot 2022 = \\boxed{4044}\\)." ]
USAMO-2023-1
https://artofproblemsolving.com/wiki/index.php/2023_USAMO_Problems/Problem_1
In an acute triangle \(ABC\), let \(M\) be the midpoint of \(\overline{BC}\). Let \(P\) be the foot of the perpendicular from \(C\) to \(AM\). Suppose that the circumcircle of triangle \(ABP\) intersects line \(BC\) at two distinct points \(B\) and \(Q\). Let \(N\) be the midpoint of \(\overline{AQ}\). Prove that \(NB=NC\).
[ "Let \\(X\\) be the foot from \\(A\\) to \\(\\overline{BC}\\). By definition, \\(\\angle AXM = \\angle MPC = 90^{\\circ}\\). Thus, \\(\\triangle AXM \\sim \\triangle CPM\\), and \\(\\triangle BMP \\sim \\triangle AMQ\\).\n\nFrom this, we have \\(\\frac{MP}{MX} = \\frac{MC}{MA} = \\frac{MP}{MQ}\\), as \\(MC=MB\\). Thus, \\(M\\) is also the midpoint of \\(XQ\\).\n\nNow, \\(NB = NC\\) if \\(N\\) lies on the perpendicular bisector of \\(\\overline{BC}\\). As \\(N\\) lies on the perpendicular bisector of \\(\\overline{XQ}\\), which is also the perpendicular bisector of \\(\\overline{BC}\\) (as \\(M\\) is also the midpoint of \\(XQ\\)), we are done. ~ Martin2001 ~ SomebodyST (minor edits)", "We are going to use barycentric coordinates on \\(\\triangle ABC\\). Let \\(A=(1,0,0)\\), \\(B=(0,1,0)\\), \\(C=(0,0,1)\\), and \\(a=BC\\), \\(b=CA\\), \\(c=AB\\). We have \\(M=\\left(0,\\frac{1}{2},\\frac{1}{2}\\right)\\) and \\(P=(x:1:1)\\) so \\(\\overrightarrow{CP}=\\left(\\frac{x}{x+2},\\frac{1}{x+2},\\frac{1}{x+2}-1\\right)\\) and \\(\\overrightarrow{AM}=\\left(-1,\\frac{1}{2},\\frac{1}{2}\\right)\\). Since \\(\\overleftrightarrow{CP}\\perp\\overleftrightarrow{AM}\\), it follows that\n\n\\[\n\\begin{align*} a^2\\left(\\frac{1}{2}\\cdot\\frac{1}{x+2}+\\frac{1}{2}\\left(\\frac{1}{x+2}-1\\right)\\right)+b^2\\left(\\frac{1}{2}\\cdot\\frac{x}{x+2}-\\left(\\frac{1}{x+2}-1\\right)\\right)\\\\ +c^2\\left(\\frac{1}{2}\\cdot\\frac{x}{x+2}-\\frac{1}{x+2}\\right)=0. \\end{align*}\n\\]\n\nSolving this gives\n\n\\[\nx=\\frac{2b^2-2c^2}{a^2-3b^2-c^2}\n\\]\n\nso\n\n\\[\nP=\\left(\\frac{b^2-c^2}{a^2-2b^2-2c^2},\\frac{a^2-3b^2-c^2}{2a^2-4b^2-4c^2},\\frac{a^2-3b^2-c^2}{2a^2-4b^2-4c^2}\\right).\n\\]\n\nThe equation for \\((ABP)\\) is\n\n\\[\n-a^2yz-b^2zx-c^2xy+ux+vy+wz=0.\n\\]\n\nPlugging in \\(A\\) and \\(B\\) gives \\(u=v=0\\). Plugging in \\(P\\) gives\n\n\\[\n\\begin{align*} -a^2\\left(\\frac{a^2-3b^2-c^2}{2a^2-4b^2-4c^2}\\right)^2-b^2\\cdot\\frac{a^2-3b^2-c^2}{2a^2-4b^2-4c^2}\\cdot\\frac{b^2-c^2}{a^2-2b^2-2c^2}\\\\ -c^2\\cdot\\frac{b^2-c^2}{a^2-2b^2-2c^2}\\cdot\\frac{a^2-3b^2-c^2}{2a^2-4b^2-4c^2}+w\\cdot\\frac{a^2-3b^2-c^2}{2a^2-4b^2-4c^2}=0 \\end{align*}\n\\]\n\nso\n\n\\[\nw=\\frac{2b^4-2c^4+a^4-3a^2b^2-a^2c^2}{2a^2-4b^2-4c^2}=\\frac{a^2}{2}-\\frac{b^2}{2}+\\frac{c^2}{2}.\n\\]\n\nNow let \\(Q=(0,t,1-t)\\) where\n\n\\[\n\\begin{align*} -a^2t(1-t)+w(1-t)&=0\\\\ \\implies t&=\\frac{w}{a^2} \\end{align*}\n\\]\n\nso \\(Q=\\left(0,\\frac{w}{a^2},1-\\frac{w}{a^2}\\right)\\). It follows that \\(N=\\left(\\frac{1}{2},\\frac{w}{2a^2},1-\\frac{w}{2a^2}\\right)\\). It suffices to prove that \\(\\overleftrightarrow{ON}\\perp\\overleftrightarrow{BC}\\). Setting \\(\\overrightarrow{O}=0\\), we get \\(\\overrightarrow{N}=\\left(\\frac{1}{2},\\frac{w}{2a^2},1-\\frac{w}{2a^2}\\right)\\). Furthermore we have \\(\\overrightarrow{CB}=(0,1,-1)\\) so it suffices to prove that\n\n\\[\n\\begin{align*} a^2\\left(-\\frac{w}{2a^2}+\\frac{1}{2}-\\frac{u}{2a^2}\\right)+b^2\\left(-\\frac{1}{2}\\right)+c^2\\left(\\frac{1}{2}\\right)=0\\\\ \\implies w=\\frac{a^2}{2}-\\frac{b^2}{2}+\\frac{c^2}{2} \\end{align*}\n\\]\n\nwhich is valid. \\(\\square\\)\n\n~KevinYang2.71" ]
USAMO-2023-2
https://artofproblemsolving.com/wiki/index.php/2023_USAMO_Problems/Problem_2
Let \(\mathbb{R}^{+}\) be the set of positive real numbers. Find all functions \(f:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+}\) such that, for all \(x, y \in \mathbb{R}^{+}\), \[ f(xy + f(x)) = xf(y) + 2 \]
[ "Make the following substitutions to the equation:\n\n1. \\((x, 1) \\rightarrow f(x + f(x)) = xf(1) + 2\\)\n\n2. \\((1, x + f(x)) \\rightarrow f(x + f(x) + f(1)) = f(x + f(x)) + 2 = xf(1) + 4\\)\n\n3. \\((x, 1 + \\frac{f(1)}{x}) \\rightarrow f(x + f(x) + f(1)) = xf\\biggl(1 + \\frac{f(1)}{x}\\biggr) + 2\\)\n\nIt then follows from (2) and (3) that \\(f(1 + \\frac{f(1)}{x}) = f(1) + \\frac{2}{x}\\), so we know that this function is linear for \\(x > 1\\). Substitute \\(f(x) = ax+b\\) and solve for \\(a\\) and \\(b\\) in the functional equation; we find that \\(f(x) = x + 1 \\forall x > 1\\).\n\nNow, we can let \\(x > 1\\) and \\(y \\le 1\\). Since \\(f(x) = x + 1\\), \\(xy + f(x) > x > 1\\), so \\(f(xy + f(x)) = xy + x + 2 = xf(y) + 2\\). It becomes clear then that \\(f(y) = y + 1\\) as well, so \\(f(x) = x + 1\\) is the only solution to the functional equation.\n\n~jkmmm3" ]
USAMO-2023-3
https://artofproblemsolving.com/wiki/index.php/2023_USAMO_Problems/Problem_3
Consider an \(n\)-by-\(n\) board of unit squares for some odd positive integer \(n\). We say that a collection \(C\) of identical dominoes is a maximal grid-aligned configuration on the board if \(C\) consists of \((n^2-1)/2\) dominoes where each domino covers exactly two neighboring squares and the dominoes don't overlap: \(C\) then covers all but one square on the board. We are allowed to slide (but not rotate) a domino on the board to cover the uncovered square, resulting in a new maximal grid-aligned configuration with another square uncovered. Let \(k(C)\) be the number of distinct maximal grid-aligned configurations obtainable from \(C\) by repeatedly sliding dominoes. Find all possible values of \(k(C)\) as a function of \(n\).
[ "We claim the answer is \\((\\frac{n+1}{2})^2\\).\n\nFirst, consider a checkerboard tiling of the board with 4 colors: R, G, B, Y. Number each column from \\(1\\) to \\(n\\) from left to right and each row from \\(1\\) to \\(n\\) from top to bottom. We color a tile R if its row and column are odd, a tile G is its row is even but its column is odd, a tile B if its row and column is even, and a tile Y if its row is odd but its column is even.\n\nLemma 1: Throughout our moves, the color of the uncolored tile stays an invariant.\n\nConsider that a domino can either only change rows or can only change columns. Therefore, sliding a domino into the hole and creating a new one has two possible colors. Of these, note that the new hole will always trivially be two tiles away from the old hole, meaning that the parity of both the row and column number stays the same. Thus, the lemma holds.\n\nLemma 2: There are more red tiles than any other color. Because each color is uniquely defined by the parity of a pair of column and row number, it satisfies to show that given an odd integer \\(n\\), there are more odd positive integers less than or equal to \\(n\\) than even ones. Obviously, this is true, and so red will have more tiles than any other color.\n\nLemma 3: For any starting configuration \\(C\\) and any blank tile \\(B\\) such that the blank tile's color matches the blank tile's color of \\(C\\), there is no more than one unique configuration \\(C'\\) that can be produced from \\(C\\) using valid moves.\n\nWe will use proof by contradiction. Assume there exists two different \\(C'\\). We can get from one of these \\(C'\\) to another using moves. However, we have to finish off with the same hole as before. Before the last move, the hole must be two tiles away from the starting hole. However, because the domino we used to move into the blank tile's spot is in the way, that hole must be congruent to the hole produced after the first move. We can induct this logic, and because there is a limited amount of tiles with the same color, eventually we will run out of tiles to apply this to. Therefore, having two distinct \\(C'\\) with the same starting hole \\(B\\) is impossible with some \\(C\\).\n\nWe will now prove that \\((\\frac{n+1}{2})^2\\) is the answer. There are \\(\\frac{n+1}{2}\\) rows and \\(\\frac{n+1}{2}\\) columns that are odd, and thus there are \\((\\frac{n+1}{2})^2\\) red tiles. Given lemma 3, this is our upper bound for a maximum. To establish that \\((\\frac{n+1}{2})^2\\) is indeed possible, we construct such a \\(C\\):\n\nIn the first column, leave the first tile up blank. Then, continuously fill in vertically oriented dominos in that column until it reaches the bottom. In the next \\(n-1\\) columns, place \\(\\frac{n-1}{2}\\) vertically oriented dominos in a row starting from the top. At the bottom row, starting with the first unfilled tile on the left, place horizontally aligned dominos in a row until you reach the right.\n\nObviously, the top left tile is red. It suffices to show that any red tile may be uncovered. For the first column, one may slide some dominos on the first column until the desired tile is uncovered. For the bottom row, all the first dominos may be slid up, and then the bottom dominos may be slid to the left until the desired red tile is uncovered. Finally, for the rest of the red tiles, the bottom red tile in the same color may be revealed, and then vertically aligned dominos in the same column may be slid down until the desired tile is revealed. Therefore, this configuration may produce \\((\\frac{n+1}{2})^2\\) different configurations with moves.\n\nHence, we have proved that \\((\\frac{n+1}{2})^2\\) is the maximum, and we are done. \\(\\blacksquare{}\\)\n\n~SigmaPiE\n\nDo not change the question yourself, SigmaPiE. The question was not maximum, it was all possible values. I changed it back to all possible values. The answer is {1,2,...,((n-1)/2)^2} U {((n+1)/2)^2}" ]
USAMO-2023-4
https://artofproblemsolving.com/wiki/index.php/2023_USAMO_Problems/Problem_4
A positive integer \(a\) is selected, and some positive integers are written on a board. Alice and Bob play the following game. On Alice's turn, she must replace some integer \(n\) on the board with \(n+a\), and on Bob's turn he must replace some even integer \(n\) on the board with \(n/2\). Alice goes first and they alternate turns. If on his turn Bob has no valid moves, the game ends. After analyzing the integers on the board, Bob realizes that, regardless of what moves Alice makes, he will be able to force the game to end eventually. Show that, in fact, for this value of \(a\) and these integers on the board, the game is guaranteed to end regardless of Alice's or Bob's moves.
