diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzigdr" "b/data_all_eng_slimpj/shuffled/split2/finalzzigdr" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzigdr" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction} \\label{Ss.Intro}\n\nWe consider {\\it combinatorial games} of two players;\nthey take turns alternating and one who makes the last move wins.\nBoth players have perfect information and there are no moves of chance.\nA game is called {\\it impartial} if both players\nhave the same possible moves in each position and\n\\emph{acyclic} if it is impossible to revisit the same position.\nIn this paper we consider only impartial acyclic combinatorial games\nand call them simply games.\nA more detailed introduction to combinatorial games can be found in \\cite{BCG01-04, Con76}.\n\nIf there is a move from position $x$ to $y$, we write $x \\to y$.\nFor a set $S$ of nonnegative integers the {\\it minimum excluded value} of $S$\nis defined as the smallest nonnegative integer that is not in $S$ and\nis denoted by $\\mex (S)$. In particular, $\\mex(\\emptyset) = 0$.\nThe {\\it Sprague-Grundy} (SG) value of a position $x$\nin a game $G$ is defined recursively by $$\\G(x) = \\mex \\{g(y) \\mid x \\to y\\}.$$\nA position of SG value $t$ is called a {\\it $t$-position};\n$0$-positions are also known as $\\P$-positions.\nA player who moves into a $\\P$-position can win the game.\nThe SG function is instrumental in the theory of disjunctive sums of games;\nsee \\cite{BCG01-04, Con76, Gru39, Spr35, Spr37}.\n\n\nA classical example is {\\sc Nim} studied by Bouton \\cite{Bou901}.\nA position in {\\sc Nim} consists of $n$ piles of tokens.\nTwo players alternately choose one of the piles and remove\nan arbitrary (positive) number of tokens from that pile.\nBouton characterized the $\\P$-positions and in fact described the SG function of {\\sc Nim}.\n\nMoore \\cite{Moo910} introduced a generalization in which\na player can remove tokens from at least one and at most $k$ of the piles, for some fixed $k M(x) = 2$ for $k = 2, n = 3, x = (2,3,3)$.\n\nFor the case of $n = k+1$ Jenkyns and Mayberry \\cite{JM80}\nprovided a formula for the SG function of {\\sc Nim}$^\\leq_{k+1, k}$.\nAn alternative proof for a slightly more general game\nwas given recently in \\cite{BGHM15}.\n\nIn this paper we introduce another generalization of {\\sc Nim}.\nGiven positive integers $n$ and $k$ such that $1 \\leq k \\leq n$,\nwe define {\\sc Exact $k$-Nim}, denoted by {\\sc Nim$^=_{n,k}$}, as follows.\nGiven $n$ piles of tokens, by one move a player chooses exactly $k$ piles and\nremoves arbitrary positive number of tokens from each of them.\nThe game terminates when there are less than $k$ nonempty piles.\n{\\sc Nim}$^=_{n, k}$ turns into the standard {\\sc Nim} when $k=1$ and\nit is the trivial one-pile {\\sc Nim} when $k=n$.\n\n\\subsection*{Main results}\n\nGiven a position $x\\in \\Z_\\geq^n$ of {\\sc Nim$^=_{n,k}$},\nwe denote by $T_{n,k}(x)$ the maximum number of consecutive moves\none can make starting with $x$.\nWe call $T_{n,k}$ the \\emph{Tetris} function of the game.\n\nThe following two theorems characterize the SG function of {\\sc Nim$^=_{n,k}$} for $2k\\geq n$.\n\\begin{theorem} \\label{thm.n<2k}\nIf $n < 2k$ then the SG function of {\\sc Nim$^=_{n,k}$} is\nequal to its Tetris function, $\\G(x) = T_{n,k}(x)$.\n\\end{theorem}\n\n\\begin{theorem} \\label{thm.n=2k}\nLet $k \\geq 2$, $n = 2k$, and\nlet $x = (x_1, \\ldots, x_n)$ be a position of {\\sc Nim}$^=_{2k, k}$.\nSet\n\\begin{align}\nu(x) &= T_{2k,k}(x), \\label{eq:u}\\\\\nm(x) &= \\min_{1\\leq i\\leq 2k} x_i,\\label{eq:m}\\\\\ny(x) &= T_{2k,k}\\big(x_1 - m(x), \\ldots, x_{2k} - m(x)\\big),\\label{eq:y}\\\\\nz(x) &= 1+\\binom{y(x)+1}{2}, \\text{ and}\\label{eq:z}\\\\\nv(x) &= \\big(z(x)-1\\big) +\\big[\\big(m(x)-z(x)\\big)\\mod \\big(y(x)+1\\big)\\big].\\label{eq:v}\n\\end{align}\nThen the SG function of {\\sc Nim}$^=_{2k,k}$ is given by formula\n\\begin{equation}\\label{eq-binomial;n=2k-Formula}\n\\G(x) ~=~ \\begin{cases}\nu(x), & \\text{if~~} m(x) < z(x);\\\\\nv(x), & \\text{if~~} m(x) \\geq z(x).\n\\end{cases}\n\\end{equation}\n\\end{theorem}\n\nNote that this formula fails for $k = 1$.\nIn this case {\\sc Nim}$^=_{2, 1}$ is the standard $2$-pile {\\sc Nim} and\nits SG function is the modulo $2$ sum of the two coordinates of a position,\nas described by Bouton.\nThis function is different from the one described by the above formula.\n\nWe also would like to remark that the above formula is surprisingly\nsimilar to the one given by Jenkyns and Mayberry\nin \\cite{JM80} for {\\sc Nim}$^\\leq_{k+1, k}$.\n\n\\bigskip\n\nThe above result implies a simple characterization of\nthe $0$- and $1$-positions of {\\sc Nim}$^=_{2k, k}$.\nA position $x = (x_1, \\ldots, x_n)$ is said to be\n\\emph{nondecreasing} if $x_1 \\leq x_2 \\leq \\cdots \\leq x_n$.\n\n\n\\begin{corollary} \\label{cor-0,1-exact:n=2k}\nGiven a nondecreasing position $x = (x_1, \\dots, x_{2k})$ of the game {\\sc Nim}$^=_{2k, k}$,\n\\begin{enumerate} \\itemsep0em\n\\item [\\rm{(i)}] $x$ is a $\\P$-position if and only if\nmore than half of its smallest coordinates are equal, that is, $x_1 = \\cdots = x_k = x_{k+1}$.\n\\item [\\rm{(ii)}] $x$ is a $1$-position if and only if $x_1 = \\cdots = x_{k - \\ell} = 2c$ and\n$x_{k-\\ell+1} = \\cdots = x_{k+ \\ell + 1} = 2c+1$\nfor some integer $c \\in \\ZZ_\\geq$ and $\\ell \\in \\{0,1, \\ldots, k-1\\}$.\n\\end{enumerate}\n\\end{corollary}\n\nBoth statements \\rm{(i)} and \\rm{(ii)} follow from Theorem \\ref{thm.n=2k},\nbut can also be derived much simpler, directly from the definitions.\n\n\\medskip\n\nThe case of $2k < n$ looks much more difficult and it is still open.\nMoreover, we have not even been able to characterize\nthe $\\P$-positions of {\\sc Exact $k$-Nim} for $1 < k < n\/2$,\ne.g., for {\\sc Nim}$^=_{5,2}$.\n\n\\bigskip\n\nThe rest of the paper is organized as follows.\nIn Section \\ref{Ss.n<2k} we characterize the SG function of {\\sc Nim$^=_{n,k}$} for the case $2k>n$.\nIn Section \\ref{Ss.n=2k} we characterize the SG function of {\\sc Nim$^=_{n,k}$} for the case $2k=n$.\nIn Section \\ref{Ss.Moore01} we provide an alternative proof for the above stated result of Jenkyns and Mayberry \\cite{JM80}.\nFinally in Section \\ref{Ss.Tetris} we show that for a given position we can compute efficiently the corresponding SG value.\n\n\n\n\n\\section{SG function in the case of $n < 2k$} \\label{Ss.n<2k}\n\n\nFor our proof we need the following basic properties of the Tetris function.\nGiven positions $x,x'\\in\\Z^n_\\geq$ we write $x\\leq x'$ if $x_i\\leq x'_i$ holds for $i=1,...,n$.\n\n\\begin{lemma}\\label{tbasic}\nConsider two positions $x, x'\\in\\ZZ^n_\\geq$.\n\\begin{itemize}\n\\item [\\rm{(i)}] If $x'\\leq x$ then $T_{n,k}(x')\\leq T_{n,k}(x)$.\n\\item [\\rm{(ii)}] If in addition we have $\\sum_{i=1}^n(x_i-x_i')=1$, then $T_{n,k}(x)-1\\leq T_{n,k}(x')\\leq T_{n,k}(x)$.\n\\end{itemize}\n\\end{lemma}\n\\proof\nIt is immediate by the definition.\n\\qed\n\nA move in {\\sc Nim}$^=_{n, k}$ is called {\\it slow} if exactly one token is taken from each of the $k$ chosen piles.\n\n\n\n\\begin{lemma}\\label{Tetrisnotdecrease}\nConsider a position $x=(x_1, \\ldots, x_n) \\in \\Z^n_\\geq$ with some indices $i, j$ such that\n$x_i < x_j$. Let $x'=(x'_1, \\ldots, x'_n)$ be defined by\n\\begin{equation}\\label{eq-l2}\nx_l'=\n\\begin{cases}\nx_i+1, & \\text{ if } l=i, \\\\\nx_j-1, & \\text{ if } l=j, \\\\\nx_l, & \\text{ otherwise. }\\\\\n\\end{cases}\n\\end{equation}\nThen we have $T_{n,k}(x) \\leq T_{n,k}(x')$.\n\\end{lemma}\nIn other words, the Tetris function is nondecreasing when\nwe move a token from a larger pile to a smaller one.\n\n\n\\proof\nConsider any sequence of slow moves from $x \\to \\cdots \\to x''$. If $x''_j>0$ then the same sequence of slow moves can be made from $x'$ since $x'_l \\geq x_l$ for $l \\neq j$.\n\nIf $x''_j=0$ then since $x_j > x_i$, this sequence contains a slow move reducing $x_j$ but not $x_i$.\nLet us modify this move reducing $x_i$ rather than $x_j$ and keeping all other moves of the sequence unchanged.\nThe obtained sequence has the same length and consists of slow moves from $x'$. \\qed\n\nNotice that we can generalize Lemma \\ref{Tetrisnotdecrease}\nreplacing $\\pm 1$ in \\raf{eq-l2} by $\\pm \\Delta$ for any integral $\\Delta \\in [0,x_j-x_i]$.\n\n\n\\begin{lemma}\\label{bestslowmove}\nThe slow move that reduces the $k$ largest piles of $x$ reduces the Tetris value $T_{n,k}(x)$ by exactly one.\n\\end{lemma}\n\n\\proof\nLet $x'$ be the position obtained from $x$ by reducing the $k$ largest piles of $x$ by exactly one each. Let $x''$ be another position obtained by some slow move.\nBy applying \\raf{eq-l2} repeatedly, we can obtain $x'$ from $x''$ with $T_{n,k}(x'') \\leq T_{n,k}(x')$ by Lemma \\ref{Tetrisnotdecrease}.\nThis implies that $x'$ has the highest Tetris value among all positions each reachable from x by a slow move. By Lemma \\ref{tbasic},\neach slow move reduces the Tetris value by at least one and there exists a slow move reducing it by exactly one.\nHence, $T_{n,k}(x')=T_{n,k}(x)-1$. \\qed\n\n\\bigskip\n\\noindent\\textbf{Proof of Theorem }\\ref{thm.n<2k}:\nThe Tetris value is the largest number of moves one can take from a position $x$,\nimplying that $\\G(x)$ is at most $T_{n,k}(x)$.\nTherefore, it is enough to show that for all integral $g$ such that\n$0 \\leq g < T_{n,k}(x)$ there exists a move $x\\to x'$ such that $T_{n,k}(x')= g$.\n\nConsider the move $x \\to x'$ that reduces the largest $k$ piles to $0$.\nFor the resulting position we have $T_{n,k}(x')=0$ because $2k>n$.\nLet us also consider the move $x \\to x''$ that reduces the $k$ largest piles each by only $1$.\nThen we have $T_{n,k}(x'')=T_{n,k}(x)-1$ according to Lemma \\ref{bestslowmove}.\nAny position between $x'$ and $x''$ is reachable form $x$. Thus, by Lemma \\ref{tbasic} the claim follows. \\qed\n\n\\section{SG function in case of $n=2k$} \\label{Ss.n=2k}\n\nLet us consider the game {\\sc Nim}$^=_{2k, k}$, where $k\\geq 2$, and let $x$ be a position of this game.\nRecall that to $x$ we associated several parameters in \\eqref{eq:u}--\\eqref{eq:v}.\nBased on these parameters, we classify the positions into the following two types:\n\\begin{itemize} \\itemsep0em\n\\item [\\rm{(i)}] \\emph{type I}, if $m(x) m(x)$.\nObviously, any move $x \\rightarrow x'$ reduces the Tetris value by at least 1, implying $u(x) > u(x')$.\nUsing this and the definitions, we get $g(x)=u(x) > u(x')$. If $g(x') = u(x')$ then by the above inequality, we get $g(x) \\neq g(x')$.\nOn the other hand, if $g(x') = v(x')$ then we must have $m(x') \\geq z(x')$ and thus $v(x')=z(x')-1 + \\left((m(x')-z(x')) \\mod (y(x')+1)\\right)\\leq z(x')-1 + (m(x')-z(x'))=m(x')-1$.\nThus, we get $g(x) = u(x) > u(x') > m(x')-1 \\geq v(x') = g(x')$.\n\nIt remains to consider the case $z(x) \\leq m(x)$, in which case $v(x)\\leq m(x)-1$ follows by the definitions. \n\t\\begin{enumerate} \\itemsep0em\n\t\\item Suppose $z(x') > m(x')$.\n We can estimate $u(x') \\geq m(x)+m(x')$ since $2k = n$.\n Furthermore, $g(x)=v(x) \\leq m(x)-1$ and thus $g(x')=u(x') \\geq m(x)+m(x') > m(x)-1 \\geq g(x)$.\n \\item Suppose $z(x') \\leq m(x')$. Then $g(x')=v(x')$. Note that $z(x) \\leq m(x)$ and thus $g(x)=v(x)$.\n Also note that $m(x') \\leq m(x)$. We examine the last inequality.\n \t\\begin{enumerate} \\itemsep0em\n\t \\item Suppose $m(x')=m(x)$.\n \n Then $x-m(x)\\to x'-m(x')$ is a move, and hence decreases the Tetris value implying $y(x)>y(x')$.\n\t By Lemma \\ref{interval}, we have\n $z(x')+y(x')-1 < z(x)-1$\n and so\n $z(x) > z(x') + y(x')$\n implying\n $v(x) \\geq z(x)-1> z(x')-1+y(x') \\geq v(x')$,\n since\n $y(x') \\geq \\bigg(\\big(m(x') - z(x)\\big) \\mod\\big(y(x) + 1\\big)\\bigg)$,\n regardless of the value of $m(x') - z(x)$.\n\t \\item Suppose $m(x') < m(x)$. We compare $y(x)$ with $y(x')$.\n \\begin{enumerate} \\itemsep0em\n\t \\item If $y(x)= y(x')$ then $z(x)=z(x')$. By the definition of a legal move, $x'$ has at least $k$ piles not smaller than $m(x)$. Therefore, $1 \\leq\n m(x)-m(x') \\leq y(x')=y(x)$, which implies that $m(x) \\mod (y(x)+1) \\neq m(x') \\mod (y(x)+1)$ and thus $v(x) \\neq v(x')$.\n\t \\item If $y(x) \\neq y(x')$, since the $y$ values are different, by Lemma \\ref{interval} we have that $v(x)$ and $v(x')$ are in different intervals,\n therefore $v(x)\\neq v(x')$.\n \\end{enumerate}\n\t \\end{enumerate}\n\t\\end{enumerate}\n\n\\subsection{Proof of (II)}\n\nWe prove property (II) by considering type I and type II positions, separately.\n\n\\subsubsection{Type I positions: $m(x)\\nu(m-1)\\geq 0.\n\\end{equation}\nwith $\\nu$ being defined as in Corollary \\ref{delta-epsilon}.\n\nLet us next define a set $Q=Q(x)$ of pairs of integers by setting\n\\[\nQ^1(x)=\\left\\{(\\mu,\\ge)\\left| \\begin{array}{c}m-\\ge~\\leq~ \\mu~\\leq~ m\\\\\n\\ge ~\\leq \\nu(m-1)-1\n\\end{array}\\right.\\right\\}\n\\]\n\\[\nQ^2(x)=\\left\\{(\\mu,\\ge) \\left| \\begin{array}{c}m-\\eps(m-1)~\\leq~ \\mu~\\leq~ m\\\\\n\\ge ~= \\nu(m-1)\n\\end{array}\\right. \\right\\} ,\n\\]\nand defining $Q=Q^1(x)\\cup Q^2(x)$.\n\nWe show that if for a position $x^*$ we have $(m(x^*),y(x^*))\\in Q$, then $x^*$ is of type II.\nTo see this consider a pair $(\\mu,\\ge)\\in Q^1(x)$.\nThen we have by the definition of $Q^1(x)$ that $\\mu\\geq m-\\ge$ and that $\\nu(m-1)-1\\geq \\ge$ from which $z(\\nu(m-1)-1)\\geq z(\\ge)$ follows by the definition of $z$ in \\eqref{eq:z}.\nWe also have the inequality $m-\\nu(m-1)+1\\geq z(\\nu(m-1)-1)$ since $m\\geq \\binom{\\nu(m-1)+1}{2}$ by the definition of $\\nu$ in Corollary \\ref{delta-epsilon}.\nPutting these together, we obtain $\\mu\\geq z(\\ge)$ as stated.\nFor $(\\mu,\\ge)\\in Q^2(x)$ we have $\\mu\\geq m-\\epsilon(m-1)$ and $\\ge=\\nu(m-1)$ by the definition of $Q^2(x)$.\nSince $m-1=\\binom{\\nu(m-1)+1}{2}+\\epsilon(m-1)$ by Corollary \\ref{delta-epsilon},\nthe inequality $\\mu\\geq m-\\epsilon(m-1)=z(\\ge)$ follows again.\n\nWe show next that for all pairs $(\\mu,\\ge)\\in Q$ there exists a move $x\\to x^*$ such that $\\mu=m(x^*)$ and $\\ge=y(x^*)$ (and $x^*$ is of type II, as we argued in the previous paragraph.)\nFor this let us consider first $\\mu\\leq m-1$ and note that if $(\\mu,\\ge)\\in Q$ then $m-\\mu\\leq \\ge\\leq \\nu(m-1)$.\nOur plan is to use Lemma \\ref{continuity} and to cover this range of $\\ge$ values by two constructions.\n\nLet us define a pair of positions $x'\\geq x''$ by\n\\begin{align*}\nx'_i &=\n\\begin{cases}\nx_i, &\\text{ for } i=1 \\text{ and } i\\geq k+2,\\\\\n\\mu, &\\text{ for } i=2, \\\\\nx_i-1, &\\text{ for } i=3,...,k+1\n\\end{cases} \\\\\n\\intertext{and}\nx''_i &=\n\\begin{cases}\nx_i, &\\text{ for } i=1 \\text{ and } i\\geq k+2,\\\\\n\\mu, &\\text{ for } i=2,...,k+1.\n\\end{cases}\n\\end{align*}\nNote that since $\\mu\\delta\\geq x_1$ which implies $x'_n=\\delta-x_10$ for $i=k+1,...,n-1$,\nand therefore we indeed decrease exactly $k$ piles of $x$ to obtain $x'$. Thus $x'$ is reachable from $x$.\n\nFor $x'$ we have\n\\begin{align*}\n\\sum_{i=1}^{2k}\\min(x_i',\\delta) &\\geq x_1' + (k-1)\\delta + (\\delta-x_1') =k\\delta \\\\\n\\intertext{and}\n\\sum_{i=1}^{2k}\\min(x_i',\\delta+1) &\\leq x_1' + (k-1)(\\delta+1) + (\\delta - x_1)\\\\\n &= \\delta + (k-1)(\\delta +1) < k(\\delta +1).\n\\end{align*}\nTherefore, by Lemma \\ref{getmeTetris}, we have $T_{2k,k}(x')=\\delta$.\n\nSince $k\\geq 2$, $x'_{k+1}=0$ and therefore $m(x')=0$. Thus, $m(x')=0<1\\leq z(x')$ implying\nthat $x'$ is of type I, from which $g(x')=u(x')=\\delta$ follows by the above.\n\\begin{figure}[ht]\n\\centering\n\\includegraphics[height=7cm]{Doc2}\n\\caption{$x'$ is obtained by removing the gray area.} \\label{F1}\n\\end{figure}\n\n\\item $x_2 \\leq \\delta < u(x)-m(x)$ (Figure \\ref{F2}). Let $A=\\sum_{i=2}^{k+1} \\max(0,\\delta-x_i)$.\nFor $i=k+2, \\ldots ,n$ choose $a_i$ such that $0 \\leq a_i \\leq \\min(x_i-1,\\delta)$ and $\\sum_{i=k+2}^n a_i=A$.\nFirst let us prove that this is possible, or equivalently that $A \\leq \\sum_{i=k+2}^n \\min(\\delta,x_i-1)$.\nTo see this let us consider two cases.\nIf $x_{k+1}>\\delta$ then $\\sum_{i=k+2}^n \\min(\\delta,x_i-1)\\geq (k-1)\\delta \\geq A$.\nIf $x_{k+1}\\leq \\delta$ then let us define\n\\begin{equation} \\label{B.def}\nB=\\sum_{i=2}^{k+1}x_i = k\\delta - A,\n\\end{equation}\nand observe that for any integer $t\\geq\\delta >x_{k+1}$ we have\n\\[\n\\sum_{i=1}^n\\min(t,x_i) = m + B +\\sum_{i=k+2}^n\\min(t,x_i).\n\\]\nConsequently, since we have $T_{2k,k}(x)=u(x)$, by Lemma \\ref{getmeTetris} we can write for $t=u(x)$ that\n\\[\nkt \\leq \\sum_{i=1}^n\\min(t,x_i) = m(x)+B+\\sum_{i=k+2}^n\\min(t,x_i).\n\\]\nNote that if we decrease $t=u(x)$ by $1$,\nthen the left hand side decreases by $k$ while the right hand side decreases by at most $k-1$,\nhence the inequality remains valid. Let us repeat this $m(x)$ times, obtaining the inequality\n\\[\nk(u(x)-m(x))+km(x) \\leq m(x)+B+\\sum_{i=k+2}^n\\min(u(x)-m(x),x_i)+(k-1)m(x)\n\\]\nfrom which\n\\[\nk(u(x)-m(x))\\leq B+\\sum_{i=k+2}^n\\min(u(x)-m(x),x_i)\n\\]\nfollows. Let us now decrease $t=u(x)-m(x)$ further by $1$, as well as replace $x_i$ by $x_i-1$.\nThen the left hand side decreases by exactly $k$, while the right hand side decreases by at most $k$, yielding the valid inequality\n\\[\nk(u(x)-m(x)-1)\\leq B+\\sum_{i=k+2}^n\\min(u(x)-m(x)-1,x_i-1).\n\\]\nFinally, we can decrease $t=u(x)-m(x)-1$ further on both sides to $t=\\delta$ and similarly to the above argument obtain\n\\[\nk\\delta\\leq B+\\sum_{i=k+2}^n\\min(\\delta,x_i-1).\n\\]\nBy \\eqref{B.def} we obtain the claimed inequality, and hence the proof for the existence of the $a_i$ values for $k=k+2,...,n$\nthat satisfy the desired inequalities.