[ "The contrapositive of the claim is somewhat easier to conceptualize: If it is not guaranteed that the game will end (i.e. the game could potentially last forever), then Bob is not able to force the game to end (i.e. Alice can force it to last forever). So we want to prove that, if the game can potentially last indefinitely, then Alice can force it to last indefinitely. Clearly, if there is \\(1\\) number on the board initially, all moves are forced. This means the claim is true in this specific case, because if the game \"potentially\" can last forever, this means it must last forever, since the game can only be played in one way. Ergo Alice can \"force\" this to occur because it is guaranteed to occur. Now we look at all cases where there is more than \\(1\\) number on the board.\n\nCase 1: \\(v_2 (a)=0\\)\n\nThe game lasts forever here no matter what. This is true because, if the game ends, it means the board was in some position \\(P\\), Alice added \\(a\\) to some number on the board, and all the numbers now on the board are odd. If there are only odd numbers on the board in position \\(P\\), Alice will add \\(a\\) to some number on the board, making it even, meaning this cannot have been the state \\(P\\) of the board. If at least one number n on the board is even, Alice can add \\(a\\) to a number other than \\(n\\), meaning there is still at least 1 even number on the board, meaning this also cannot have been the state \\(P\\) of the board. This covers all possible boards when \\(v_2 (a)=0\\), so we're done.\n\nCase 2: \\(v_2 (a)=1\\)\n\nIf there is at least one number n on the board that is even, the game can also last forever. On any move, Alice will add \\(a\\) to this number \\(n\\) if and only if \\(v_2 (n)=1\\). This way, the new number \\(n'= n+a\\) satisfies \\(v_2 (n') \\geq 2\\). If Bob does divides \\(n'\\) until \\(v_2 (n)=1\\), Alice will again add \\(a\\) to \\(n\\) resulting in \\(v_2 (n') \\geq 2\\). This means that Alice can always keep an even number on the board for Bob to divide no matter how Bob plays.\n\nIf there is no even number on the board, then the game can clearly not last forever. No matter what Alice does, Bob will have no even number to divide after her move.\n\nGeneral Case: \\(v_2 (a)=x\\)\n\nIn general, it seems to be the case that the game can last indefinitely for some given \\(a\\) if and only if there exists some number \\(n\\) on the board such that \\(v_2 (n) \\geq v_2 (a)\\), so we should aim to prove this.\n\n1. \"If\"\n\nAlice can apply a similar strategy to the strategy used in case 2. On any move, Alice will add \\(a\\) to \\(n\\) if and only if \\(v_2 (n)=v_2 (a)\\). If she does this addition, then \\(v_2 (n') \\geq v_2 (a)+1\\), keeping an even number on the board. Even if Bob divides \\(n'\\) until \\(v_2 (n)=v_2 (a)\\), Alice will apply the same strategy and keep \\(v_2 (n') \\geq v_2 (a)+1\\). Alice's use of this strategy ensures that there always exists some number n on the board such that \\(v_2 (n) \\geq v_2 (a)\\), ensuring there always exists an even number n on the board.\n\n2.\"Only If\"\n\nIf \\(v_2 (n) < v_2 (a)\\) for all n on the board, this means that Alice can never change the value of \\(v_2 (n)\\) for any \\(n\\) on the board. Only Bob can do this, and Bob will subtract \\(1\\) from each \\(v_2 (n)\\) until they are all equal to \\(0\\) (all odd), ending the game.\n\nWe've shown that the game can last indefinitely iff there exists some number \\(n\\) on the board such that \\(v_2 (n) \\geq v_2 (a)\\), and have shown that Alice can ensure the game lasts forever in these scenarios using the above strategy. This proves the contrapositive, proving the claim.\n\n~SpencerD." ]
USAMO-2023-5
https://artofproblemsolving.com/wiki/index.php/2023_USAMO_Problems/Problem_5
Let \(n\geq3\) be an integer. We say that an arrangement of the numbers \(1\), \(2\), \(\dots\), \(n^2\) in a \(n \times n\) table is row-valid if the numbers in each row can be permuted to form an arithmetic progression, and column-valid if the numbers in each column can be permuted to form an arithmetic progression. For what values of \(n\) is it possible to transform any row-valid arrangement into a column-valid arrangement by permuting the numbers in each row?
[ "The answer is all \\(\\boxed{\\text{prime } n}\\).\n\n## Proof that primes work\n\nSuppose \\(n=p\\) is prime. Then, let the arithmetic progressions in the \\(i\\)th row have least term \\(a_i\\) and common difference \\(d_i\\). For each cell with integer \\(k\\), assign a monomial \\(x^k\\). The sum of the monomials is\n\n\\[\nx(1+x+\\ldots+x^{n^2-1}) = \\sum_{i=1}^n x^{a_i}(1+x^{d_i}+\\ldots+x^{(n-1)d_i}),\n\\]\n\nwhere the LHS is obtained by summing over all cells and the RHS is obtained by summing over all rows. Let \\(S\\) be the set of \\(p\\)th roots of unity that are not \\(1\\); then, the LHS of the above equivalence vanishes over \\(S\\) and the RHS is\n\n\\[\n\\sum_{p \\mid d_i} x^{a_i}.\n\\]\n\nReducing the exponents (mod \\(p\\)) in the above expression yields\n\n\\[\nf(x) := \\sum_{p \\mid d_i} x^{a_i \\pmod{p}} = 0\n\\]\n\nwhen \\(x \\in S\\). Note that \\(\\prod_{s \\in S} (x-s)=1+x+\\ldots+x^{p-1}\\) is a factor of \\(f(x)\\), and as \\(f\\) has degree less than \\(p\\), \\(f\\) is either identically 0 or \\(f(x)=1+x+\\ldots+x^{p-1}\\).\n\n- If \\(f\\) is identically 0, then \\(p\\) never divides \\(d_i\\). Thus, no two elements in each row are congruent \\(\\pmod{p}\\), so all residues are represented in each row. Now we can rearrange the grid so that column \\(i\\) consists of all numbers \\(i \\pmod{p}\\), which works.\n\n- If \\(f(x)=1+x+\\ldots+x^{p-1}\\), then \\(p\\) always divides \\(d_i\\). It is clear that each \\(d_i\\) must be \\(p\\), so each row represents a single residue \\(\\pmod{p}\\). Thus, we can rearrange the grid so that column \\(i\\) contains all consecutive numbers from \\(1 + (i-1)p\\) to \\(ip\\), which works.\n\nAll in all, any prime \\(n\\) satisfies the hypotheses of the problem.\n\n## Proof that composites do not work\n\nLet \\(n=ab\\). Look at the term \\(a^2b+ab\\); we claim it cannot be part of a column that has cells forming an arithmetic sequence after any appropriate rearranging. After such a rearrangement, if the column it is in has common difference \\(d<ab=n\\), then \\(a^2b+ab-d\\) must also be in its column, which is impossible. If the column has difference \\(d > ab = n\\), then no element in the next row can be in its column. If the common difference is \\(d = ab = n\\), then \\(a^2b + ab - 2d = a^2b - ab\\) and \\(a^2b + ab - d = a^2b\\), which are both in the row above it, must both be in the same column, which is impossible. Therefore, the grid is not column-valid after any rearrangement, which completes the proof.\n\n~ Leo.Euler" ]
USAMO-2023-6
https://artofproblemsolving.com/wiki/index.php/2023_USAMO_Problems/Problem_6
Let ABC be a triangle with incenter \(I\) and excenters \(I_a\), \(I_b\), \(I_c\) opposite \(A\), \(B\), and \(C\), respectively. Given an arbitrary point \(D\) on the circumcircle of \(\triangle ABC\) that does not lie on any of the lines \(II_{a}\), \(I_{b}I_{c}\), or \(BC\), suppose the circumcircles of \(\triangle DIIa\) and \(\triangle DI_bI_c\) intersect at two distinct points \(D\) and \(F\). If \(E\) is the intersection of lines \(DF\) and \(BC\), prove that \(\angle BAD = \angle EAC\).
[ "\\[\n[asy] size(500); pair A,B,C,D,E,F,G,H,I,J,K,IA,IB,IC,P,Q; B=(0,0); C=(8,0); A=intersectionpoint(Circle(B,6),Circle(C,9)); I=incenter(A,B,C); path c=circumcircle(A,B,C); J=intersectionpoint(I--(4*I-3*A),c); IA=2*J-I; IB=2*intersectionpoint(I--(4*I-3*B),c)-I; IC=2*intersectionpoint(I--(4*I-3*C),c)-I; K=intersectionpoint(IB--IC,c); D=intersectionpoint(I--(I+(10,-12)),c); path c1=circumcircle(D,I,IA),c2=circumcircle(D,IB,IC); F=intersectionpoints(c1,c2)[1]; E=extension(B,C,D,F); G=intersectionpoint(c1,c); H=intersectionpoint(c2,c); P=extension(A,I,B,C); Q=extension(IB,IC,B,C); draw(A--B--C--A); draw(c); draw(A--J); draw(circumcircle(D,I,IA)); draw(circumcircle(D,IB,IC)); draw(D--F); draw(B--Q--IB); draw(G--J,dashed); draw(H--K,dashed); dot(\"$A$\",A,dir(A-circumcenter(A,B,C))); dot(\"$B$\",B,1/2*dir(B-dir(circumcenter(A,B,C))*dir(90)+dir(B-C))); dot(\"$C$\",C,dir(C-circumcenter(A,B,C))*dir(15)); dot(\"$D$\",D,dir(dir(90)*dir(circumcenter(D,I,IA)-D)+dir(90)*dir(D-circumcenter(D,IB,IC)))); dot(\"$E$\",E,dir(dir(H-K)+dir(B-C))); dot(\"$F$\",F,dir(dir(90)*dir(F-circumcenter(D,I,IA))+dir(90)*dir(circumcenter(D,IB,IC)-F))); dot(\"$G$\",G,dir(dir(90)*dir(G-circumcenter(D,I,IA))+dir(90)*dir(circumcenter(A,B,C)-G))); dot(\"$H$\",H,dir(dir(90)*dir(circumcenter(D,IB,IC)-H)+dir(90)*dir(H-circumcenter(A,B,C)))); dot(\"$I$\",I,dir(dir(90)*dir(circumcenter(D,I,IA)-I)+dir(A-I))); dot(\"$J$\",J,dir(J-circumcenter(A,B,C))); dot(\"$K$\",K,dir(K-circumcenter(A,B,C))); dot(\"$I_A$\",IA,dir(IA-circumcenter(D,I,IA))); dot(\"$I_B$\",IB,dir(dir(IB-IC)+dir(IB-IA))); dot(\"$I_C$\",IC,dir(dir(90)*dir(circumcenter(D,IB,IC)-IC)+dir(IC-IB))); dot(\"$P$\",P,dir(dir(A-I)+dir(C-B))); dot(\"$Q$\",Q,dir(dir(IC-IB)+dir(B-C))); [/asy]\n\\]\n\nConsider points \\(G,H,J,K,P,\\) and \\(Q\\) such that the intersections of the circumcircle of \\(\\triangle{}ABC\\) with the circumcircle of \\(\\triangle{}DII_A\\) are \\(D\\) and \\(G\\), the intersections of the circumcircle of \\(\\triangle{}ABC\\) with the circumcircle of \\(\\triangle{}DI_BI_C\\) are \\(D\\) and \\(H\\), the intersections of the circumcircle of \\(\\triangle{}ABC\\) with line \\(\\overline{II_A}\\) are \\(A\\) and \\(J\\), the intersections of the circumcircle of \\(\\triangle{}ABC\\) with line \\(\\overline{I_BI_C}\\) are \\(A\\) and \\(K\\), the intersection of lines \\(\\overline{II_A}\\) and \\(\\overline{BC}\\) is \\(P\\), and the intersection of lines \\(\\overline{I_BI_C}\\) and \\(\\overline{BC}\\) is \\(Q\\).\n\nSince \\(IBI_AC\\) is cyclic, the pairwise radical axes of the circumcircles of \\(\\triangle{}DII_A,\\triangle{}ABC,\\) and \\(IBI_AC\\) concur. The pairwise radical axes of these circles are \\(\\overline{GD},\\overline{II_A},\\) and \\(\\overline{BC}\\), so \\(G,P,\\) and \\(D\\) are collinear. Similarly, since \\(BCI_BI_C\\) is cyclic, the pairwise radical axes of the cirucmcircles of \\(\\triangle{}DI_BI_C,\\triangle{}ABC,\\) and \\(BCI_BI_C\\) concur. The pairwise radical axes of these circles are \\(\\overline{HD},\\overline{I_BI_C},\\) and \\(\\overline{BC}\\), so \\(H,Q,\\) and \\(D\\) are collinear. This means that \\(-1=(Q,P;B,C)\\stackrel{D}{=}(H,G;B,C)\\), so the tangents to the circumcircle of \\(\\triangle{}ABC\\) at \\(G\\) and \\(H\\) intersect on \\(\\overline{BC}\\). Let this intersection be \\(X\\). Also, let the intersection of the tangents to the circumcircle of \\(\\triangle{}ABC\\) at \\(K\\) and \\(J\\) be a point at infinity on \\(\\overline{BC}\\) called \\(Y\\) and let the intersection of lines \\(\\overline{KG}\\) and \\(\\overline{}HJ\\) be \\(Z\\). Then, let the intersection of lines \\(\\overline{GJ}\\) and \\(\\overline{HK}\\) be \\(E'\\). By Pascal's Theorem on \\(GGJHHK\\) and \\(GJJHKK\\), we get that \\(X,E',\\) and \\(Z\\) are collinear and that \\(E',Y,\\) and \\(Z\\) are collinear, so \\(E',X,\\) and \\(Y\\) are collinear, meaning that \\(E'\\) lies on \\(\\overline{BC}\\) since both \\(X\\) and \\(Y\\) lie on \\(\\overline{BC}\\).\n\n\\[\n[asy] size(500); pair A,B,C,D,E,F,G,H,I,J,K,IA,IB,IC,P,Q,GP; B=(0,0); C=(8,0); A=intersectionpoint(Circle(B,6),Circle(C,9)); I=incenter(A,B,C); path c=circumcircle(A,B,C); J=intersectionpoint(I--(4*I-3*A),c); IA=2*J-I; IB=2*intersectionpoint(I--(4*I-3*B),c)-I; IC=2*intersectionpoint(I--(4*I-3*C),c)-I; K=intersectionpoint(IB--IC,c); D=intersectionpoint(I--(I+(10,-12)),c); path c1=circumcircle(D,I,IA),c2=circumcircle(D,IB,IC); F=intersectionpoints(c1,c2)[1]; E=extension(B,C,D,F); G=intersectionpoint(c1,c); H=intersectionpoint(c2,c); P=extension(A,I,B,C); Q=extension(IB,IC,B,C); GP=extension(A,2*foot(G,A,I)-G,B,C); draw(A--B--C--A); draw(c); draw(A--J--G--D); draw(C--GP); draw(circumcircle(A,E,J),dashed); dot(\"$A$\",A,dir(A-circumcenter(A,B,C))); dot(\"$B$\",B,dir(B-circumcenter(A,B,C))); dot(\"$C$\",C,dir(dir(90)*dir(circumcenter(A,B,C)-C)+dir(C-B))); dot(\"$D$\",D,dir(dir(90)*dir(circumcenter(D,I,IA)-D)+dir(90)*dir(D-circumcenter(D,IB,IC)))); dot(\"$E'$\",E,dir(dir(J-G)+dir(B-C))); dot(\"$G$\",G,dir(dir(90)*dir(G-circumcenter(D,I,IA))+dir(90)*dir(circumcenter(A,B,C)-G))); dot(\"$I$\",I,dir(dir(90)*dir(circumcenter(D,I,IA)-I)+dir(A-I))); dot(\"$J$\",J,dir(J-circumcenter(A,B,C))); dot(\"$P$\",P,dir(dir(A-I)+dir(C-B))); dot(\"$G'$\",GP,dir(GP-B)); [/asy]\n\\]\n\nConsider the transformation which is the composition of an inversion centered at \\(A\\) and a reflection over the angle bisector of \\(\\angle{}CAB\\) that sends \\(B\\) to \\(C\\) and \\(C\\) to \\(B\\). We claim that this sends \\(D\\) to \\(E'\\) and \\(E'\\) to \\(D\\). It is sufficient to prove that if the transformation sends \\(G\\) to \\(G'\\), then \\(AE'JG'\\) is cyclic. Notice that \\(\\triangle{}AGB\\sim\\triangle{}ACG'\\) since \\(\\angle{}GAB=\\angle{}G'AC\\) and \\(\\tfrac{AG'}{AC}=\\tfrac{\\frac{AB\\cdot{}AC}{AG}}{AC}=\\tfrac{AB}{AG}\\). Therefore, we get that \\(\\angle{}AG'E'=\\angle{}ABG=\\angle{}AJE'\\), so \\(AE'JG'\\) is cyclic, proving the claim. This means that \\(\\angle{}BAE'=\\angle{}CAD\\).\n\n\\[\n[asy] size(500); pair A,B,C,D,E,F,G,H,I,J,K,IA,IB,IC,P,Q,GP,BP,CP,DP; B=(0,0); C=(8,0); A=intersectionpoint(Circle(B,6),Circle(C,9)); I=incenter(A,B,C); path c=circumcircle(A,B,C); J=intersectionpoint(I--(4*I-3*A),c); IA=2*J-I; IB=2*intersectionpoint(I--(4*I-3*B),c)-I; IC=2*intersectionpoint(I--(4*I-3*C),c)-I; K=intersectionpoint(IB--IC,c); D=intersectionpoint(I--(I+(10,-12)),c); path c1=circumcircle(D,I,IA),c2=circumcircle(D,IB,IC); F=intersectionpoints(c1,c2)[1]; E=extension(B,C,D,F); G=intersectionpoint(c1,c); H=intersectionpoint(c2,c); P=extension(A,I,B,C); Q=extension(IB,IC,B,C); BP=2*foot(B,IB,IC)-B; CP=2*foot(C,IB,IC)-C; DP=2*foot(D,IB,IC)-D; draw(A--B--C--A); draw(E--DP); draw(BP--A--CP); draw(IB--IC); draw(c); draw(circumcircle(B,IB,IC)); draw(circumcircle(E,IB,IC)); dot(\"$A$\",A,2*dir(dir(IB-A)+dir(C-A))); dot(\"$B$\",B,dir(B-circumcenter(A,B,C))); dot(\"$C$\",C,dir(dir(90)*dir(circumcenter(A,B,C)-C)+dir(C-B))); dot(\"$D$\",D,dir(dir(90)*dir(circumcenter(D,I,IA)-D)+dir(90)*dir(D-circumcenter(D,IB,IC)))); dot(\"$E'$\",E,dir(B-C)*dir(90)); dot(\"$I_B$\",IB,dir(dir(IB-IC)+dir(IB-IA))); dot(\"$I_C$\",IC,dir(dir(90)*dir(circumcenter(D,IB,IC)-IC)+dir(IC-IB))); dot(\"$B'$\",BP,dir(BP-circumcenter(B,IB,IC))); dot(\"$C'$\",CP,dir(CP-circumcenter(B,IB,IC))); dot(\"$D'$\",DP,dir(DP-E)); [/asy]\n\\]\n\nWe claim that \\(\\angle{}I_BE'I_C+\\angle{}I_BDI_C=180^\\circ\\). Construct \\(D'\\) to be the intersection of line \\(\\overline{AE'}\\) and the circumcircle of \\(\\triangle{}E'I_BI_C\\) and let \\(B'\\) and \\(C'\\) be the intersections of lines \\(\\overline{AC}\\) and \\(\\overline{AB}\\) with the circumcircle of \\(\\triangle{}BI_BI_C\\). Since \\(B'\\) and \\(C'\\) are the reflections of \\(B\\) and \\(C\\) over \\(\\overline{I_BI_C}\\), it is sufficient to prove that \\(A,B',C',D'\\) are concyclic. Since \\(\\overline{B'C},\\overline{D'E'},\\) and \\(\\overline{I_BI_C}\\) concur and \\(D',E',I_B,I_C\\) and \\(I_B,I_C,B',C\\) are concyclic, we have that \\(B',C,D',E'\\) are concyclic, so \\(\\angle{}B'D'A=\\angle{}ACE'=\\angle{}AC'B'\\), so \\(A,B',C',D'\\) are concyclic, proving the claim. We can similarly get that \\(\\angle{}IE'I_A=\\angle{}IDI_A\\).