\n\nLet us now consider the position $x'$ defined by\n\\[\nx'_i =\n\\begin{cases}\n0, & \\textrm{ for } i=1; \\\\\nx_i, &\\textrm{ for }i=2, \\ldots, k+1; \\\\\na_i, &\\textrm{ for }i=k+2, \\ldots ,n.\n\\end{cases}\n\\]\nBy the above arguments $x\\to x'$ is a move in the game.\nThe equality $g(x')=u(x')=\\delta$ now follows by the above analysis and Lemma \\ref{getmeTetris}, completing our proof in this case.\n\n\\begin{figure}[ht]\n\\centering\n\\includegraphics[height=9cm]{Doc3}\n\\caption{$x'$ is obtained by removing the gray area.} \\label{F2}\n\\end{figure}\n\n\n\\item $u(x)-m(x) \\leq \\delta < u(x)$ (Figure \\ref{F3}).\nLet us define position $x'$ as follow\n$$\nx_i' =\n\\begin{cases}\nx_i-u(x)+\\delta, &\\textrm{ for } i \\in I_1=\\{i \\mid i\\leq k, x_{i+k} \\leq \\delta\\}; \\\\\nx_i-u(x)+\\delta, &\\textrm{ for } i \\in I_2=\\{i \\mid i > k, \\delta < x_i \\leq u(x)\\}; \\\\\n\\delta , &\\textrm{ for }i \\in I_3=\\{i \\mid i > k, x_i > u(x) \\}; \\\\\nx_i, &\\textrm{ otherwise.}\n\\end{cases}\n$$\nNote that $x_i \\geq m(x)\\geq u(x)-\\delta$, therefore $x_i'$ are all nonnegative.\nIt is easy to see that $I_1+k,I_2,I_3$ form a partition of $\\{ k+1, \\ldots , n \\}$.\nWe have reduced $x_i$ for all $i \\in I_1 \\cup I_2 \\cup I_3$, therefore $x\\to x'$ is a move.\n\nNext we note that the above construction implies that\n\\[\n\\sum_{i=1}^n \\min(x_i',\\delta) \\geq \\sum_{i=1}^n\\min(x_i,u(x)) - k (u(x)-\\delta)\\geq k \\delta\n\\]\nwhere the last inequality follows by the fact that $u(x)=T_{2k,k}(x)$, and hence $\\sum_{i=1}^n\\min(x_i,u(x))\\geq ku(x)$ by Lemma \\ref{getmeTetris}.\nSimilarly, we get\n\\[\n\\sum_{i=1}^n \\min(x_i',\\delta+1) \\leq \\sum_{i=1}^n\\min(x_i,u(x)+1) - k (u(x)-\\delta)< k (\\delta+1)\n\\]\nsince $\\sum_{i=1}^n\\min(x_i,u(x)+1)< k(u(x)+1)$.\nTherefore $T_{n,k}(x')=u(x')= \\delta$ follows by Lemma \\ref{getmeTetris}.\n\nNote that $x_{k+1}\\leq \\delta$, since otherwise $u(x)\\geq \\delta+m(x)$ would follow, contradicting our choice of $\\delta$.\nIt follows that $x_1 \\in I_1$ and thus $m(x')=m(x)-u(x)+\\deltam(x)$.\nWe consider two positions $\\bar{x}$ and $\\hat{x}$ reachable from $x$ defined as\n$$\\bar{x}_i =\n\\begin{cases}\nx_i, & \\textrm{ if } i=1, \\ldots ,k;\\\\\nm(x), & \\textrm{ if } i\\geq k+1; \\\\\n\\end{cases}\n~~~~\\hat{x}_i =\n\\begin{cases}\nx_i, & \\textrm{ if } i=1, \\ldots ,k; \\\\\nx_i-1, & \\textrm{ if } i\\geq k+1. \\\\\n\\end{cases}\n$$\nSimilarly to the previous cases, we have $y(\\bar{x})=0$ and\n$y(\\hat{x}) = y(x)-1$, and thus by Lemma \\ref{continuity} it follows that\nfor all $y \\in[0,y(x)-1]$ there exists an $x' \\in [\\bar{x},\\hat{x}]$ such that $y=y(x')$.\n\nIf we put all three cases together we cover all values $(m,y) \\in D$.\n\\qed\n\n\\medskip\n\nLet us set\n\\begin{equation} \\label{v(m,y)}\nv(m,y)= {y+1 \\choose 2}+ \\Big[ \\Big(m-1-{y+1 \\choose 2}\\Big)\\mod (y+1)\\Big].\n\\end{equation}\nNote that if $m=m(x)$ and $y=y(x)$ then $v(m,y)=v(x)$.\nFurthermore, we have\n$$V(y):=\\{v(m,y) | m \\in \\ZZ_\\geq\\}=\\left[ {y+1 \\choose 2},{y+2 \\choose 2} \\right).$$\nTherefore the sets $V(y)$, $y \\in \\ZZ_\\geq$ form a partition of $\\ZZ_\\geq$, as shown in Lemma \\ref{interval}.\n\n\\begin{lemma}\\label{Dstar}\nIf $m(x) \\geq z(x)$, then every $(m,y) \\in D(x)$ satisfies the following relations\n$$m\\geq {y+1 \\choose 2}+1~~\\textrm{ and }~~\\{v(m,y) \\mid (m,y) \\in D(x) \\} = [0,v(x)).$$\n\\end{lemma}\n\n\\proof\nLet us first consider $(m,y) \\in D_1(x)$.\nBy the definition of $D_1$ and the assumption of $m(x) \\geq z(x)={y(x)+1 \\choose 2}+1$ we get\n$$m\\geq m(x)-y \\geq {y(x)+1 \\choose 2}+1-y \\geq {y+2 \\choose 2}+1-y \\geq {y+1 \\choose 2}+1.$$\nBy the definition of $D_1(x)$, for all $y \\in [0,y(x))$ we have $(m,y) \\in D_1(x)$ for all $m \\in [m(x)-y,m(x)]$. Hence, by (\\ref{v(m,y)}) we have $\\{v(m,y) \\mid m \\in [m(x)-y,m(x)]\\}=V(y)$.\nThus, by Lemma~\\ref{interval} we get $$\\{v(m,y) \\mid (m,y) \\in D_1(x) \\}=\\bigcup_{y \\in [0,y(x))}V(y)=[0,z(x)-1).$$\n\n\\smallskip\n\nLet us next consider $(m,y) \\in D_2(x)$.\nBy the definition of $ v(x)$ we can write $v(x)=z(x)-1+r $, where $r=m(x)-z(x)-\\lambda(y(x)+1)$ for some $\\lambda \\in \\ZZ_\\geq$. This implies $m(x)-r\\geq z(x)$. By the definition of $D_2(x)$ we have $m\\geq m(x)-r$.\nThese two inequalities imply $m\\geq {y+1 \\choose 2} +1$.\nSince $m \\in [m(x)-r, m(x))$ takes $r$ consecutive values, we have\n$\\{v(m,y) \\mid (m,y) \\in D_2(x) \\}=[z(x)-1,v(x))$.\n\\qed\n\nThe above lemma implies that any position $x'$ with $m(x')=m$, $y(x')=y$ for some $(m,y)\\in D$\nis a type II position. Hence $g(x')=v(x')=v(m,y)$.\nThus the second claim in the above lemma together with Lemma \\ref{existsmoveinD} implies that for any $0\\leq \\delta x_{ij}$, then we must have a $j'>j$ such that $x'_{ij'} 0$, there exists a move $x \\to x'$ such that $M(x') = 0$.\n\\end{itemize}\n\nTo show \\rm{(i0)}, let us consider a move $x \\to x'$\nfrom a position $x$ with $M(x)=0$.\n\nLet $j$ be the highest binary bit such that\n$x_{ij}$ and $x'_{ij}$ differ for some $i$. Such a $j$ must exist since in a move we must change at least one components.\nBy Lemma \\ref{moore1} we have $x'_{ij}\\leq x_{ij}$ for all $i=1,...,n$,\nimplying $1\\leq \\sum_{i=1}^n(x_{ij}-x'_{ij}) \\leq k$ because in a move we can change at most $k$ components. Therefore,\n$(\\sum_{i}x'_{ij} \\mod (k+1)) \\not= 0$ and, thus, $M(x') > 0$.\n\n\\medskip\n\nTo show \\rm{(a0)}, let us consider a position $x$ with $M(x)>0$. We will construct a move $x \\to x'$ such that $M(x') = 0$.\n\n\\begin{notation} \\label{T.N}\nLet $t_1, \\dots , t_p$ denote the bits $j$ such that $y_j\\not=0$, assuming $t_1 > \\dots > t_p$.\nSet $N = \\{1,2,\\ldots,n\\}$.\n\\end{notation}\n\n\nThe following algorithm defines index sets $\\emptyset=I_0\\subseteq I_1\\subseteq \\cdots I_p \\subseteq N$\nsuch that we can compute a move $x \\to x'$ with $M(x')=0$, by decreasing components $i\\in I_p$. Define $O_j=\\{i\\in I_j\\mid x_{it_{j+1}}=1\\}$, and set $\\ga_j = |I_j|$ and $\\gb_j=|O_j|$.\n\n\n\n\\medskip\n\\noindent\n{\\bf Step 0}. Initialize $I_0=\\emptyset$\nand hence, $\\ga_0 = \\gb_0 = 0$, and set $x'_i:=x_i$ for all $i \\in N$.\n\n\\smallskip\n\\noindent\n{\\bf Step 1}. For $j=1, \\dots , p$, construct $I_j$ and update $x'$ as follows.\n\n\\smallskip\n\n{\\bf Case 1.} If $y_{t_j} \\leq \\beta_{j-1}$, then let $I_j:=I_{j-1}$,\nchoose $y_{t_j}$ many indices $i\\in O_{j-1}$,\nand update $x'_{it_j}:=0$.\n\n\\smallskip\n\n{\\bf Case 2.} If $y_{t_j} > \\beta_{j-1}$ and $(k+1) -y_{t_j} \\leq \\ga_{j-1}-\\gb_{j-1}$, then\nlet $I_j:=I_{j-1}$, choose $(k+1)-y_{t_j}$ many indices $i\\in I_{j-1}\\setminus O_{j-1}$, and update $x'_{i{t_j}}:=1$.\n\n\\smallskip\n\n{\\bf Case 3.} If $y_{t_j} > \\beta_{j-1}$ and $y_{t_j}-\\beta_{j-1} \\leq k-\\ga_{j-1}$, then\nlet $I_j$ be the index set obtained from $I_{j-1}$ by adding\n$y_{t_j} - \\beta_{j-1}$ many indices $i$ from $N \\setminus I_{j-1}$ such that\n$x_{i{t_j}}=1$. Update $x'_{i{t_j}}:=0$ for $i\\in (I_j\\setminus I_{j-1})\\cup O_{j-1}$.\n\n\n\\medskip\n\\noindent\nWe first note that the three cases above are exclusive and\ncover all possible $y_{t_j}$, $\\ga_{j-1}$ and $\\gb_{j-1}$ values.\nMoreover, it is easily seen that\na position $x'$ after the execution of the algorithm satisfies $M(x')=0$, and\n$x'_i=x_i$ holds for $i \\not\\in I_p$.\n\nNote that we increase the set $I_j$ only in Case 3, in which case we have\n\\[\n|I_j|=\\alpha_j=(y_{t_j}-\\beta_{j-1})+\\alpha_{j-1} \\leq k,\n\\]\nimplying $|I_p|\\leq k$.\n\nNote also that in Case 3 we must have at least $y_{t_j} - \\beta_{j-1}$ many indices $i\\in N\\setminus I_{j-1}$ with $x_{it_j}=1$ by the definition of $y_{t_j}$ in \\eqref{e0001}.\n\nIt remains to show that $x' < x$.\nAssume that $i$ is an index such that $x'_i \\not= x_i$.\nThen some $j$ satisfies $i \\not\\in I_{j-1}$ and $i \\in I_j$.\nThis implies that $x'_i$ was first updated during the $j$th iteration of Step 1.\nNamely, the $t_j$th bit of $x'_i$ is modified from $1$ to $0$.\nSince $x'_{it}=x_{it}$ holds for all $t$ with $t > t_j$,\nwe have $x'_i < x_i$, which completes the proof.\n\\qed\n\n\\subsection{Proof of Theorem \\ref{t-Moore-m=0,1} for $m=1$}\n\\label{ss-p-t1-1}\n\nNow, let us prove that\n$\\cG(x)=1$ if and only if $M(x)=1$.\n\n\\smallskip\nThe proof in the previous subsection implies that for a position $x$ with $M(x)=1$ there exists a move $x\\to x'$ such that $M(x')=0$.\nBy the properties of the SG function, it remains to show that\n\\begin{itemize}\n\\item [\\rm{(i1)}] for any position $x$ with $M(x) = 1$, there exists\nno move $x \\to x'$ such that $M(x') = 1$;\n\\item [\\rm{(a1)}] for any position $x$ with $M(x) > 1$, there\nexists a move $x \\to x'$ such that $M(x')=1$;\n\\end{itemize}\n\nWe prove \\rm{(i1)} similarly to (i0).\nLet us assume that $M(x)=1$ holds for a position $x$\nand consider a move $x \\to x'$.