\n\n\\[\n[asy] size(500); pair A,B,C,D,E,F,G,H,I,J,K,IA,IB,IC,P,Q,JP,KP; B=(0,0); C=(8,0); A=intersectionpoint(Circle(B,6),Circle(C,9)); I=incenter(A,B,C); path c=circumcircle(A,B,C); J=intersectionpoint(I--(4*I-3*A),c); IA=2*J-I; IB=2*intersectionpoint(I--(4*I-3*B),c)-I; IC=2*intersectionpoint(I--(4*I-3*C),c)-I; K=intersectionpoint(IB--IC,c); D=intersectionpoint(I--(I+(10,-12)),c); path c1=circumcircle(D,I,IA),c2=circumcircle(D,IB,IC); F=intersectionpoints(c1,c2)[1]; E=extension(B,C,D,F); G=intersectionpoint(c1,c); H=intersectionpoint(c2,c); P=extension(A,I,B,C); Q=extension(IB,IC,B,C); JP=2*J-E; KP=2*K-E; draw(A--B--C--A); draw(c); draw(A--J); draw(circumcircle(D,I,IA)); draw(circumcircle(D,IB,IC)); draw(D--F,dashed); draw(B--Q--IB); draw(G--JP); draw(H--KP); dot(\"$A$\",A,dir(A-circumcenter(A,B,C))); dot(\"$B$\",B,1/2*dir(B-dir(circumcenter(A,B,C))*dir(90)+dir(B-C))); dot(\"$C$\",C,dir(C-circumcenter(A,B,C))*dir(15)); dot(\"$D$\",D,dir(dir(90)*dir(circumcenter(D,I,IA)-D)+dir(90)*dir(D-circumcenter(D,IB,IC)))); dot(\"$E'$\",E,dir(dir(H-K)+dir(B-C))); dot(\"$F$\",F,dir(dir(90)*dir(F-circumcenter(D,I,IA))+dir(90)*dir(circumcenter(D,IB,IC)-F))); dot(\"$G$\",G,dir(dir(90)*dir(G-circumcenter(D,I,IA))+dir(90)*dir(circumcenter(A,B,C)-G))); dot(\"$H$\",H,dir(dir(90)*dir(circumcenter(D,IB,IC)-H)+dir(90)*dir(H-circumcenter(A,B,C)))); dot(\"$I$\",I,dir(dir(90)*dir(circumcenter(D,I,IA)-I)+dir(A-I))); dot(\"$J$\",J,dir(dir(circumcenter(A,B,C)-J)*dir(90)+dir(J-G))); dot(\"$K$\",K,dir(dir(K-circumcenter(A,B,C))*dir(90)+dir(K-H))); dot(\"$I_A$\",IA,dir(IA-circumcenter(D,I,IA))); dot(\"$I_B$\",IB,dir(dir(IB-IC)+dir(IB-IA))); dot(\"$I_C$\",IC,dir(dir(90)*dir(circumcenter(D,IB,IC)-IC)+dir(IC-IB))); dot(\"$P$\",P,dir(dir(A-I)+dir(C-B))); dot(\"$Q$\",Q,dir(dir(IC-IB)+dir(B-C))); dot(\"$J'$\",JP,dir(JP-circumcenter(D,I,IA))); dot(\"$K'$\",KP,dir(KP-circumcenter(D,IB,IC))); [/asy]\n\\]\n\nLet line \\(\\overline{E'J}\\) intersect the circumcircle of \\(\\triangle{}DII_A\\) at \\(G\\) and \\(J'\\). Notice that \\(J\\) is the midpoint of \\(\\overline{II_A}\\) and \\(\\angle{}IE'I_A=\\angle{}IDI_A=\\angle{}IJ'I_A\\), so \\(IE'I_AJ'\\) is a parallelogram with center \\(J\\), so \\(\\tfrac{EJ}{EJ'}=\\tfrac{1}{2}\\). Similarly, we get that if line \\(\\overline{E'K}\\) intersects the circumcircle of \\(\\triangle{}DI_BI_C\\) at \\(H\\) and \\(K'\\), we have that \\(\\tfrac{EK}{EK'}=\\tfrac{1}{2}\\), so \\(\\overline{KJ}\\parallel\\overline{K'J'}\\), so \\(\\angle{}HGJ'=\\angle{}HGJ=\\angle{}HKJ=\\angle{}HK'J'\\), so \\(G,H,J',K'\\) are concyclic. Then, the pairwise radical axes of the circumcircles of \\(\\triangle{}DII_A,\\triangle{}DI_BI_C,\\) and \\(GHJ'K'\\) are \\(\\overline{DF},\\overline{HK'},\\) and \\(\\overline{GJ'}\\), so \\(\\overline{DF},\\overline{HK'},\\) and \\(\\overline{GJ'}\\) concur, so \\(\\overline{DF},\\overline{HK},\\) and \\(\\overline{GJ}\\) concur, so \\(E=E'\\). We are then done since \\(\\angle{}BAE'=\\angle{}CAD\\).\n\n~Zhaom", "Set \\(\\triangle ABC\\) as the reference triangle, and let \\(D=(x_0,y_0,z_0)\\) with homogenized coordinates. To find the equation of circle \\(DII_A\\), we note that \\(I=(a:b:c)\\) (not homogenized) and \\(I_A=(-a:b:c)\\). Thus, for this circle with equation \\(a^2yz+b^2xz+c^2xy=(x+y+z)(u_1x+v_1y+w_1z)\\), we compute that\n\n\\[\nu_1=bc,\n\\]\n\n\\[\nv_1=\\frac{bc^2x_0}{bz_0-cy_0},\n\\]\n\n\\[\nw_1=\\frac{b^2cx_0}{cy_0-bz_0}.\n\\]\n\nFor circle \\(DI_BI_C\\) with equation \\(a^2yz+b^2xz+c^2xy=(x+y+z)(u_2x+v_2y+w_2z)\\),, we find that\n\n\\[\nu_2=-bc,\n\\]\n\n\\[\nv_2=\\frac{bc^2x_0}{bz_0+cy_0},\n\\]\n\n\\[\nw_2=\\frac{b^2cx_0}{cy_0+bz_0}.\n\\]\n\nThe equation of the radical axis is \\(ux+vy+wz=0\\) with \\(u=u_1-u_2\\), \\(v=v_1-v_2\\), and \\(w=w_1-w_2\\). We want to consider the intersection of this line with line \\(\\overline{BC}\\), so set \\(x=0\\). The equation reduces to \\(vy+wz=0\\). We see that \\(v=\\frac{2bc^3x_0y_0}{b^2z_0^2-c^2y_0^2}\\) and \\(w=\\frac{2b^3cx_0z_0}{c^2y_0^2-b^2z_0^2}\\), so\n\n\\[\n\\frac{v}{w}=\\frac{b^2z_0}{c^2y_0},\n\\]\n\nwhich is the required condition for \\(\\overline{AD}\\) and \\(\\overline{AE}\\) to be isogonal.\n\n~MathIsFun286" ]
USAMO-2024-1
https://artofproblemsolving.com/wiki/index.php/2024_USAMO_Problems/Problem_1
Find all integers \(n \geq 3\) such that the following property holds: if we list the divisors of \(n !\) in increasing order as \(1=d_1<d_2<\cdots<d_k=n!\), then we have \[ d_2-d_1 \leq d_3-d_2 \leq \cdots \leq d_k-d_{k-1} . \]
[ "We can start by verifying that \\(n=3\\) and \\(n=4\\) work by listing out the factors of \\(3!\\) and \\(4!\\). We can also see that \\(n=5\\) does not work because the terms \\(15, 20\\), and \\(24\\) are consecutive factors of \\(5!\\). Also, \\(n=6\\) does not work because the terms \\(6, 8\\), and \\(9\\) appear consecutively in the factors of \\(6!\\).\n\nNote that if we have a prime number \\(p>n\\) and an integer \\(k>p\\) such that both \\(k\\) and \\(k+1\\) are factors of \\(n!\\), then the condition cannot be satisfied.\n\nIf \\(n\\geq7\\) is odd, then \\((2)(\\frac{n-1}{2})(n-1)=n^2-2n+1\\) is a factor of \\(n!\\). Also, \\((n-2)(n)=n^2-2n\\) is a factor of \\(n!\\). Since \\(2n<n^2-2n\\) for all \\(n\\geq7\\), we can use Bertrand's Postulate to show that there is at least one prime number \\(p\\) such that \\(n<p<n^2-2n\\). Since we have two consecutive factors of \\(n!\\) and a prime number between the smaller of these factors and \\(n\\), the condition will not be satisfied for all odd \\(n\\geq7\\).\n\nIf \\(n\\geq8\\) is even, then \\((2)(\\frac{n-2}{2})(n-2)=n^2-4n+4\\) is a factor of \\(n!\\). Also, \\((n-3)(n-1)=n^2-4n+3\\) is a factor of \\(n!\\). Since \\(2n<n^2-4n+3\\) for all \\(n\\geq8\\), we can use Bertrand's Postulate again to show that there is at least one prime number \\(p\\) such that \\(n<p<n^2-4n+3\\). Since we have two consecutive factors of \\(n!\\) and a prime number between the smaller of these factors and \\(n\\), the condition will not be satisfied for all even \\(n\\geq8\\).\n\nTherefore, the only numbers that work are \\(n=3\\) and \\(n=4\\).\n\n~alexanderruan\n\nAs listed above, \\(n=3\\) and \\(n=4\\) work but \\(n=5\\) and \\(n=6\\) don't. Assume \\(n>6\\), let \\(p\\) be the smallest positive integer that doesn't divide \\(n!\\). It is easy to see \\(p\\) is the smallest prime greater than \\(n\\). It is easy to see that both \\(p-1\\) and \\(p+1\\) are divisors of \\(n!\\) Let \\(q\\) be the smallest positive integer greater than \\(p\\) that is 2 mod 6. We see that either \\(q=p+3\\) or \\(q=p+1\\) but \\(q\\), \\(q+1\\), \\(q+2\\) must all be divisors of \\(n!\\) giving a difference of 1 after a difference of 2. Therefore, all \\(n>6\\) fail. ~mathophobia", "We claim only \\(n = 3\\) and \\(n = 4\\) are the only two solutions. First, it is clear that both solutions work.\n\nNext, we claim that \\(n < 5\\). For \\(n \\geq 5\\), let \\(x\\) be the smallest \\(x\\) such that \\(x+1\\) is not a factor of \\(n!\\). Let the smallest factor larger than \\(x\\) be \\(x+k\\).\n\nNow we consider \\(\\frac{n!}{x-1}\\), \\(\\frac{n!}{x}\\) and \\(\\frac{n!}{x+k}\\). Since \\(\\frac{n!}{x-1} > \\frac{n!}{x} > \\frac{n!}{x+k}\\), if \\(n\\) were to satisfy the conditions, then \\(\\frac{n!}{x-1}-\\frac{n!}{x} \\geq \\frac{n!}{x} - \\frac{n!}{x+k}\\). However, note that this is not true for \\(x \\geq 5\\) and \\(k > 1\\).\n\nNote that the inequality is equivalent to showing \\(\\frac{1}{x(x-1)} \\geq \\frac{k}{x(x+k)}\\), which simplifies to \\(x+k \\geq kx-k\\), or \\(\\frac{x}{x-2} \\geq k \\geq 2\\). This implies \\(x \\geq 2x-4\\), \\(x \\leq 4\\), a contradiction, since the set of numbers \\(\\{1, 2, 3, 4, 5\\}\\) are all factors of \\(n!\\), and the value of \\(x\\) must exist. Hence, no solutions for \\(n \\geq 5\\)." ]
USAMO-2024-2
https://artofproblemsolving.com/wiki/index.php/2024_USAMO_Problems/Problem_2
Let \(S_1, S_2, \ldots, S_{100}\) be finite sets of integers whose intersection is not empty. For each non-empty \(T \subseteq\left\{S_1, S_2, \ldots, S_{100}\right\}\), the size of the intersection of the sets in \(T\) is a multiple of the number of sets in \(T\). What is the least possible number of elements that are in at least 50 sets?
[ "Let's determine the smallest possible number of elements that appear in at least 50 of the sets.\n\nFirst, we establish some notation. For any subset \\(T \\subseteq \\{S_1, S_2, \\ldots, S_{100}\\}\\), let \\(\\cap T\\) denote the intersection of all sets in \\(T\\). By the problem statement, \\(|\\cap T|\\) is divisible by \\(|T|\\).\n\nWe'll use a binary encoding approach. For each element \\(x\\) in our universe, define its \"signature\" as a binary string of length 100, where the \\(i\\)-th bit is 1 if \\(x \\in S_i\\) and 0 otherwise.\n\nFor any signature \\(v\\) with exactly \\(k\\) ones, corresponding elements must appear in exactly \\(k\\) sets. The problem condition requires that for any subset of sets corresponding to positions where 1's appear in some signature \\(u\\), the number of elements having signature \\(v\\) that contains all 1's from \\(u\\) must be divisible by \\(|u|\\).\n\nLet's denote by \\(f(v)\\) the number of elements with signature \\(v\\). The problem condition translates to: for any signature \\(u\\), \\(|u|\\) divides \\(\\sum_{v \\supseteq u} f(v)\\).\n\nKey claim: For any signature \\(u\\) with \\(|u| = 50\\), we need \\(f(u) \\geq 50\\).\n\nProof: By the divisibility condition, \\(\\sum_{v \\supseteq u} f(v)\\) must be divisible by 50. For any \\(v\\) strictly containing \\(u\\), if we sum over all such \\(u\\) of size 50, each \\(f(v)\\) with \\(|v| > 50\\) gets counted multiple times. To satisfy the divisibility condition for each individual \\(u\\), we need \\(f(u) \\geq 50\\) for each signature \\(u\\) with exactly 50 ones.\n\nSince there are \\(\\binom{100}{50}\\) different ways to choose 50 positions for ones in the signature, and each such signature must correspond to at least 50 elements, the minimum number of elements appearing in at least 50 sets is \\(50 \\binom{100}{50}\\).\n\nTherefore, the answer is \\(\\boxed{50 \\binom{100}{50}}\\).\n\n~ brandonyee" ]
USAMO-2024-3
https://artofproblemsolving.com/wiki/index.php/2024_USAMO_Problems/Problem_3
Let \(m\) be a positive integer. A triangulation of a polygon is \(m\)-balanced if its triangles can be colored with \(m\) colors in such a way that the sum of the areas of all triangles of the same color is the same for each of the \(m\) colors. Find all positive integers \(n\) for which there exists an \(m\)-balanced triangulation of a regular \(n\)-gon. Note: A triangulation of a convex polygon \(\mathcal{P}\) with \(n \geq 3\) sides is any partitioning of \(\mathcal{P}\) into \(n-2\) triangles by \(n-3\) diagonals of \(\mathcal{P}\) that do not intersect in the polygon's interior.
[ "The answer is if and only if \\(m\\) is a proper divisor of \\(n\\).\n\nWe represent the vertices of the regular \\(n\\)-gon using complex numbers. Let \\(\\omega = \\exp(2\\pi i/n)\\) be a primitive \\(n\\)-th root of unity, and label the vertices as \\(1, \\omega, \\omega^2, \\ldots, \\omega^{n-1}\\).\n\nKey Lemma:\n\nThe triangle with vertices \\(\\omega^k, \\omega^{k+a}, \\omega^{k+b}\\) has signed area:\n\n\\[\nT(a, b) := \\frac{(\\omega^a - 1)(\\omega^b - 1)(\\omega^{-b} - \\omega^{-a})}{4i}\n\\]\n\nProof of Lemma:\n\nWe can rotate by \\(\\omega^{-k}\\) to assume without loss of generality that \\(k = 0\\). Then we apply the complex shoelace formula to the triangle with vertices \\(1, \\omega^a, \\omega^b\\):\n\n\\[\n\\frac{i}{4} \\det \\begin{bmatrix} 1 & 1 & 1 \\\\ \\omega^a & \\omega^{-a} & 1 \\\\ \\omega^b & \\omega^{-b} & 1 \\end{bmatrix} = \\frac{i}{4} \\det \\begin{bmatrix} 0 & 0 & 1 \\\\ \\omega^a - 1 & \\omega^{-a} - 1 & 1 \\\\ \\omega^b - 1 & \\omega^{-b} - 1 & 1 \\end{bmatrix}\n\\]\n\nwhich simplifies to the expression for \\(T(a, b)\\) above.\n\nConstruction (Sufficiency):\n\nTo construct a valid triangulation and coloring when \\(m\\) divides \\(n\\), we draw all the diagonals from vertex 1, and then color the resulting triangles with the \\(m\\) colors in cyclic order.\n\nFor example, when \\(n = 9\\) and \\(m = 3\\), we get a coloring with red, green, blue as shown in the image, repeating cyclically.\n\nTo prove this works, we fix a residue \\(r \\mod m\\) corresponding to one of the colors. Then the total area of triangles with this color is:\n\n\\[\n\\sum_{\\substack{0 \\leq j < n \\\\ j \\equiv r \\mod m}} \\text{Area}(\\omega^0, \\omega^j, \\omega^{j+1}) = \\sum_{\\substack{0 \\leq j < n \\\\ j \\equiv r \\mod m}} T(j, j+1)\n\\]\n\nAfter algebraic manipulation:\n\n\\[\n= \\frac{\\omega - 1}{4i} \\left(\\frac{n}{m}(1 + \\omega^{-1}) + \\sum_{\\substack{0 \\leq j < n \\\\ j \\equiv r \\mod m}} (\\omega^{-1-j} - \\omega^j)\\right)\n\\]\n\nWhen \\(m\\) divides \\(n\\), we can show that the inner sum vanishes, and the total area of each color equals:\n\n\\[\n\\sum_{\\substack{0 \\leq j < n \\\\ j \\equiv r \\mod m}} \\text{Area}(\\omega^0, \\omega^j, \\omega^{j+1}) = \\frac{n}{m} \\cdot \\frac{\\omega - \\omega^{-1}}{4i}\n\\]\n\nSince the right-hand side does not depend on the residue \\(r\\), this shows all colors have equal area.\n\nNecessity:\n\nTo prove that \\(m\\) must divide \\(n\\), we note that if there were a valid triangulation and coloring, the total area of each color would equal:\n\n\\[\nS := \\frac{n}{m} \\cdot \\frac{\\omega - \\omega^{-1}}{4i}\n\\]\n\nWe then prove the following claim:\n\nThe number \\(4i \\cdot S\\) is not an algebraic integer when \\(m \\nmid n\\).\n\nThis can be shown using the fact that \\(K = \\mathbb{Q}(\\omega)\\) is a number field with ring of integers \\(\\mathcal{O}_K = \\mathbb{Z}[\\omega]\\). We demonstrate that:\n\n\\[\n\\omega \\cdot 4i \\cdot S = \\frac{n}{m}\\omega^2 - \\frac{n}{m}\n\\]\n\nis not an algebraic integer when \\(m\\) doesn't divide \\(n\\) because \\(\\frac{n}{m} \\not\\in \\mathbb{Z}\\).\n\nHowever, for any triangle in our construction, \\(4i \\cdot T(a,b)\\) is always an algebraic integer. Since a finite sum of algebraic integers is also an algebraic integer, the sum of expressions of the form \\(4i \\cdot T(a,b)\\) can never equal \\(4i \\cdot S\\) unless \\(m\\) divides \\(n\\).\n\nWe can also show this directly by noting that \\(T(a,b)\\) is always divisible by \\(\\frac{\\omega - \\omega^{-1}}{4i}\\), which means we need \\(\\frac{n}{m}\\) to be an algebraic integer. A rational number is an algebraic integer if and only if it's a rational integer, so \\(\\frac{n}{m}\\) is an algebraic integer exactly when \\(m\\) divides \\(n\\).\n\nTherefore, a valid triangulation and coloring exists if and only if \\(m\\) is a proper divisor of \\(n\\).\n\n~ brandonyee" ]
USAMO-2024-4
https://artofproblemsolving.com/wiki/index.php/2024_USAMO_Problems/Problem_4
Let \(m\) and \(n\) be positive integers. A circular necklace contains \(m n\) beads, each either red or blue. It turned out that no matter how the necklace was cut into \(m\) blocks of \(n\) consecutive beads, each block had a distinct number of red beads. Determine, with proof, all possible values of the ordered pair \((m, n)\).