\nLet $j$ be the highest binary bit such that\n$x_{ij}$ and $x'_{ij}$ differ for some $i$.\nThen by Lemma \\ref{moore1} we have $x'_{ij} \\leq x_{ij}$ for all $i$, and\n$1 \\leq \\sum_{i}(x_{ij}-x'_{ij}) \\leq k$.\nHence, $\\sum_{i}x'_{ij} \\not= \\sum_{i}x_{ij} (\\mod (k+1))$ and, thus, $M(x') \\not= 1$.\n\n\nTo show \\rm{(a1)}, let us consider a position $x$ with $M(x) > 1$.\nSimilarly to \\rm{(a0)}, we will algorithmically construct a move $x \\to x'$ such that $M(x')=1$.\n\nLet again $t_1, \\dots , t_{p-1}$ denote the bits $j>0$ such that $y_j>0$,\nwhere we assume that $t_1 > \\dots > t_{p-1}$, and add $t_p=0$.\n\nThe algorithm remains the same, as for \\rm{(a0)},\nexcept for $j=p$, when $t_p=0$. We detail below the computation of $I_p$ from $I_{p-1}$:\n\n\n\\medskip\n\\noindent\n\n{\\bf Case 1.} If $y_0> 1$ and $y_0-1 \\leq \\gb_{p-1}$,\nthen let $I_p:=I_{p-1}$,\nchoose $y_0-1$ many indices ${i}$ from $I_p$ such that\n$x'_{i0}=1$, and update $x'_{i0}:=0$ for such indices.\n\n\\smallskip\n\n{\\bf Case 2.} If $y_0> 1$ and $(k+2)-y_0 \\leq \\ga_{p-1}-\\beta_{p-1}$,\nlet $I_p:=I_{p-1}$, choose $(k+2)-y_{0}$ many indices ${i}$ from $I_p$ such\nthat $x'_{i0}=0$ and update $x'_{i0}:= 1$ for such indices.\n\n\\smallskip\n\n{\\bf Case 3.} If $y_0> 1$ and $0< (y_0-1) -\\gb_{p-1} \\leq k-\\ga_{p-1}$, then\nlet $I_p$ be an index set obtained from $I_{p-1}$ by adding $(y_{0}-1)-\\beta_{p-1}$ many\nindices $i$ from $N \\setminus I_{p-1}$ such that $x_{i{0}}=1$, and\nupdate $x'_{i0}:=0$ for all $i \\in I_p$ with $x'_{i0}=1$.\n\n\\smallskip\n\n{\\bf Case 4.} If $y_0=0$ and $\\ga_{p-1} > \\gb_{p-1}$,\nthen let $I_p:=I_{p-1}$,\nchoose an index ${i}$ from $I_p$ such that $x'_{i0}=0$, and update $x'_{i0}:=1$.\n\n\\smallskip\n\n{\\bf Case 5.} If $y_0=0$ and $\\ga_{p-1} = \\gb_{p-1}=k$,\nthen let $I_p:=I_{p-1}$, and update $x'_{i0}:=0$ for all $i \\in I_p$.\n\n\\smallskip\n\n{\\bf Case 6.} If $y_0=0$ and $\\ga_{p-1} = \\gb_{p-1} 0$ since otherwise $M(x)=0$, giving a contradiction.\nThus, in Case 6, we can choose $k - \\ga_{p-1}$ many\nindices $i$ from $N \\setminus I_{p-1}$ such that $x_{i{0}}=1$.\nLet $x'$ be a position obtained by the algorithm.\nThen, clearly $M(x')=1$ and $x \\to x'$ is a move. This completes the proof of (a1).\n\\qed\n\n\\section{More on the Tetris function} \\label{Ss.Tetris}\n\nIn this section we show that the SG function described in Theorems \\ref{thm.n<2k}\nand \\ref{thm.n=2k} can in fact be computed efficiently, in polynomial time. For this we need to show that the Tetris function can be computed in polynomial time for these games.\nWe also prove that for a given position $x$ and integer $0\\leq g < T_{n,k}(x)$\nwe can compute in polynomial time a move $x\\to x'$ such that $T_{n,k}(x')=g$.\nFinally, in subsection \\ref{tetris-degree} we recall\nsome relations to degree sequences of graphs and hypergraphs.\n\n\\subsection{Computing the Tetris function in polynomial time} \\label{comp-tet-in-p}\n\nLet us recall that to a position $x$ we associated a shifted position $\\bar x$ after Corollary \\ref{delta-epsilon} with the property that $T_{n,k}(x)=\\bar x_{n-k+1}$.\nThe procedure described there is a non-polynomial algorithm.\nHowever $\\bar x$ and consequently $T_{n,k}(x)=\\bar x_{n-k+1}$ can be computed in a more efficient way.\n\n\\begin{theorem}\\label{Tetrispolynomial}\nGiven a position $x$ we can compute $T_{n,k}(x)$ in linear time in $n$.\n\\end{theorem}\n\n\\proof\nWe can assume without loss of generality that $x$ is a nondecreasing position. We show that the corresponding $\\bar x$ can be constructed in linear time in $n$, and thus the claim follows by the equality $T_{n,k}(x)=\\bar x_{n-k+1}$.\n\nRecall that the input size is $\\log(\\prod_{i = 1}^n x_i)$. Let $s=\\sum_{i=1}^{n-k} x_i$ be the number of tokens we shift on top of the largest $k$ piles; see Figure \\ref{ADD(1)}.\nWe know that for some $\\ell n\/2$.\nWe examine the question of how to move to some position of given Tetris value $g$.\n\nWe denote by\n\\begin{enumerate} \\itemsep0em\n\\item [\\rm{(i)}] $x^\\ell$ the position obtained from $x$ by removing all tokens from the largest $k$ piles of $x$, and by\n\\item [\\rm{(ii)}] $x^u$ the position obtained from $x$ by decreasing the largest $k$ piles by one unit each.\n\\end{enumerate}\n\nConsider the set\n$W = \\{T_{n,k}(x') \\mid \\forall x': x\\to x'\\}$.\nBy Lemma \\ref{tbasic} (i) we have $T_{n,k}(x^\\ell) = \\min(W)$.\nBy Lemma \\ref{bestslowmove} we have $T_{n,k}(x^u)=T_{n,k}(x)-1$. As in the proof of\nTheorem \\ref{thm.n<2k}, we can argue that for\nevery value $g$ such that $T_{n,k}(x^\\ell)\\leq g \\leq T_{n,k}(x)-1$ there exists a move $x\\to x'$ such that $x^\\ell \\leq x' \\leq x$ and $T_{n,k}(x')=g$.\n\n\n\\begin{theorem}\\label{achieveTetrisvalue} \nGiven $g \\in W$, computing a position $x'$ such that $x^\\ell \\leq x' \\leq x$ and $T_{n,k}(x')=g$ can be done in $O\\big(n\\log(\\sum_{i=n-k+1}^nx_i)\\big)$ time.\n\\end{theorem}\n\\proof\nWe have $T_{n,k}(x^\\ell)\\leq g \\leq T_{n,k}(x^u)$.\nUsing the monotonicity of the Tetris function we perform a binary search in the space of positions between $x^\\ell$ and $x^u$.\nIn a general step we compute $L=\\sum_{i=n-k+1}^nx^\\ell_i$ and $U=\\sum_{i=n-k+1}^nx^u_i$, set $M=\\lfloor\\frac{L+U}{2}\\rfloor$ and compute $y_i=int(\\frac{x^\\ell_i+x^u_i}{2})$ for $i=n-k+1$,\nwhere $int(\\cdot)$ is a rounding to a nearest integer value in such a way that $\\sum_{i=n-k+1}^ny_i=M$.\nFinally, we set $y_i=x_i$ for $i0}$ (see, for instance, \\cite[Theorem\n II.6.1]{Te15}). There are two downsides of this method. First, the\nerror term is suboptimal under assumption of the GRH. Second, it would need some effort to make the constant $c$ actually explicit. \n\nHowever, as $|\\Delta|<10^{15}$, we are only interested in the case\nwhere $[a,b]$ is a subinterval of ${[1, 2 \\cdot 10^7]}$. In this\nrange, a simple \\textsf{SAGE} script using the \\textsf{MPFI} library\n\\cite{MPFI,sagemath} can be used to improve on\n\\eqref{equation::selbergdelange} computationally (see Proposition\n\\ref{proposition::selbergdelange}). As a consequence, we obtain\n$|\\Delta|<10^{10}$ for any singular unit of discriminant $\\Delta$ in Theorem \\ref{thmidrange}.\n\nThis is still not sufficient to check all remaining cases, at least\nwith modest computational means. The range is nevertheless small\nenough to use a counting algorithm in order to bound\n${\\mathcal{C}}_{10^{-3}}(\\Delta)$ for all discriminants $\\Delta$ satisfying\n$|\\Delta|< 10^{10}$, see Lemma \\ref{lem:lowrangeCebound}. {This still\nneeds an appropriate counting strategy, as determining\n${\\mathcal{C}}_{10^{-3}}(\\Delta)$ for each discriminant is rather slow,\ncomparable to computing separately each class number ${\\mathcal{C}}(\\Delta)$ in the\nsame range. Our trick is to bound all ${\\mathcal{C}}_{10^{-3}}(\\Delta)$\nsimultaneously by running through a set containing all imaginary quadratic\n$\\tau \\in \\mathcal{F}$ satisfying ${|\\tau-\\zeta_3|< \\varepsilon}$ or\n${|\\tau-\\zeta_6|< \\varepsilon}$ and such that $j(\\tau)$ is of discriminant\n$\\Delta$ with $|\\Delta|<10^{10}$. For each $\\tau$ encountered, we\ncompute its discriminant $\\Delta(\\tau)$ after the fact and increment\nour counter for ${\\mathcal{C}}_{10^{-3}}(\\Delta(\\tau))$. The thus obtained\nbounds for ${\\mathcal{C}}_{10^{-3}}(\\Delta)$ refine once again our previous\ninequalities, and allow us to conclude that $|\\Delta|<10^7$. Repeating this\nprocedure once again, with a slightly changed $\\varepsilon$, we\nachieve even $|\\Delta|<3 \\cdot 10^5$ in Theorem \\ref{thlowrange}.\nThese remaining cases can now be dealt with directly, for which we use\na \\textsf{PARI}~\\cite{PARI} program to prove Theorem \\ref{thm:habeggerspari}, completing thereby the proof of Theorem~\\ref{thmain}. \n\n\\bigskip\n\nIt is very probable that our argument can be adapted to solve a more general problem: given an algebraic integer~$\\beta$, determine the singular moduli~$\\alpha$ such that ${\\alpha-\\beta}$ is a unit; or at least bound effectively the discriminants of such~$\\alpha$. For instance, one may ask whether~$0$ is the only singular modulus~$\\alpha$ such that ${\\alpha-1}$ is a unit. In the general case, as explained in~\\cite{hab:junit}, this would require lower bounds for elliptic logarithmic forms, but when~$\\beta$ itself is a singular modulus, our argument extends almost without changes. One may go further and obtain an effective version of Theorem~2 from~\\cite{hab:junit}, which is an analogue of Siegel's Finiteness Theorem for\nspecial points.