[ "We need to determine all possible positive integer pairs \\((m, n)\\) such that there exists a circular necklace of \\(mn\\) beads, each colored red or blue, satisfying the following condition:\n\nNo matter how the necklace is cut into \\(m\\) blocks of \\(n\\) consecutive beads, each block has a distinct number of red beads.\n\nNecessary Condition:\n\n1. Maximum Possible Distinct Counts: In a block of \\(n\\) beads, the number of red beads can range from \\(0\\) to \\(n\\). Therefore, there are \\(n + 1\\) possible distinct counts of red beads in a block. Since we have \\(m\\) blocks, the maximum number of distinct counts must be at least \\(m\\). Thus, we must have:\n\n\\[\nm \\leq n + 1\n\\]\n\nSufficient Construction:\n\nWe will show that for all positive integers \\(m\\) and \\(n\\) satisfying \\(m \\leq n + 1\\), such a necklace exists.\n\n1. Construct Blocks: Create \\(m\\) blocks, each containing \\(n\\) beads. Assign to each block a unique number of red beads, ranging from \\(0\\) to \\(m - 1\\).\n\n2. Design the Necklace: Arrange these \\(m\\) blocks in a fixed order to form the necklace. Since the necklace is circular, cutting it at different points results in cyclic permutations of the blocks.\n\n3. Verification: In any cut, the sequence of blocks (and thus the counts of red beads) is a cyclic shift of the original sequence. Therefore, in each partition, the blocks will have distinct numbers of red beads.\n\nExample:\n\nLet's construct a necklace for \\(m = 3\\) and \\(n = 2\\):\n\nBlocks: Block 1: \\(0\\) red beads (BB) Block 2: \\(1\\) red bead (RB) Block 3: \\(2\\) red beads (RR)\n\nNecklace Arrangement: Place the blocks in order: BB-RB-RR.\n\nVerification: Any cut of the necklace into \\(3\\) blocks of \\(2\\) beads will have blocks with red bead counts of \\(0\\), \\(1\\), and \\(2\\).\n\nConclusion:\n\nAll ordered pairs \\((m, n)\\) where \\(m \\leq n + 1\\) satisfy the condition. Therefore, the possible values of \\((m, n)\\) are all positive integers such that \\(m \\leq n + 1\\).\n\nFinal Answer:\n\nExactly all positive integers \\(m\\) and \\(n\\) with \\(m \\leq n + 1\\); these are all possible ordered pairs \\((m, n)\\).\n\n~Athmyx" ]
USAMO-2024-5
https://artofproblemsolving.com/wiki/index.php/2024_USAMO_Problems/Problem_5
Point \(D\) is selected inside acute triangle \(ABC\) so that \(\angle DAC=\angle ACB\) and \(\angle BDC=90^\circ+\angle BAC\). Point \(E\) is chosen on ray \(BD\) so that \(AE=EC\). Let \(M\) be the midpoint of \(BC\). Show that line \(AB\) is tangent to the circumcircle of triangle \(BEM\).
[ "Let \\(\\angle DBT = \\alpha\\) and \\(\\angle BEM = \\beta\\). Extend AD intersects BC at point T, then TC = TA, TE is perpendicular to AC\n\nThus, AB is the tangent of the circle BEM\n\nThen the question is equivalent as the \\(\\angle ABT\\) is the auxillary angle of \\(\\angle BEM\\).\n\nontinued" ]
USAMO-2024-6
https://artofproblemsolving.com/wiki/index.php/2024_USAMO_Problems/Problem_6
Continued Let \(n>2\) be an integer and let \(\ell \in\{1,2, \ldots, n\}\). A collection \(A_1, \ldots, A_k\) of (not necessarily distinct) subsets of \(\{1,2, \ldots, n\}\) is called \(\ell\)-large if \(\left|A_i\right| \geq \ell\) for all \(1 \leq i \leq k\). Find, in terms of \(n\) and \(\ell\), the largest real number \(c\) such that the inequality \[ \sum_{i=1}^k \sum_{j=1}^k x_i x_j \frac{\left|A_i \cap A_j\right|^2}{\left|A_i\right| \cdot\left|A_j\right|} \geq c\left(\sum_{i=1}^k x_i\right)^2 \] holds for all positive integers \(k\), all nonnegative real numbers \(x_1, \ldots, x_k\), and all \(\ell\)-large collections \(A_1, \ldots, A_k\) of subsets of \(\{1,2, \ldots, n\}\). Note: For a finite set \(S,|S|\) denotes the number of elements in \(S\).
[]
USAMO-2025-1
https://artofproblemsolving.com/wiki/index.php/2025_USAMO_Problems/Problem_1
Let \(k\) and \(d\) be positive integers. Prove that there exists a positive integer \(N\) such that for every odd integer \(n>N\), the digits in the base-\(2n\) representation of \(n^k\) are all greater than \(d\).
[ "We define a remainder operation \\(\\,a \\bmod b\\,\\) to be the remainder when \\(a\\) is divided by \\(b\\). Also, let \\(\\lfloor x\\rfloor\\) be the usual floor function.\n\nBase-\\((2n)\\) Representation:\n\n\\[\nn^k \\;=\\; a_{k-1}\\,(2n)^{k-1} \\;+\\; a_{k-2}\\,(2n)^{k-2} \\;+\\;\\dots\\;+\\;a_1\\,(2n)\\;+\\;a_0,\n\\]\n\nwhere each \\(a_i\\) satisfies \\(0 \\le a_i < 2n.\\) Hence, the base-\\((2n)\\) representation of \\(n^k\\) is \\(a_{k-1}\\,a_{k-2}\\,\\dots\\,a_1\\,a_0.\\)\n\nLeading Digit:\n\n\\[\na_{k-1} \\;=\\; \\left\\lfloor \\dfrac{n^k}{(2n)^{k-1}} \\right\\rfloor \\;=\\; \\left\\lfloor \\dfrac{n}{2^{k-1}} \\right\\rfloor.\n\\]\n\nGeneral Digit Formula: For \\(0 \\le i < k,\\)\n\n\\[\na_i \\;=\\; \\left\\lfloor \\dfrac{\\,n^k \\bmod (2n)^{\\,i+1}}{(2n)^i} \\right\\rfloor.\n\\]\n\nBecause \\(n\\) is odd, one can show\n\n\\[\nn^k \\bmod (2n)^{\\,i+1} \\;\\ge\\; n^{\\,i+1},\n\\]\n\nwhich implies\n\n\\[\na_i \\;\\ge\\; \\left\\lfloor \\dfrac{n^{\\,i+1}}{(2n)^i} \\right\\rfloor \\;=\\; \\left\\lfloor \\dfrac{n}{2^i} \\right\\rfloor \\;\\ge\\; 2^{\\,k-1-i} \\left\\lfloor \\dfrac{n}{2^{\\,k-1}} \\right\\rfloor.\n\\]\n\nAs \\(n\\) grows large, \\(\\bigl\\lfloor n / 2^{\\,k-1}\\bigr\\rfloor\\) becomes arbitrarily big, so each digit \\(a_i\\) eventually exceeds any fixed \\(d.\\) Hence there exists an integer \\(N\\) such that for all odd \\(n > N,\\) the digits \\(a_i\\) in the base-\\((2n)\\) representation of \\(n^k\\) are all \\(> d.\\) This completes the proof. ~Rex" ]
USAMO-2025-2
https://artofproblemsolving.com/wiki/index.php/2025_USAMO_Problems/Problem_2
Let \(n\) and \(k\) be positive integers with \(k<n\). Let \(P(x)\) be a polynomial of degree \(n\) with real coefficients, nonzero constant term, and no repeated roots. Suppose that for any real numbers \(a_0, a_1, \dots, a_k\) such that the polynomial \(a_kx^k+\cdots+a_1x+a_0\) divides \(P(x)\), the product \(a_0a_1\cdots a_k\) is zero. Prove that \(P(x)\) has a nonreal root.
[ "We proceed by contradiction. Assume that all roots of \\(P(x)\\) are real. Let the distinct roots be \\(r_1, r_2, \\ldots, r_n\\), all nonzero since the constant term is nonzero.\n\nConsider any subset of \\(k\\) roots \\(\\{r_{i_1}, r_{i_2}, \\ldots, r_{i_k}\\}\\) and form the polynomial:\n\n\\[\nQ(x) = \\prod_{j=1}^k (x - r_{i_j}) =a_{k} x^k + a_{k-1}x^{k-1} + \\cdots + a_0\n\\]\n\nBy Vieta's formulas:\n\n\\(a_0 = (-1)^k \\prod_{j=1}^k r_{i_j} \\neq 0\\) \\(a_{k-1} = -\\sum_{j=1}^k r_{i_j}\\) \\(a_{k-2} = \\sum_{1\\leq m<n\\leq k} r_{i_m}r_{i_n}\\)\n\nThe given condition requires that \\(a_0a_1\\cdots a_k = 0\\). Since \\(a_0 \\neq 0\\), at least one other coefficient must be zero.\n\nCase \\(k=2\\): For any pair of roots \\((r_i, r_j)\\), we have:\n\n\\[\nQ(x) = x^2 - (r_i+r_j)x + r_ir_j\n\\]\n\nThe condition implies \\(-r_ir_j(r_i+r_j) = 0\\), so \\(r_i + r_j = 0\\) for all pairs. But with \\(n \\geq 3\\), considering three roots \\(r_1, r_2, r_3\\) gives:\n\n\\[\nr_1 + r_2 = 0 \\quad \\text{and} \\quad r_1 + r_3 = 0 \\implies r_2 = r_3\n\\]\n\ncontradicting distinct roots. In General \\(k\\): For any \\(k\\) roots, some symmetric sum must be zero. For \\(k=3\\), this would require:\n\n\\[\nr_i + r_j + r_m = 0 \\quad \\text{for all triples}\n\\]\n\nwhich leads to contradictions for any \\(n \\geq 4\\) as it would force roots to be equal.\n\nThus, our initial assumption is false, and \\(P(x)\\) must have at least one nonreal root.~Jonathan" ]
USAMO-2025-3
https://artofproblemsolving.com/wiki/index.php/2025_USAMO_Problems/Problem_3
Alice the architect and Bob the builder play a game. First, Alice chooses two points \(P\) and \(Q\) in the plane and a subset \(\mathcal{S}\) of the plane, which are announced to Bob. Next, Bob marks infinitely many points in the plane, designating each a city. He may not place two cities within distance at most one unit of each other, and no three cities he places may be collinear. Finally, roads are constructed between the cities as follows: for each pair \(A,\,B\) of cities, they are connected with a road along the line segment \(AB\) if and only if the following condition holds: For every city \(C\) distinct from \(A\) and \(B\), there exists \(R\in\mathcal{S}\) such that \(\triangle PQR\) is directly similar to either \(\triangle ABC\) or \(\triangle BAC\). Alice wins the game if (i) the resulting roads allow for travel between any pair of cities via a finite sequence of roads and (ii) no two roads cross. Otherwise, Bob wins. Determine, with proof, which player has a winning strategy. Note: \(\triangle UVW\) is directly similar to \(\triangle XYZ\) if there exists a sequence of rotations, translations, and dilations sending \(U\) to \(X\), \(V\) to \(Y\), and \(W\) to \(Z\).