\n\n\nThe famous work of Gross-Zagier and Dorman \\cite{Dorman1988,Gross1985} inspires the following problem: determine all couples $(\\alpha,\\beta)$ of singular moduli such that ${\\alpha-\\beta}$ is a unit; presumably, there is none. As indicated above, when~$\\beta$ is fixed and~$\\alpha$ varying, a version of our argument does the job, but if we let both~$\\alpha$ and~$\\beta$ vary, the problem seems more intricate. Very recently Yingkun Li~\\cite{Li18} made important progress: he proved that ${\\alpha-\\beta}$ is not a unit if the discriminants of~$\\alpha$ and~$\\beta$ are fundamental and coprime. In particular, his result implies the following partial version of our Theorem~\\ref{thmain}: the discriminant of a singular unit must be either non-fundamental or divisible by~$3$. \n\nAnother natural problem is extending our work to $S$-units. Recall that, given a finite set~$S$ of prime numbers, a non-zero algebraic number is called an $S$-unit if both its denominator and numerator are composed of prime ideals dividing primes from~$S$. Recently\nHerrero,\nMenares and\nRivera-Letelier announced the proof of finiteness of the set of singular $S$-units (that is, singular moduli that are $S$-units) for any finite set of primes~$S$. However, to the best of our knowledge, their argument is not effective as of now. \n\n\n\\bigskip\n\nFinally, let us discuss an application of Theorem~\\ref{thmain} to\neffective results of Andr\\'e-Oort type. A point\n$(\\alpha_1,\\ldots,\\alpha_n) \\in \\mathbb{C}^n$ is called special if\neach $\\alpha_i$, $i \\in \\{ 1, \\dots, n\\}$, is a singular modulus. Since singular moduli are algebraic integers, the following statement is an immediate consequence of our main result. \n\n\n\n\n\\begin{corollary} For each polynomial ${P}$ in unknowns $X_2,\\ldots,\n X_n$ and coefficients that are algebraic integers in~$\\C$,\n \n the hypersurface defined by \n\\begin{equation*}\nX_1 P(X_1, \\dots, X_n) = 1\n\\end{equation*}\n contains no special points.\n\\end{corollary}\n\nIn particular, $\\alpha_1^{a_1}\\cdots \\alpha_n^{a_n}\\not=1$\nfor all special points $(\\alpha_1,\\ldots,\\alpha_n)$\nand all integers $a_1\\ge 1,\\ldots,a_n \\ge 1$. \nThis corollary exhibits a rather general class of algebraic varieties\nof arbitrary dimension and degree for which the celebrated theorem of\nPila~\\cite{Pila2011a} can be proved effectively and even explicitly.\nIt is complementary to other recent effective results of Andr\\'e-Oort type~\\cite{Bilu2017,Bi18}.\n\n\\paragraph{Plan of the article}\nIn Section~\\ref{sceps} we obtain an explicit version of the\nestimate~\\eqref{ecepsrough}. In Section~\\ref{supper} we obtain an\nupper estimate for the height of a singular unit. In\nSection~\\ref{slower} we obtain explicit versions of the lower estimates~\\eqref{ecolm} and~\\eqref{etriv}. In Section~\\ref{stenfourteen} we use all previous results to bound the discriminant of a singular unit as ${|\\Delta|<10^{15}}$. This bound is reduced to $10^{10}$ in Section~\\ref{sec:midrange} and to $3\\cdot10^5$ in Section~\\ref{slowrange}. Finally, in Section~\\ref{sfinal} we show that the discriminant of a singular unit satisfies ${|\\Delta|>3\\cdot10^5}$. \n\n\\paragraph{Convention}\nIn this article we fix, once and for all, an embedding ${\\bar\\Q\\hookrightarrow \\C}$; this means that all algebraic numbers in this article are viewed as elements of~$\\C$.\n\n\\paragraph{Acknowledgments}\nYuri Bilu was partially supported by the University of Basel, the Fields Institute (Toronto), and the Xiamen University. Lars K\u00fchne was supported by the Max-Planck Institute for Mathematics, the Fields Institute, and the Swiss National Science Foundation through an Ambizione grant. We thank Ricardo Menares and Amalia Pizarro for many useful conversations, Florian Luca and Aleksandar Ivic for helpful suggestions, Bill Allombert and Karim Belabas for a \\textsf{PARI} tutorial, and Jean-Louis Nicolas and Cyril Mauvillain for helping to access Robin's thesis~\\cite{Ro83a}. Finally, we thank both anonymous referees for encouraging reports and helpful suggestions. \n\n\n\\section{An estimate for \\texorpdfstring{${\\mathcal{C}}_\\varepsilon(\\Delta)$}{Ceps(Delta)}}\n\\label{sceps}\nLet~$\\Delta$ be a negative integer satisfying ${\\Delta\\equiv 0,1\\bmod 4}$ and \n$$\n{\\mathcal{O}}_\\Delta=\\Z[(\\Delta+\\sqrt\\Delta)\/2] \n$$\nthe imaginary quadratic order of discriminant~$\\Delta$. Then ${\\Delta=Df^2}$, where~$D$ is the discriminant of the imaginary quadratic field ${\\Q(\\sqrt\\Delta)}$ (the ``fundamental discriminant'') and ${f=[{\\mathcal{O}}_D:{\\mathcal{O}}_\\Delta]}$ is the conductor. We denote by ${\\mathcal{C}}(\\Delta)$ the class number of the order~${\\mathcal{O}}_\\Delta$. \n\n\n\nUp to $\\C$-isomorphism there exist ${\\mathcal{C}}(\\Delta)$ elliptic curves with CM by~${\\mathcal{O}}_\\Delta$. The $j$-invariants of these curves are called \\textsl{singular moduli} of discriminant~$\\Delta$.\nThe singular moduli of discriminant~$\\Delta$ form a full Galois orbit over~$\\Q$\nof cardinality ${\\mathcal{C}}(\\Delta)$, see \\cite[Proposition~13.2]{Cox}. \n\n\nLet~${\\mathcal{F}}$ be the standard fundamental domain in the Poincar\u00e9 plane, that is, the open hyperbolic triangle with vertices ${\\zeta_3,\\zeta_6,i\\infty}$, together with the geodesics $[i,\\zeta_6]$ and ${[\\zeta_6,i\\infty)}$; here \n$$\n\\zeta_3=e^{2\\pi i\/3}=\\frac{-1+\\sqrt{-3}}2, \\quad \\zeta_6=e^{\\pi i\/3}=\\frac{1+\\sqrt{-3}}2. \n$$\nEvery singular modulus can be uniquely presented as ${j(\\tau)}$, where ${\\tau\\in {\\mathcal{F}}}$. \n\n\nNow fix ${\\varepsilon\\in (0,1\/3]}$ and denote by ${\\mathcal{C}}_\\varepsilon(\\Delta)$ the number of singular moduli of discriminant~$\\Delta$ that can be presented as ${j(\\tau)}$ where ${\\tau\\in {\\mathcal{F}}}$ satisfies\n\\begin{equation}\n\\label{etauzeze}\n\\min \\{|\\tau-\\zeta_3|,|\\tau-\\zeta_6|\\}< \\varepsilon. \n\\end{equation}\nIn this section we bound this quantity.\n\n\nDefine the \\textsl{modified conductor}~${\\tilde f}$ by\n\\begin{equation}\n\\label{etilf}\n{\\tilde f}=\n\\begin{cases}\nf,& D\\equiv 1\\bmod 4,\\\\\n2f, &D\\equiv 0\\bmod 4. \n\\end{cases}\n\\end{equation}\nThen ${\\Delta\/{\\tilde f}^2}$ is a square-free integer. \n\n\n\n\n\\begin{theorem}\n\\label{thceps}\nFor ${\\varepsilon\\in (0,1\/3]}$ we have\n\\begin{equation}\n\\label{enewbound}\n{\\mathcal{C}}_\\varepsilon(\\Delta) \\le F\\left( \\frac{16}3\\frac{\\sigma_1({\\tilde f})}{{\\tilde f}}|\\Delta|^{1\/2}\\varepsilon^2+\\frac83|\\Delta|^{1\/2}\\varepsilon+ 8|\\Delta\/3|^{1\/4}\\sigma_0({\\tilde f})\\varepsilon+4\\right),\n\\end{equation}\nwhere \n\\begin{equation}\n\\label{ecapitalf}\nF=F(\\Delta) = \\max \\bigl\\{2^{\\omega(a)}: a\\le |\\Delta|^{1\/2}\\bigr\\}.\n\\end{equation}\n\\end{theorem}\n\\begin{corollary}\n\\label{cseps}\nIn the set-up of Theorem~\\ref{thceps} assume that ${|\\Delta|\\ge 10^{14}}$. Then \n\\begin{equation}\n\\label{enewboundsimple}\n{\\mathcal{C}}_\\varepsilon(\\Delta) \\le F\\left( 9.83|\\Delta|^{1\/2} \\varepsilon^2\\log\\log(|\\Delta|^{1\/2})+3.605|\\Delta|^{1\/2}\\varepsilon+ 4\\right).\n\\end{equation}\n\\end{corollary}\n\n\n\\subsection{Some lemmas}\nWe need some lemmas. For a prime number~$\\ell$ and a non-zero integer~$n$ we denote by ${\\mathrm{ord}}_\\ell(n)$ the $\\ell$-adic order of~$n$; that is, ${\\ell^{{\\mathrm{ord}}_\\ell(n)}\\,\\|\\,n}$. \n\n\\begin{lemma}\n\\label{lrootsmodle}\nLet~$\\ell$ be a prime number, ${e \\ge 1}$ an integer, and~$\\Delta$ a\nnon-zero integer with ${\\nu={\\mathrm{ord}}_\\ell\\Delta}$.\nThen the set of ${b\\in \\Z}$ satisfying \n${b^2\\equiv \\Delta\\bmod \\ell^e}$ \nis a union of at most~$2$ residue classes modulo ${\\ell^{e-\\lfloor\n \\min\\{e,\\nu\\}\/2\\rfloor}}$ in all cases except when ${\\ell=2}$ and\n${e\\ge 3}$; in this latter case it is a union of most~$4$ such\nclasses. Finally, the set of $b$ equals a single residue class modulo ${\\ell^{e-\\lfloor\n \\min\\{e,\\nu\\}\/2\\rfloor}}$\nif $\\nu\\ge e$. \n\\end{lemma}\n\\begin{proof}\nWe suppose first that ${\\nu=0}$, that is, ${\\ell\\nmid \\Delta}$. In this case we have to count the number of elements in the multiplicative group $(\\Z\/\\ell^e\\Z)^\\times$ whose\nsquare is represented by~$\\Delta$. If ${\\ell\\ge 3}$ or ${\\ell^e\\in \\{2,4\\}}$, then $(\\Z\/\\ell^e\\Z)^\\times$ is a cyclic group.\nThen there are at most~$2$ square roots and this implies our claim. If ${\\ell=2}$ and ${e \\ge 3}$, then ${(\\Z\/2^e\\Z)^\\times\\cong \\Z\/2\\Z\\times \\Z\/2^{e-2}\\Z}$, and there are at most~$4$ square roots, as desired.\n\nNow assume that ${\\nu {|\\Delta\/3|^{1\/4}}. \n\\end{cases}\n$$\nHence\n\\begin{align}\n\\sum_{d^2\\mid \\Delta}d\\cdot \\#(I\\cap d^2\\Z) &\\le \\sum_{\\genfrac{}{}{0pt}{}{d\\mid{\\tilde f}}{d\\le|\\Delta\/3|^{1\/4}}}d\\left(\\frac23\\frac{|\\Delta|^{1\/2}}{d^2}\\varepsilon+1\\right)\\nonumber\\\\\n&\\le \\frac23|\\Delta|^{1\/2}\\varepsilon\\sum_{d\\mid{\\tilde f}}d^{-1}+ \\sum_{\\genfrac{}{}{0pt}{}{d\\mid{\\tilde f}}{d\\le |\\Delta\/3|^{1\/4}}}d\\nonumber\\\\\n\\label{estim3}\n&\\le \\frac23\\frac{\\sigma_1({\\tilde f})}{{\\tilde f}}|\\Delta|^{1\/2}\\varepsilon+ |\\Delta\/3|^{1\/4}\\sigma_0({\\tilde f}). \n\\end{align}\nFinally, Lemma~\\ref{ltrivi} implies that\n\\begin{equation}\n\\label{eintini}\n\\#(I\\cap\\Z)\\le \\frac23|\\Delta|^{1\/2}\\varepsilon+1. \n\\end{equation}\nPutting the estimates~\\eqref{estim1},~\\eqref{estim2},~\\eqref{estim3} and~\\eqref{eintini} together, we obtain~\\eqref{enewbound}. \\qed\n\n\\subsection{Proof of Corollary~\\ref{cseps}}\n\nWe need to estimate $\\sigma_0({\\tilde f})$ and $\\sigma_1({\\tilde f})$ in terms of~$|\\Delta|$. The following lemma uses a simple estimate for $\\sigma_0(n)$ due to Nicolas and Robin~\\cite{NR83}. Much sharper estimates can be found in Robin's thesis~\\cite{Ro83a}. \n\n\\begin{lemma}\n\\label{lsigzersimple}\nFor ${|\\Delta|\\ge 10^{14}}$ we have \n\\begin{align}\n\\label{euppersigzer}\n\\sigma_0({\\tilde f})&\\le |\\Delta|^{0.192},\\\\\n\\label{euppersigone}\n{\\sigma_1({\\tilde f})}\/{\\tilde f} &\\le 1.842\\log\\log(|\\Delta|^{1\/2}). \n\\end{align}\n\\end{lemma}\n\n\\begin{proof}\nFor proving~\\eqref{euppersigzer} may assume that ${{\\tilde f}\\ge 16}$, otherwise there is nothing to prove. In~\\cite{NR83} it is proved that for ${n\\ge 3}$ we have\n$$\n\\frac{\\log\\sigma_0(n)}{\\log2} \\le 1.538\\frac{\\log n}{\\log\\log n}.\n$$\nThe function ${x\\mapsto (\\log x)\/(\\log\\log x)}$ is increasing for ${x\\ge 16}$. Since \n$$\n|\\Delta|\\ge 10^{14}, \\qquad 16\\le {\\tilde f}\\le |\\Delta|^{1\/2},\n$$ \nthis gives\n\\begin{align*}\n\\log\\sigma_0({\\tilde f}) \n&\\le 1.538\\log 2\\frac{\\log(|\\Delta|^{1\/2})}{\\log\\log(|\\Delta|^{1\/2})}\\\\\n&\\le \\frac{1.538}2\\log 2\\frac{\\log|\\Delta|}{\\log\\log(10^{7})}\\\\\n&<0.192\\log|\\Delta|,\n\\end{align*}\nas wanted. \n\nFor proving~\\eqref{euppersigone} we use the estimate \n${\\sigma_1(n)\\le 1.842n\\log\\log n}$\nwhich holds for ${n\\ge 121}$, see \\cite[Theorem~1.3]{AFJ07}. This\nproves~\\eqref{euppersigone} for ${{\\tilde f}\\ge 121}$. For ${{\\tilde f}\\le 120}$\none can check directly that ${\\sigma_1({\\tilde f})\/{\\tilde f}\\le 3}$ so that inequality~\\eqref{euppersigone} is also\ntrue in this case.\n\\end{proof}\n\n\n\\begin{proof}[Proof of Corollary~\\ref{cseps}]\nIf ${|\\Delta|\\ge 10^{14}}$ then Lemma~\\ref{lsigzersimple} implies that \n\\begin{align*}\n&8\\left|\\frac\\Delta3\\right|^{1\/4}\\sigma_0({\\tilde f}) \\le \\frac8{3^{1\/4}}|\\Delta|^{0.442}\\le \\frac8{3^{1\/4}\\cdot10^{0.812}}|\\Delta|^{1\/2}\\le 0.938|\\Delta|^{1\/2},\\\\\n&\\frac{16}3\\frac{\\sigma_1({\\tilde f})}{{\\tilde f}} \\le \\frac{16}3\\cdot 1.842 \\log\\log(|\\Delta|^{1\/2})\\le9.83 \\log\\log(|\\Delta|^{1\/2}).\n\\end{align*}\nSubstituting all this to~\\eqref{enewbound}, we obtain~\\eqref{enewboundsimple}. \n\\end{proof}\n\n\n\n\\section{An upper bound for the height of a singular unit}\n\\label{supper}\n\nIn this section we obtain a fully explicit version of estimate~\\eqref{eupperheight}.\nWe use the notation ${\\mathcal{C}}(\\Delta)$, ${\\mathcal{C}}_\\varepsilon(\\Delta)$, ${\\mathcal{F}}$, $\\zeta_3$, $\\zeta_6$ introduced in Section~\\ref{sceps}. \n\nLet~$\\alpha$ be a complex algebraic number of degree~$m$ whose minimal polynomial over~$\\Z$ is \n$$\nP(x)=a_mx^m+\\cdots+a_0=a_m(x-\\alpha_1)\\cdots (x-\\alpha_m) \\in \\Z[x].\n$$\nHere ${\\gcd(a_0,a_1, \\ldots, a_m)=1}$ and ${\\alpha_1, \\ldots, \\alpha_m\\in \\C}$ are the conjugates of~$\\alpha$ over~$\\Q$. Then the height of~$\\alpha$ is defined by\n$$\n{\\mathrm{h}}(\\alpha) =\\frac1m\\left(\\log|a_m|+\\sum_{k=1}^m \\log^+|\\alpha_k|\\right), \n$$ \nwhere ${\\log^+(\\cdot)=\\log\\max\\{1,\\cdot\\}}$. \nIf~$\\alpha$ is an algebraic integer then \n$$\n{\\mathrm{h}}(\\alpha) =\\frac1m\\sum_{k=1}^m \\log^+|\\alpha_k|. \n$$ \nIt is known that ${{\\mathrm{h}}(\\alpha)={\\mathrm{h}}(\\alpha^{-1})}$ when ${\\alpha\\ne 0}$. \n\n\\begin{theorem}\n\\label{theceps}\nLet~$\\alpha$ be a singular unit of discriminant~$\\Delta$, and~$\\varepsilon$ a real number satisfying ${0<\\varepsilon \\le 4\\cdot10^{-3}}$. \nThen \n\\begin{equation}\n\\label{eupperheightbis}\n{\\mathrm{h}}(\\alpha)\\le 3\\frac{{\\mathcal{C}}_\\varepsilon(\\Delta)}{{\\mathcal{C}}(\\Delta)}\\log|\\Delta|+3\\log(\\varepsilon^{-1})-10.66. \n\\end{equation}\n\\end{theorem}\n\nCombining this with Corollary~\\ref{cseps} and optimizing~$\\varepsilon$, we obtain the following consequence. \n\n\\begin{corollary}\n\\label{cupper}\nIn the set-up of Theorem~\\ref{theceps} assume that ${|\\Delta|\\ge 10^{14}}$. Then\n\\begin{align}\n\\label{eupdel}\n{\\mathrm{h}}(\\alpha)&\\le \\frac{12A}{{\\mathcal{C}}(\\Delta)}+3\\log\\frac{A|\\Delta|^{1\/2}}{{\\mathcal{C}}(\\Delta)}-3.77,\n\\end{align}\nwhere ${A= F\\log|\\Delta|}$ and~$F$ is defined in~\\eqref{ecapitalf}. \n\\end{corollary}\n\n\n\n\n\n\\subsection{Proof of Theorem~\\ref{theceps}}\n\nWe start from some simple lemmas.\n\n\n\\begin{lemma}\n\\label{lanal}\nFor ${z\\in {\\mathcal{F}}}$ we have \n$$\n|j(z)|\\ge 42700\\bigl(\\min\\{|z-\\zeta_3|,|z-\\zeta_6|,4\\cdot10^{-3}\\}\\bigr)^3. \n$$\n\\end{lemma}\n\n\\begin{proof}\nThis is an easy modification of Proposition~2.2 from~\\cite{BLP16}; just replace therein $10^{-3}$ by ${4\\cdot10^{-3}}$. \n\\end{proof}\n\nIn the next lemma we use the notation $T_\\Delta$ and ${\\tau(a,b,c)}$ introduced before Lemma~\\ref{lgauss}. \n\n\\begin{lemma}\n\\label{lliouv}\nAssume that ${\\Delta\\ne -3}$. Let ${\\tau=\\tau(a,b,c)}$, where ${(a,b,c)\\in T_\\Delta}$. Let~$\\zeta$ be one of the numbers~$\\zeta_3$ or~$\\zeta_6$. \nThen\n$$\n|\\tau-\\zeta|\\ge \\frac{\\sqrt3}{4|\\Delta|}. \n$$\n\\end{lemma}\n\n\\begin{proof}\nWe have\n$$\n|\\tau-\\zeta|\\ge|\\Im\\tau-\\Im\\zeta|= \\left|\\frac{\\sqrt{|\\Delta|}}{2a}-\\frac{\\sqrt3}{2}\\right|= \\frac{\\bigl||\\Delta|-3a^2\\bigr|}{2a(\\sqrt{|\\Delta|}+a\\sqrt3)}.\n$$\nSince ${\\Delta\\ne -3}$ we have ${\\Delta\\ne -3a^2}$, see item~\\ref{iadelta} of Lemma~\\ref{lgauss}. Hence\n$$\n|\\tau-\\zeta|\\ge \\frac{1}{2a(\\sqrt{|\\Delta|}+a\\sqrt3)}\\ge \\frac{\\sqrt3}{4|\\Delta|}, \n$$\nthe last inequality being again by item~\\ref{iadelta} of Lemma~\\ref{lgauss}. \n\\end{proof}\n\nNow we are ready to prove Theorem~\\ref{theceps}. \n\n\\begin{proof}[Proof of Theorem~\\ref{theceps}]\nLet ${\\alpha=\\alpha_1, \\alpha_2, \\ldots, \\alpha_m\\in \\C}$ be the conjugates of~$\\alpha$ over~$\\Q$. Then ${m={\\mathcal{C}}(\\Delta)}$ and ${\\alpha_1, \\ldots, \\alpha_m}$ is the full list of singular moduli of discriminant~$\\Delta$. Write them as ${j(\\tau_1), \\ldots, j(\\tau_m)}$, where ${\\tau_1, \\ldots, \\tau_m\\in {\\mathcal{F}}}$. \n\nSince~$\\alpha$ is a unit, we have \n$$\n{\\mathrm{h}}(\\alpha)={\\mathrm{h}}(\\alpha^{-1}) = \\frac1m \\sum_{k=1}^m\\log^+|\\alpha_k^{-1}|,\n$$\nHence \n{\\footnotesize\n\\begin{align}\n{\\mathrm{h}}(\\alpha) &= \\frac{1}{{\\mathcal{C}}(\\Delta)}\\sum_{k=1}^m\\log^+|j(\\tau_k)^{-1}|\\nonumber\\\\\n\\label{etwosums}\n&= \\frac{1}{{\\mathcal{C}}(\\Delta)}\\left(\\sum_{\\genfrac{}{}{0pt}{}{1\\le k\\le m}{\\min\\{|\\tau_k-\\zeta_3|, |\\tau_k-\\zeta_6|\\}< \\varepsilon}}+\\sum_{\\genfrac{}{}{0pt}{}{1\\le k\\le m}{\\min\\{|\\tau_k-\\zeta_3|, |\\tau_k-\\zeta_6|\\}\\ge \\varepsilon}}\\right)\\log^+|j(\\tau_k)^{-1}|.\n\\end{align}\n}%\nWe estimate each of the two sums separately.\n\nSince ${\\varepsilon\\le 4\\cdot10^{-3}}$, Lemma~\\ref{lanal} implies that each term in the second sum satisfies \n$$\n\\log^+|j(\\tau_k)^{-1}| \\le 3\\log(\\varepsilon^{-1})-\\log42700 \\le 3\\log(\\varepsilon^{-1})-10.66. \n$$\nHence\n$$\n\\sum_{\\genfrac{}{}{0pt}{}{1\\le k\\le m}{\\min\\{|\\tau_k-\\zeta_3|, |\\tau_k-\\zeta_6|\\}\\ge \\varepsilon}}\\log^+|j(\\tau_k)^{-1}|\\le ({\\mathcal{C}}(\\Delta)-{\\mathcal{C}}_\\varepsilon(\\Delta))\\bigl(3\\log(\\varepsilon^{-1})-10.66\\bigr). \n$$\nSince ${\\varepsilon\\le 4\\cdot10^{-3}}$ we have\n ${3\\log(\\varepsilon^{-1})>10.66}$, which implies that \n\\begin{equation}\n\\label{esecondsum}\n\\sum_{\\genfrac{}{}{0pt}{}{1\\le k\\le m}{\\min\\{|\\tau_k-\\zeta_3|, |\\tau_k-\\zeta_6|\\}\\ge \\varepsilon}}\\log^+|j(\\tau_k)^{-1}|\\le {\\mathcal{C}}(\\Delta)\\bigl(3\\log(\\varepsilon^{-1})-10.66\\bigr). \n\\end{equation}\nAs for the first sum, Lemmas~\\ref{lanal} and~\\ref{lliouv} imply that each term in this sum satisfies\n$$\n\\log^+|j(\\tau_k)^{-1}| \\le \\max\\left\\{0, 3\\log\\frac{4|\\Delta|}{\\sqrt3}-\\log42700\\right\\} \\le 3\\log|\\Delta|. \n$$ \nNote that we may use here Lemma~3.4 because the only singular modulus of discriminant $-3$ is~$0$, which is not a unit. \n\nSince the first sum has ${\\mathcal{C}}_\\varepsilon(\\Delta)$ terms, this implies the estimate \n\\begin{equation}\n\\label{efirstsum}\n\\sum_{\\genfrac{}{}{0pt}{}{1\\le k\\le m}{\\min\\{|\\tau_k-\\zeta_3|, |\\tau_k-\\zeta_6|\\}< \\varepsilon}}\\log^+|j(\\tau_k)^{-1}|\\le 3{\\mathcal{C}}_\\varepsilon(\\Delta)\\log|\\Delta|. \n\\end{equation}\nSubstituting~\\eqref{esecondsum} and~\\eqref{efirstsum} into~\\eqref{etwosums}, we obtain~\\eqref{eupperheightbis}.