[ "https://artofproblemsolving.com/wiki/index.php/File:2025_USAMO_PROBLEM_3.jpeg\n\nThe answer is that Alice wins. Let's define a Bob-set V to be a set of points in the plane with no three collinear and with all distances at least 1. The point of the problem is to prove the following fact. Claim — Given a Bob-set V ⊆ R 2 , consider the Bob-graph with vertex set V defined as follows: draw edge ab if and only if the disk with diameter ab contains no other points of V on or inside it. Then the Bob-graph is (i) connected, and (ii) planar. Proving this claim shows that Alice wins since Alice can specify S to be the set of points outside the disk of diameter P Q. Proof that every Bob-graph is connected. Assume for contradiction the graph is disconnected. Let p and q be two points in different connected components. Since pq is not an edge, there exists a third point r inside the disk with diameter pq. Hence, r is in a different connected component from at least one of p or q — let's say point p. Then we repeat the same argument on the disk with diameter pr to find a new point s, non-adjacent to either p or r. See the figure below, where the X'ed out dashed edges indicate points which are not only non-adjacent but in different connected components.\n\nIn this way we generate an infinite sequence of distances δ1, δ2, δ3, . . . among the non-edges in the picture above. By the “Pythagorean theorem” (or really the inequality for it), we have δi2 ≤ δi-12- 1 and this eventually generates a contradiction for large i, since we get 0 ≤ δi2≤ δi2- (i - 1). Proof that every Bob-graph is planar. Assume for contradiction edges ac and bd meet, meaning abcd is a convex quadrilateral. WLOG assume ∠bad ≥ 90◦ (each quadrilateral has an angle at least 90◦ ). Then the disk with diameter bd contains a, contradiction.\n\nRemark. In real life, the Bob-graph is actually called the Gabriel graph. Note that we never require the Bob-set to be infinite; the solution works unchanged for finite Bob-sets. However, there are approaches that work for finite Bob-sets that don't work for infinite sets, such as the relative neighbor graph, in which one joins a and b iff there is no c such that d(a, b) ≤ max{d(a, c), d(b, c)}. In other words, edges are blocked by triangles where ab is the longest edge (rather than by triangles where ab is the longest edge of a right or obtuse triangle as in the Gabriel graph). The relative neighbor graph has fewer edges than the Gabriel graph, so it is planar too. When the Bob-set is finite, the relative distance graph is still connected. The same argument above works where the distances now satisfy δ1 > δ2 > . . . instead, and since there are finitely many distances one arrives at a contradiction.\n\nHowever for infinite Bob-sets the descending condition is insufficient, and connectedness actually fails altogether. A counterexample is to start by taking An ≈ (2n, 0) and Bn ≈ (2n + 1, √ 3) for all n ≥ 1, then perturb all the points slightly so that B1A1 > A1A2 > A2B1 > B1B2 > B2A2 > A2A3 > A3B2 > B2B3 > B3A3 > · · · . In that case, {An} and {Bn} will be disconnected from each other: none of the edges AnBn or B n A n+1 are formed. In this case the relative neighbor graph consists of the edges A1A2A3A4 · · · and B1B2B3B4 · · · . That's why for the present problem, the inequality δi2 ≤ δi-12- 1 plays such an important role, because it causes the (squared) distances to decrease appreciably enough to give the final contradiction." ]
USAMO-2025-4
https://artofproblemsolving.com/wiki/index.php/2025_USAMO_Problems/Problem_4
Let \(H\) be the orthocenter of acute triangle \(ABC\), let \(F\) be the foot of the altitude from \(C\) to \(AB\), and let \(P\) be the reflection of \(H\) across \(BC\). Suppose that the circumcircle of triangle \(FAP\) intersects line \(BC\) at two distinct points \(X\) and \(Y\). Prove that \(C\) is the midpoint of \(XY\).
[ "Let AP intersects BC at D. Extend FC to the point E on the circumcircle \\(\\omega\\) of \\(FAP\\). Since \\(H\\) is the orthocenter of \\(\\Delta ABC\\), we know that \\(HD = DP\\) or \\(HP = 2HD\\), and \\(AH \\cdot HD = CH \\cdot HF\\). Next we use the power of H in \\(\\omega\\): \\(AH \\cdot HP = CH \\cdot HE\\). These relations imply that \\(HE = 2HF\\).\n\nHence \\(C, D\\) are midpoints of \\(HE, HP\\) respectively. By midline theorem, \\(CD // EP\\). Since \\(AD \\perp CD\\), we have \\(AD \\perp EP\\). This implies that \\(\\angle APE = 90^{\\circ}\\). Consequently, \\(AE\\) is the diameter of \\(\\omega\\). Let \\(G\\) be the midpoint of \\(AE\\) which is also the center of \\(\\omega\\). \\(G,C\\) are midpoints of \\(AE, EH\\) respectively. By the midline theorem again, we have \\(GC//AH\\), consequently, \\(GC \\perp BC\\). This implies that \\(GC\\) is the perpendicular bisector of the chord \\(XY\\) hence \\(C\\) is the midpoint of \\(XY\\). ~ Dr. Shi davincimath.com", "Denote \\(O_1\\) as the center of \\((ABC)\\), \\(O_2\\) as the center of \\(FAP\\), \\(K\\) as the midpoint of \\(AF\\), \\(M\\) as the midpoint of \\(AC\\), and \\(N\\) as the midpoint of \\(AP\\). It suffices to show that \\(\\angle{O_2CB}=90\\).\n\nClaim: \\(O_1MO_2C\\) is cyclic.\n\nProof: Since \\(AK=FK\\) and \\(AM=MC\\), KM is a midline of \\(\\triangle{AFC}\\) and \\(KM\\parallel FC\\). \\(KO_2\\parallel FC\\) as well since \\(\\angle{AKO_2}=\\angle{AFC}=90\\), so \\(M\\) lies on \\(KO_2\\). Next, note that \\(P\\) lies on \\((ABC)\\), so the perpendicular bisector of \\(AP\\) through \\(N\\) passes through \\(O_1\\). In other words, \\(N, O_1\\), and \\(O_2\\) are collinear. Since \\(NO_2\\) and \\(BC\\) are both perpendicular to \\(AP\\), it follows that they are parallel. Since \\(KO_2\\parallel FC\\) and \\(NO_2\\parallel BC\\), then \\(\\angle KO_2N=\\angle{FCB}\\). Finally, we have that\n\n\\[\n\\angle{MO_2O_1}=\\angle{KO_2N}=\\angle{FCB}=90-B=\\angle{MCO_1},\n\\]\n\nand thus \\(O_1MO_2C\\) is cyclic. It follows that \\(\\angle O_1O_2C=\\angle{O_1MC}=90\\), so \\(\\angle{O_2CB}=180-\\angle{O_1MC}=90\\), as desired.\n\n-mop", "Connect \\(HP\\) and have \\(HP\\) intersect \\(XY\\) at \\(W\\). Also extend \\(FC\\) past point \\(C\\) and have it intersect with the circle at point \\(D\\).\n\nSince \\(P\\) is the reflection of \\(H\\) over \\(BC\\), we know that \\(HP\\perp XY\\). Since \\(H\\) is the orthocenter, we can draw the altitude and tell that \\(A\\), \\(H\\), and \\(P\\) are collinear. We know \\(m\\angle{AFH} = m\\angle{CWH}=90^{\\circ}\\) and \\(m\\angle{FHA} = m\\angle{WHC}\\), so \\(\\triangle{AHF} \\sim \\triangle{CHW}\\) by AA, so \\(m\\angle{FAH} = m\\angle{WCH}\\).\n\n\\(m\\angle{FAP} = \\frac{1}{2}m\\overarc{FP} = \\frac{1}{2}(m\\overarc{FX} + m\\overarc{XP})\\) and \\(m\\angle{FCX} = \\frac{1}{2}(m\\overarc{FX} + m\\overarc{DY})\\). From this, we can tell that \\(m\\overarc{XP} = m\\overarc{DY}\\). Therefore, \\(m\\overarc{XD} = m\\overarc{PY}\\) and \\(XD = PY\\).\n\nIf we connect \\(HY\\), we can tell that that \\(PY = HY\\) due to \\(P\\) being the reflection of \\(H\\) and \\(WY\\) being perpendicular to \\(HP\\), so \\(XD = HY\\). In addition, \\(m\\angle{HYW} = m\\angle{PYW} = \\frac{1}{2} m\\overarc{XP} = \\frac{1}{2} m\\overarc{DY} = m\\angle{YXD}\\). Also, \\(m\\angle{HCY} = m\\angle{XCD}\\) because they are vertical angles.\n\nSo, \\(\\triangle{HCY} \\cong \\triangle{XCD}\\) because of SAA. From this we can conclude that \\(XC = CY\\), so \\(C\\) is the midpoint of \\(XY\\).", "Let \\(D\\) be the foot of the altitude from \\(A\\) to \\(BC.\\) By Power of a Point, we have\n\n\\[\nBF\\cdot BA = BX \\cdot BY = (BC-CX)(BC+CY) = BC^2 + BC(CY-CX) - CX\\cdot CY\n\\]\n\nand\n\n\\[\nDB\\cdot DC = DX\\cdot DY = (CX-CD)(CD+CY) = CX\\cdot CY - CD(CY-CX) - CD^2.\n\\]\n\nAdding, we get\n\n\\[\nBF \\cdot BA + DB\\cdot DC = BC^2-DC^2 + (BC-CD)(CY-CX).\n\\]\n\nIt is well known that the reflection of \\(H\\) over \\(AB,\\) which we denote by \\(P,\\) lies on \\((ABC).\\) Then, let \\(\\angle{BAP} = \\angle{BCP} = \\angle{BCH} = \\theta.\\) We have\n\n\\[\nBF\\cdot BA + DB\\cdot DC = (BC\\sin\\theta)\\cdot BA + DB\\cdot DC = BC\\cdot (BA \\sin\\theta) + DB\\cdot DC = DB\\cdot (BC+DC) = (BC-DC)(BC+DC) = BC^2-DC^2.\n\\]\n\nThus, \\((BC-CD)(CY-CX)=0,\\) and since \\(BC\\neq CD\\) we have \\(CY=CX.\\) Hence, \\(C\\) is the midpoint of \\(XY.\\) ~TThB0501", "Let Q be the antipode of B. Claim — AHQC is a parallelogram, and AP CQ is an isosceles trapezoid.\n\nProof. As AH ⊥ BC ⊥ CQ and CF ⊥ AB ⊥ AQ. Let M be the midpoint of QC.\n\nClaim — Point M is the circumcenter of triangle AFP.\n\nProof. It's clear that MA = MP from the isosceles trapezoid.\n\nAs for MA = MF, let N denote the midpoint of AF; then MN is a midline of the parallelogram, so MN ⊥ AF. Since CM ⊥ BC and M is the center of (AF P), it follows CX = CY .", "Quick angle chasing gives \\(\\angle FAP = \\angle BCF = \\angle BCP = \\alpha\\). Let \\(O\\) be the circumcenter of \\(\\triangle AFP\\).