\n\\end{proof}\n\n\n\\subsection{Proof of Corollary~\\ref{cupper}}\n\nTo prove the corollary we need a lower bound for the quantity~$F$ defined in Theorem~\\ref{thceps} and an upper bound for the class number ${\\mathcal{C}}(\\Delta)$. \n\n\\begin{lemma}\n\\label{lf}\nAssume that ${|\\Delta|\\ge 10^{14}}$. Then ${F\\ge |\\Delta|^{0.34\/\\log\\log(|\\Delta|^{1\/2})}}$ and ${F\\ge 18.54\\log\\log(|\\Delta|^{1\/2})}$. \n\\end{lemma}\n\n\\begin{proof}\nDefine, as usual \n\\begin{equation}\n\\label{ethetapi}\n\\vartheta(x)=\\sum_{p\\le x}\\log p, \\qquad \\pi(x)=\\sum_{p\\le x}1. \n\\end{equation}\nThen \n\\begin{align}\n\\vartheta(x) & \\le 1.017 x &&(x>0), \\nonumber\\\\\n\\label{elowerpi}\n\\pi(x) &\\ge \\frac{x}{\\log x} && (x\\ge 17),\n\\end{align} \nsee \\cite{RS62}, Theorem~9 on page~71 and Corollary~1 after Theorem~2 on page~69. Estimate~\\eqref{elowerpi} implies that\n\\begin{equation}\n\\label{elowerpibis}\n\\pi(x) \\ge 0.99995\\frac{x}{\\log x} \\qquad (x\\ge 13).\n\\end{equation}\nSetting here \n$$\nx=\\frac{\\log(|\\Delta|^{1\/2})}{1.017} , \\qquad N =\\prod_{p\\le x}p,\n$$\nwe obtain ${N\\le |\\Delta|^{1\/2}}$ and \n$$\n\\omega(N) =\\pi(x) \\ge \\frac{0.99995\\log(|\\Delta|^{1\/2})}{1.017\\log\\log(|\\Delta|^{1\/2})}. \n$$\nNote that ${x\\ge \\bigl(\\log(10^{7})\\bigr)\/1.017>15}$, so we are allowed to use~\\eqref{elowerpibis}. \nWe obtain \n$$\nF\\ge 2^{\\omega(N)} \\ge |\\Delta|^{\\frac{0.99995\\log2}{2\\cdot1.017\\log\\log(|\\Delta|^{1\/2})}}\\ge |\\Delta|^{0.34\/\\log\\log(|\\Delta|^{1\/2})},\n$$\nproving the first estimate.\n\n\nTo prove the second estimate, we deduce from the first estimate that \n\\begin{equation}\n\\label{elog-log}\n\\log F-\\log\\log\\log(|\\Delta|^{1\/2})\\ge 0.68 \\frac{u}{\\log u}-\\log\\log u,\n\\end{equation} \nwhere we set ${u=\\log(|\\Delta|^{1\/2})}$. The right-hand side of~\\eqref{elog-log}, viewed as a function in~$u$, is increasing for ${u\\ge \\log(10^{7})}$. Hence \n$$\n\\log F-\\log\\log\\log(|\\Delta|^{1\/2})\\ge 0.68 \\frac{\\log(10^{7})}{\\log \\log(10^{7})}-\\log\\log \\log(10^{7}) \\ge 2.92,\n$$\nand ${F\\ge e^{2.92}\\log\\log(|\\Delta|^{1\/2}) \\ge 18.54 \\log\\log(|\\Delta|^{1\/2})}$. \n\\end{proof}\n\n\\begin{lemma}\n\\label{lcd}\nFor ${\\Delta\\ne -3,-4}$ we have \n$$\n{\\mathcal{C}}(\\Delta) \\le \\pi^{-1}|\\Delta|^{1\/2}(2+\\log|\\Delta|).\n$$ \n\\end{lemma}\n\n\\begin{proof}\nThis follows from Theorems~10.1 and~14.3 in \\cite[Chapter~12]{Hu82}. Note that in~\\cite{Hu82} the right-hand side has an extra factor $\\omega\/2$, where~$\\omega$ is the number of roots of unity in the imaginary quadratic order of discriminant~$\\Delta$. Since we assume that ${\\Delta \\ne -3,-4 }$, we have ${\\omega=2}$, so we may omit this factor. \n\\end{proof}\n\n\n\\begin{proof}[Proof of Corollary~\\ref{cupper}]\nSubstituting the estimate for ${\\mathcal{C}}_\\varepsilon(\\Delta)$ from~\\eqref{enewboundsimple} into~\\eqref{eupperheightbis}, we obtain the estimate\n\\begin{align*}\n{\\mathrm{h}}(\\alpha)&\\le 3A\\frac{9.83|\\Delta|^{1\/2} \\varepsilon^2\\log\\log(|\\Delta|^{1\/2})+3.605|\\Delta|^{1\/2}\\varepsilon+ 4}{{\\mathcal{C}}(\\Delta)}\n\\\\&\\hphantom{\\le}\n+3\\log(\\varepsilon^{-1})-10.66\n\\end{align*}\nwith ${A=F\\log|\\Delta|}$. Specifying \n$$\n\\varepsilon=0.27\\frac{{\\mathcal{C}}(\\Delta)}{A|\\Delta|^{1\/2}}\n$$\n(this is a nearly optimal value, and it satisfies ${\\varepsilon\\le 4\\cdot10^{-3}}$ as verified below), we obtain, using Lemmas~\\ref{lf} and~\\ref{lcd},\n\\begin{align*}\n{\\mathrm{h}}(\\alpha)&\\le 3\\cdot 9.83 \\cdot (0.27)^2\\frac{\\log\\log(|\\Delta|^{1\/2})}F\\frac{{\\mathcal{C}}(\\Delta)}{|\\Delta|^{1\/2}\\log|\\Delta|}+3\\cdot3.605\\cdot0.27\\\\ \n&\\hphantom{\\le}+\\frac{12A}{{\\mathcal{C}}(\\Delta)} +3\\log\\frac{A|\\Delta|^{1\/2}}{{\\mathcal{C}}(\\Delta)}-3\\log0.27-10.66\\\\\n&\\le \\frac{12A}{{\\mathcal{C}}(\\Delta)} +3\\log\\frac{A|\\Delta|^{1\/2}}{{\\mathcal{C}}(\\Delta)}+ \\frac{3\\cdot 9.83 \\cdot (0.27)^2\\cdot0.34}{18.54}+3\\cdot3.605\\cdot0.27\\\\ \n&\\hphantom{\\le}-3\\log0.27-10.66\\\\\n&\\le \\frac{12A}{{\\mathcal{C}}(\\Delta)} +3\\log\\frac{A|\\Delta|^{1\/2}}{{\\mathcal{C}}(\\Delta)}-3.77,\n\\end{align*}\nas wanted. \n\n \nWe only have to verify that ${\\varepsilon\\le 4\\cdot10^{-3}}$. We have\n${F\\ge 256}$ when ${|\\Delta|\\ge 10^{14}}$. Using Lemma~\\ref{lcd}, we obtain \n\\begin{align*}\n\\varepsilon=\\frac{0.27{\\mathcal{C}}(\\Delta)}{|\\Delta|^{1\/2}\\log |\\Delta|}\\frac1F\n\\le 0.27\\pi^{-1}\\frac{2+\\log(10^{14})}{\\log(10^{14})}\\cdot \\frac1{256}\n<4\\cdot 10^{-4}.\n\\end{align*}\nThe proof is complete.\n\\end{proof}\n\n\n\n\n\\section{Lower bounds for the height of a singular modulus}\n\\label{slower}\n\nNow we establish explicit lower bounds of the form~\\eqref{ecolm} and~\\eqref{etriv}. \n\n\\subsection{The ``easy'' bound}\n\nWe start by proving a bound of the form~\\eqref{etriv}. \n\n\\begin{proposition}\n\\label{ptriv}\nLet~$\\alpha$ be a singular modulus of discriminant~$\\Delta$. \nAssume that ${|\\Delta|\\ge 16}$. Then \n\\begin{equation}\n\\label{etrivbis}\n{\\mathrm{h}}(\\alpha) \\ge \\frac{\\pi|\\Delta|^{1\/2}-0.01}{{\\mathcal{C}}(\\Delta)}.\n\\end{equation}\n\\end{proposition}\n\n\nWe need a simple lemma. \n\n\\begin{lemma}\n\\label{lttsn}\nFor ${z\\in {\\mathcal{F}}}$ with imaginary part~$y$ we have \n$$\n\\bigl||j(z)|-e^{2\\pi y}\\bigr|\\le 2079.\n$$ \nIf ${y\\ge 2}$ then we also have ${|j(z)|\\ge 0.992 e^{2\\pi y}}$. \n\\end{lemma}\n\n\\begin{proof}\nThe first statement is Lemma~1 of~\\cite{BMZ13}, and the second one is an immediate consequence. \n\\end{proof}\n\n\n\n\\begin{proof}[Proof of Proposition~\\ref{ptriv}]\nOne of the conjugates of~$\\alpha$ over~$\\Q$ is equal to ${j((b+\\sqrt\\Delta)\/2)}$, with ${b=1}$ for~$\\Delta$ odd, and ${b=0}$ for~$\\Delta$ even; it corresponds to the element ${(1,b,(-\\Delta+b^2)\/4)}$ of the set $T_\\Delta$. \nHence\n$$\n{\\mathrm{h}}(\\alpha) \\ge \\frac{\\log|j((b+\\sqrt\\Delta)\/2)|}{{\\mathcal{C}}(\\Delta)}.\n$$\nUsing Lemma~\\ref{lttsn}, we obtain\n$$\n\\log|j((b+\\sqrt\\Delta)\/2)|\\ge \\pi|\\Delta|^{1\/2}+\\log 0.992 \\ge \\pi|\\Delta|^{1\/2}-0.01.\n$$\nWhence the result. \n\\end{proof}\n\n\\subsection{The ``hard'' bound}\n\n\\label{sshard}\n\nWe are left with bound~\\eqref{ecolm}. We are going to prove the following. \n\n\\begin{proposition}\n \\label{phlb2}\nLet~$\\alpha$ be a singular modulus of discriminant~$\\Delta$.\nThen\n\\begin{equation}\n\\label{ecolmbis}\n{\\mathrm{h}}(\\alpha) \\ge \\frac{3}{\\sqrt{5}}\\log|\\Delta| -9.79. \n\\end{equation}\n\\end{proposition}\n\n\nThe proof of Proposition~\\ref{phlb2} relies on the\nfact that it is possible to evaluate the Faltings height of an elliptic curve\nwith complex multiplication precisely, due to the work of Colmez~\\cite{Colmez} and Nakkajima-Taguchi~\\cite{NT}; for an exact statement see \\cite[Lemma 4.1]{Ha10}. \n\nLet~$E$ be an elliptic curve with CM by an order of discriminant~$\\Delta$. \nWe let ${\\mathrm{h}}_F(E)$ denote the stable Faltings height of~$E$ (using Deligne's normalization~\\cite{De85}). \nThe above-mentioned explicit formula for ${\\mathrm{h}}_F(E)$ is used in~\\cite{HJM:6} to obtain the lower bound \n\\begin{equation}\n\\label{efaltlowerold}\n{\\mathrm{h}}_F(E) \\ge \\frac{1}{4\\sqrt5} \\log|\\Delta| - 5.93,\n\\end{equation}\nsee Lemma 14(ii) therein. Unfortunately, this bound is numerically too weak for our purposes.\n\nProposition~\\ref{phlb2} will be deduced from the following numerical refinement of~\\eqref{efaltlowerold}.\n\n\\begin{proposition}\n\\label{pfaltlower}\nLet~$E$ be an elliptic curve with CM by an order of discriminant~$\\Delta$.\nThen \n\\begin{equation}\n\\label{efaltlowernew}\n{\\mathrm{h}}_F(E) \\ge \\frac{1}{4\\sqrt5} \\log|\\Delta| -\\gamma -\\frac{\\log(2\\pi)}{2}-\\left(\\frac1{2\\sqrt5}-\\frac16\\right)\\log2,\n\\end{equation}\nwhere ${\\gamma=0.57721\\ldots}$ is the Euler constant. \n\\end{proposition}\n\n\nLet us first show how Proposition~\\ref{pfaltlower} implies Proposition~\\ref{phlb2}. \n\n\n\\begin{proof}[Proof of Proposition~\\ref{phlb2} (assuming Proposition~\\ref{pfaltlower})]\nLet~$E$ be an elliptic curve with ${j(E)=\\alpha}$. \nWe only need to relate ${\\mathrm{h}}_F(E)$ to ${\\mathrm{h}}(j(E))$. For this purpose we use Lemma~7.9 of Gaudron and R\\'emond~\\cite{GaudronRemond:periodes}\\footnote{The reader should be warned that our ${\\mathrm{h}}_F(E)$ is denoted ${\\mathrm{h}}(E)$ in~\\cite{GaudronRemond:periodes}.}.\nIn our\nnotation they\nshow that \n\\begin{equation}\n\\label{egr}\n{\\mathrm{h}}_F(E)\\le {\\mathrm{h}}(j(E))\/12 - 0.72 \n\\end{equation}\nA quick calculation yields our\nclaim. \n\\end{proof}\n\n\n\nTo prove Proposition~\\ref{pfaltlower} we need a technical lemma.\nSet \n\\begin{equation}\n\\label{elambda}\n\\lambda=\\frac12-\\frac1{2\\sqrt5},\n\\end{equation}\nand define the additive arithmetical functions $\\beta(n)$ and $\\delta(n)$ by \n\\begin{equation}\n\\label{ebetadelta}\n\\beta(p^k) =\\frac{\\log p}{p+1}\\frac{1-p^{-k}}{1-p^{-1}}, \\quad \\beta(n)=\\sum_{p^k\\|n}\\beta(p^k), \\quad \\delta(n)= \\lambda\\log n -\\beta(n).