\n\nThus \\(\\angle FOP= 2 \\angle FAP = 2\\alpha\\) (because O and A lie on the same side of segment \\(AF\\)).\n\nAs \\(\\angle FOP = \\angle FCP = 2 \\alpha\\), the quadrilateral FOCP is cyclic.\n\nObserve that \\(2\\angle FPO = 180-2\\alpha\\), so \\(\\angle FPO = 90^{\\circ} - \\alpha\\).\n\nFrom the properties of cyclic quadrilaterals, \\(\\angle FCO = \\angle FPO\\)\n\nThus \\(\\angle BCO = \\angle BCF + \\angle FCO = \\alpha + (90^{\\circ} - \\alpha) =90^{\\circ}\\). Therefore \\(OC\\) is perpendicular to chord \\(XY\\), \\(XC=CY\\).", "Let the line perpendicular to \\(BC\\) and going through \\(C\\) be line \\(l\\). Let the midpoint of \\(AF\\) be \\(M\\), and let the line perpendicular to \\(AF\\) and going through \\(M\\) be line \\(n\\). Let \\(O\\) be the intersection of \\(l\\) and \\(n\\), and let \\(D\\) be the intersections of \\(n\\) and \\(AH\\).\n\nBecause \\(AH\\parallel l\\) and \\(CF\\parallel n\\), \\(CODH\\) is a parallelogram. From this, we get \\(OC=DH\\). Looking at triangle \\(\\triangle{FAP}\\), because \\(n\\parallel HF\\), and \\(M\\) is the midpoint of \\(AF\\), it is clear that \\(D\\) is the midpoint of \\(AH\\).\n\nLet the intersection of \\(HP\\) and \\(BC\\) be \\(Z\\). \\(PZ=ZH\\) because \\(P\\) is a reflection of \\(H\\) across \\(BC\\). \\(AP=AH+HP=2DH+2PZ\\)\n\nSubstituting the equation from earlier, we get \\(AP=2OC+2PZ\\)\n\nDraw a line perpendicular to \\(OC\\) starting from \\(O\\). Let the intersection of that line and \\(AP\\) be \\(E\\). Because \\(BC\\) is perpendicular to \\(OC\\) which is perpendicular to \\(OE\\), \\(BC\\parallel OE\\). By the same logic, \\(AP\\parallel OC\\). This means \\(OCZE\\) is a parallelogram (specifically a rectangle). Therefore \\(EZ=OC\\). \\(EP=EZ+PZ\\) and by substituting, we get \\(EP=OC+PZ\\), and because \\(AP=2OC+2PZ\\), that means \\(2EP=AP\\), so \\(E\\) is a midpoint of \\(AP\\). This means \\(OE\\) is the perpendicular bisector of \\(AP\\), and because \\(AP\\) is a chord of the circumcircle of \\(\\triangle{FAP}\\), \\(OE\\) goes through the center of the circle. By the same logic, \\(n\\) also goes through the center of the circle, since it is the perpendicular bisector of \\(AF\\). This means the center of the circle is \\(O\\), because it is the only point on both \\(OE\\) and \\(n\\). Because \\(O\\) is on line \\(OC\\) and line \\(OC\\) is perpendicular to line \\(BC\\), line \\(OC\\) must perpendicularly bisect the chord of circle \\(O\\) containing \\(C\\) that lies on line \\(BC\\) (basically \\(OC\\) must perpendicularly bisect \\(XY\\)). This means \\(C\\) is the midpoint of \\(XY\\).\n\n-Nolan Lin" ]
USAMO-2025-5
https://artofproblemsolving.com/wiki/index.php/2025_USAMO_Problems/Problem_5
Determine, with proof, all positive integers \(k\) such that \[ \frac{1}{n+1} \sum_{i=0}^n \binom{n}{i}^k \] is an integer for every positive integer \(n.\)
[ "https://artofproblemsolving.com/wiki/index.php/File:2025_USAMO_PROBLEM_5_1.jpg\n\nhttps://artofproblemsolving.com/wiki/index.php/File:2025_USAMO_PROBLEM_5_2.jpg", "Let\n\n\\[\nX=\\{(A_1,\\dots,A_k):A_j\\subseteq\\{1,\\dots,n\\},\\;|A_1|=\\cdots=|A_k|\\}.\n\\]\n\nThen\n\n\\[\n|X|=\\sum_{i=0}^n\\binom{n}{i}^k.\n\\]\n\nThe cyclic group \\(C_{n+1}\\) acts on \\(\\{1,\\dots,n+1\\}\\) by “add 1 mod \\(n+1\\),” and hence diagonally on \\(X\\).\n\nBy Burnside's lemma,\n\n\\[\n|X/C_{n+1}|=\\frac1{n+1}\\sum_{g\\in C_{n+1}}|\\mathrm{Fix}(g)|.\n\\]\n\nIf \\(g\\) has order \\(d\\mid(n+1)\\) then \\(\\mathrm{Fix}(g)\\) consists of choosing each \\(A_j\\) as a union of the \\((n+1)/d\\) orbits of \\(g\\), so\n\n\\[\n|\\mathrm{Fix}(g)|=\\sum_{i=0}^n\\binom{d}{i}^k.\n\\]\n\nThus\n\n\\[\n(n+1)\\,|X/C_{n+1}|=\\sum_{d\\mid(n+1)}\\varphi\\Bigl(\\tfrac{n+1}d\\Bigr)\\, \\sum_{i=0}^n\\binom{d}{i}^k.\n\\]\n\nCase 1: \\(k=2m\\) even\n\nWe prove by induction on the divisor‐lattice of \\(n+1\\) that for every proper divisor \\(d<n+1\\),\n\n\\[\nd\\mid\\sum_{i=0}^n\\binom{d}{i}^{2m}.\n\\]\n\nThen each term \\(\\varphi\\bigl((n+1)/d\\bigr)\\sum_i\\binom{d}{i}^{2m}\\) is divisible by \\(\\varphi\\bigl((n+1)/d\\bigr)\\,d\\), and since\n\n\\[\n\\sum_{d\\mid(n+1),\\,d<n+1}\\varphi\\Bigl(\\tfrac{n+1}d\\Bigr)\\,d =(n+1)-\\varphi(1)\\cdot1 =n+1-1,\n\\]\n\nthe entire sum on the right is congruent modulo \\(n+1\\) to\n\n\\[\n\\varphi(1)\\sum_{i=0}^n\\binom{n+1}{i}^{2m} =1\\cdot\\sum_{i=0}^n\\binom{n+1}{i}^{2m}.\n\\]\n\nBut the left side \\((n+1)\\,|X/C_{n+1}|\\equiv0\\pmod{n+1}\\), so\n\n\\[\nn+1\\;\\bigm\\vert\\;\\sum_{i=0}^n\\binom{n+1}{i}^{2m}.\n\\]\n\nA re‐index \\(n+1\\mapsto n\\) yields\n\n\\[\nn+1\\;\\bigm\\vert\\;\\sum_{i=0}^n\\binom{n}{i}^{2m},\n\\]\n\nhence\n\n\\[\nA_{2m}(n)=\\frac1{n+1}\\sum_{i=0}^n\\binom{n}{i}^{2m}\\in\\mathbb Z.\n\\]\n\nCase 2: \\(k\\) odd\n\nTake \\(n=2\\). Then\n\n\\[\nA_k(2)=\\frac{\\binom21^k+\\binom22^k+\\binom23^k}{3} =\\frac{1+2^k+1}{3} =\\frac{2+2^k}{3}.\n\\]\n\nIf \\(k\\) is odd then \\(2^k\\equiv2\\pmod3\\), so \\(2+2^k\\equiv4\\equiv1\\pmod3\\), whence\n\n\\[\nA_k(2)\\notin\\mathbb Z.\n\\]\n\nConclusion\n\n\\[\nA_k(n)\\in\\mathbb Z\\text{ for all }n\\ge1\\iff k\\text{ even.}\n\\]" ]
USAMO-2025-6
https://artofproblemsolving.com/wiki/index.php/2025_USAMO_Problems/Problem_6
Let \(m\) and \(n\) be positive integers with \(m\geq n\). There are \(m\) cupcakes of different flavors arranged around a circle and \(n\) people who like cupcakes. Each person assigns a nonnegative real number score to each cupcake, depending on how much they like the cupcake. Suppose that for each person \(P\), it is possible to partition the circle of \(m\) cupcakes into \(n\) groups of consecutive cupcakes so that the sum of \(P\)'s scores of the cupcakes in each group is at least \(1\). Prove that it is possible to distribute the \(m\) cupcakes to the \(n\) people so that each person \(P\) receives cupcakes of total score at least \(1\) with respect to \(P\).
[ "https://artofproblemsolving.com/wiki/index.php/File:2025_USAMO_PROBLEM_6.jpeg\n\nArbitrarily pick any one person — call her Pip — and her n arcs. The initial idea is to try to apply Hall's marriage lemma to match the n people with Pip's arcs (such that each such person is happy with their matched arc). To that end, construct the obvious bipartite graph G between the people and the arcs for Pip. We now consider the following algorithm, which takes several steps. If a perfect matching of G exists, we're done! We're probably not that lucky. Per Hall's condition, this means there is a bad set B1 of people, who are compatible with fewer than |B1| of the arcs. Then delete B1 and the neighbors of B1, then try to find a matching on the remaining graph. If a matching exists now, terminate the algorithm. Otherwise, that means there's another bad set B2 for the remaining graph. We again delete B2 and the fewer than B2 neighbors. Repeat until some perfect matching M is possible in the remaining graph, i.e. there are no more bad sets (and then terminate once that occurs). Since Pip is a universal vertex, it's impossible to delete Pip, so the algorithm does indeed terminate with nonempty M. We commit to assigning each of person in M their matched arc (in particular if there are no bad sets at all, the problem is already solved). Now we finish the problem by induction on n (for the remaining people) by simply deleting the arcs used up by M. To see why this deletion-induction works, consider any particular person Quinn not in M. By definition, Quinn is not happy with any of the arcs in M. So when an arc A of M is deleted, it had value less than 1 for Quinn so in particular it couldn't contain entirely any of Quinn's arcs. Hence at most one endpoint among Quinn's arcs was in the deleted arc A. When this happens, this causes two arcs of Quinn to merge, and the merged value is (≥ 1) + (≥ 1) - (≤ 1) ≥ 1 meaning the induction is OK.\n\nRemark. This deletion argument can be thought of in some special cases even before the realization of Hall, in the case where M has only one person (Pip). This amounts to saying that if one of Pip's arcs isn't liked by anybody, then that arc can be deleted and the induction carries through.\n\nRemark. Conversely, it should be reasonable to expect Hall's theorem to be helpful even before finding the deletion argument. While working on this problem, one of the first things I said was: “We should let Hall do the heavy lifting for us: find a way to make n groups that satisfy Hall's condition, rather than an assignment of n groups to n people.” As a general heuristic, for any type of “compatible matching” problem, Hall's condition is usually the go-to tool. (It is much easier to verify Hall's condition than actually find the matching yourself.) Actually in most competition problems, if one realizes one is in a Hall setting, one is usually close to finishing the problem. This is a relatively rare example in which one needs an additional idea to go alongside Hall's theorem." ]