\n\\end{equation}\n\n\\begin{lemma}\n\\label{ltechn} \nFor every positive integer~$n$ we have \n$$\n\\delta(n) \\ge\\delta(2)= \\left(\\frac16-\\frac1{2\\sqrt5}\\right)\\log2. \n$$\n\\end{lemma}\n\n\\begin{proof}\nSince ${1\/3>\\lambda>1\/4}$, we have ${\\delta(2)<0}$ and ${\\delta(p)>0}$ for all primes ${p\\ge 3}$. Also, for ${k\\ge 1}$ and any prime~$p$ we have \n$$\n\\delta(p^{k+1})-\\delta(p^k)= \\left(\\lambda-\\frac1{p^k(p+1)}\\right)\\log p >0.\n$$\nSince ${\\delta(4)>0}$, this proves that ${\\delta(p^k)>0}$ for every prime power ${p^k\\ne 2}$, whence the result. \n\\end{proof}\n\nProposition~\\ref{pfaltlower} is an immediate consequence of Lemma~\\ref{ltechn} and the following statement. \n\n\\begin{proposition}\n\\label{pfaltlowerf}\nIn the set-up of Proposition~\\ref{pfaltlower} we have\n\\begin{equation}\n\\label{efaltlowernewf}\n{\\mathrm{h}}_F(E) \\ge \\frac{1}{4\\sqrt5} \\log|\\Delta| + \\lambda \\log f- \\beta(f) -\\gamma -\\frac{\\log(2\\pi)}{2}.\n\\end{equation}\n\\end{proposition}\nSince \n$$\n\\lambda \\log f- \\beta(f)\\ge-\\left(\\frac1{2\\sqrt5}-\\frac16\\right)\\log2\n$$\nby Lemma~\\ref{ltechn}, this implies Proposition~\\ref{pfaltlower}. \n\n\\begin{proof}[Proof of Proposition~\\ref{pfaltlowerf}]\nWrite ${\\Delta=Df^2}$ with~$D$ the fundamental discriminant and~$f$ the conductor. Define \n$$\ne_f(p) = \\frac{1-\\chi(p)}{p-\\chi(p)}\\frac{1-p^{-{\\mathrm{ord}}_p(f)}}{1-p^{-1}}, \\qquad c(f) = \\frac 12 \\left(\\sum_{p\\mid f} e_f(p)\\log p\\right),\n$$\nwhere ${\\chi(\\cdot)=(D\/\\cdot)}$ is Kronecker's symbol.\n\n\nIn the proof of Lemma~14 of~\\cite{HJM:6}\\footnote{Note that our~$D$ is written~$\\Delta$ in~\\cite{HJM:6}.},\nthe stable Faltings height of~$E$ is estimated as \n\\begin{align*}\n{\\mathrm{h}}_F(E) &\\ge \\frac{1}{4\\sqrt5}\\log|D| + \\frac 12 \\log f\n-c(f)\n-\\gamma -\\frac{\\log(2\\pi)}{2},\\\\\n&=\\frac{1}{4\\sqrt5}\\log|\\Delta| + \\lambda \\log f\n-c(f)\n-\\gamma -\\frac{\\log(2\\pi)}{2}.\n\\end{align*}\nThus, to establish~\\eqref{efaltlowernewf}, we only have to prove that ${c(f)\\le \\beta(f)}$. \nWe have \n$$\n\\frac{1-\\chi(p)}{p-\\chi(p)} = \n\\begin{cases}\n0, &\\chi(p)=1,\\\\\n1\/p, &\\chi(p)=0,\\\\\n 2\/(p+1), &\\chi(p)=-1. \n\\end{cases}\n$$\nHence\n$$\n\\frac{1-\\chi(p)}{p-\\chi(p)} \\le \\frac2{p+1}\n$$\nin any case. This implies that ${c(f)\\le \\beta(f)}$.\nThe proposition is proved. \n\\end{proof}\n\n\n\\section{The estimate \\texorpdfstring{${|\\Delta|< 10^{15}}$}{|Delta|<10to15}}\n\\label{stenfourteen}\n\n\n\nIn this section we obtain the first explicit upper bound for the\ndiscriminant of a singular unit. \n\n\n\\begin{theorem}\n \\label{thhighrange}\nLet~$\\Delta$ be the discriminant of a singular unit. Then\n ${|\\Delta|<10^{15}}$. \n\\end{theorem}\n\n{\\sloppy\n\nThroughout this section~$\\Delta$ is the discriminant of a singular unit~$\\alpha$, and we assume that ${X=|\\Delta|\\ge 10^{15}}$, as otherwise there is nothing to prove. \nOur principal tools will be the upper estimate~\\eqref{eupdel}\nand the lower estimates~\\eqref{etrivbis},~\\eqref{ecolmbis}. We reproduce them here for convenience:\n\\begin{align}\n\\label{eupx}\n{\\mathrm{h}}(\\alpha)&\\le \\frac{12A}{{\\mathcal{C}}(\\Delta)}+3\\log\\frac{AX^{1\/2}}{{\\mathcal{C}}(\\Delta)}-3.77,\\\\\n\\label{etrivx}\n{\\mathrm{h}}(\\alpha)&\\ge \\frac{\\pi X^{1\/2}-0.01}{{\\mathcal{C}}(\\Delta)},\\\\\n\\label{ecolmx}\n{\\mathrm{h}}(\\alpha)&\\ge \\frac{3}{\\sqrt5}\\log X-9.79.\n\\end{align}\nNote that our assumption ${X\\ge 10^{15}}$ implies that the right-hand side of~\\eqref{ecolmx} is positive.\n\n}\n\n\n\\subsection{The main inequality}\n\\label{ssinequality}\n\nRecall that ${A= F\\log X}$. \nMinding~$0.01$ in~\\eqref{etrivx} we deduce from~\\eqref{eupx},~\\eqref{etrivx} and~\\eqref{ecolmx} the inequality\n$$\n\\frac{12A}{{\\mathcal{C}}(\\Delta)}+3\\log\\frac{AX^{1\/2}}{{\\mathcal{C}}(\\Delta)}-3.76 \\ge \\max\\left\\{\\frac{\\pi X^{1\/2}}{{\\mathcal{C}}(\\Delta)}, \\frac{3}{\\sqrt5}\\log X-9.78\\right\\}. \n$$\nDenoting\n\\begin{equation}\n\\label{ey}\nY=\\max\\left\\{\\frac{\\pi X^{1\/2}}{{\\mathcal{C}}(\\Delta)}, \\frac{3}{\\sqrt5}\\log X-9.78\\right\\}, \n\\end{equation}\nwe re-write this as \n\\begin{equation}\n\\label{enotyet}\n\\frac{12A\/{\\mathcal{C}}(\\Delta)}{Y}+\\frac{3\\log A-3.76}Y+ \\frac{\\log(X^{1\/2}\/{\\mathcal{C}}(\\Delta))}{Y} \\ge 1. \n\\end{equation}\nNote that ${3\\log A-3.76>0}$, because ${A\\ge \\log X\\ge \\log(10^{15}) >30}$. Hence we may replace~$Y$ by ${\\frac{3}{\\sqrt5}\\log X-9.78}$ in the middle term of the left-hand side in~\\eqref{enotyet}. Similarly, in the first term we may replace~$Y$ by ${\\pi X^{1\/2}\/{\\mathcal{C}}(\\Delta)}$, and in the third term we may replace ${X^{1\/2}\/{\\mathcal{C}}(\\Delta)}$ by ${\\pi^{-1}Y}$. We obtain \n\\begin{equation}\n\\label{einequality}\n12\\pi^{-1} AX^{-1\/2}+ \\frac{3\\log A-3.76}{\\frac{3}{\\sqrt5}\\log X-9.78} + 3\\frac{\\log(\\pi^{-1}Y)}{Y}\\ge 1. \n\\end{equation}\n\n\nTo show that~\\eqref{einequality} is not possible for ${X\\ge 10^{15}}$, we will bound from above each of the three terms in its left-hand side. To begin with, we bound~$A$. \n\n\n\\subsection{Bounding~$F$ and~$A$}\n\nRecall that ${F=\\max\\{2^{\\omega(a)}: a\\le X^{1\/2}\\}}$ and ${A=F\\log X}$. \n\nLet \n${N_1 = 2\\cdot3\\cdot5\\cdots 1129}$ be the product of the first $189$ prime numbers. Define the real number~$c_1$ from \n\\begin{align*}\n\\omega(N_1)&=\\frac{\\log N_1}{\\log\\log N_1 -c_1}.\n\\end{align*}\nA calculation shows that ${c_1 < 1.1713142}$. \nRobin \\cite[Th\u00e9or\u00e8me~13]{Ro83} proved that\n\\begin{equation*}\n\\omega(n) \\le \\frac{\\log n}{\\log\\log n-c_1} \n\\end{equation*}\nfor ${n\\ge 26}$. \nThis implies that \n\\begin{align}\n\\label{eboundf}\n\\frac{\\log F}{\\log 2} &\\le \\frac12\\frac{\\log X}{\\log\\log X-c_1-\\log2} ,\\\\\n\\label{ebounda}\n\\log A &\\le \\frac{\\log2}2\\frac{\\log X}{\\log\\log X-c_1-\\log2}+\\log\\log X. \n\\end{align}\nIndeed, the function \n$$\ng(x)= \\frac{\\log x}{\\log\\log x-c_1}\n$$\nis strictly increasing for ${x\\ge 6500}$ and ${g(6500)>8}$. \nIf ${a\\le X^{1\/2}}$ then either ${a\\le 6500}$ in which case \n${\\omega(a) \\le 5A>0$ we remark that\n\\begin{equation}\n\\label{eremark}\nB\\log B - A \\log A \\le (B-A)(1+\\log B).\n\\end{equation}\nWhen ${A\\ge 2}$ estimate~\\eqref{esqrtab} follows immediately\nfrom~\\eqref{esqrt} and~\\eqref{eremark}.\n\nLet us assume that ${A<2}$, hence $\\lfloor A\\rfloor$ is $0$ or $1$\nand $S(\\lfloor A\\rfloor)=0$ or $1$, respectively. For $B\\ge 2$ we find\n$$\n\\sum_{A 0}$, as ${X\\ge 10^{5}}$, \nto get\n\\begin{equation}\n \\label{eq:1lowerbound}\\begin{aligned}\n 1&\\le \\frac{3}{\\pi}(8\\varepsilon^2 + 0.811\\varepsilon) (\\log X)^2\n +\\frac{3}{\\pi}(28\\varepsilon^2 + 2.829\\varepsilon) \\log X\\\\\n &\\hphantom{\\le}+ \\frac{267}{\\pi}\\varepsilon\\frac{\\log X}{X^{1\/8}} \n +\\frac{93.18}{\\pi X^{1\/4}}\n +\n \\frac{3\\log(\\varepsilon^{-1})-10.65}{\\frac{3}{\\sqrt 5}\\log X - 9.78}.\n\\end{aligned}\n\\end{equation}\nOur choice is ${\\varepsilon = \n10^{-4}}$.\nThe first two terms in the right-hand side of~\\eqref{eq:1lowerbound} are \nmonotonously increasing, and the remaining three terms are decreasing for ${X\\in[10^{10}, 10^{15})}$; note that ${x\\mapsto(\\log x)\/x^{1\/8}}$ is\ndecreasing for ${x \\ge 3000 > e^8}$. Using ${X < 10^{15}}$ for the first two terms and ${X\\ge 2\\cdot 10^{10}}$ for the\n remaining three terms, we see that the right-hand side of~\\eqref{eq:1lowerbound} is strictly smaller than $0.962$ if\n${X \\in [2\\cdot10^{10}, 10^{15})}$. Similarly, we infer that it is strictly smaller than $0.960$ if ${X \\in [10^{10}, 2\\cdot10^{10})}$. This completes the proof of Theorem~\\ref{thmidrange}. \n\\qed\n\n}\n\n\\section{Handling the low-range \\texorpdfstring{$3\\cdot 10^{5}\\le |\\Delta|< 10^{10}$}{3.10to5<|Delta|<10to10}}\n\\label{slowrange}\nWe now deal with the low-range ${|\\Delta|\\in [3\\cdot10^5, 10^{10})}$. For this range the upper\nbound on ${\\mathcal{C}}_\\varepsilon(\\Delta)$ arises from a computer-assisted search algorithm.\n\n \n\nWe prove the following. \n\\begin{theorem}\n \\label{thlowrange}\n Let~$\\Delta$ be the discriminant of a singular unit. Then\n ${|\\Delta|\\notin [3\\cdot10^5, 10^{10})}$. \n\\end{theorem}\n\nThe proof relies on the following lemma. \n\n\\begin{lemma}\n \\label{lem:lowrangeCebound}\n Let $\\Delta$ be the discriminant of a singular modulus.\n \\begin{enumerate}[label={(\\roman*)}]\n \\item \n \\label{itenten} If ${10^7\\le|\\Delta|< 10^{10}}$ and ${\\varepsilon = 10^{-3}}$, then\n ${{\\mathcal{C}}_\\varepsilon(\\Delta)\\le 16}$.\n \\item \n \\label{itenseven}\n If ${3\\cdot10^5\\le|\\Delta|< 10^7}$ and ${\\varepsilon = 4\\cdot 10^{-3}}$, then\n ${{\\mathcal{C}}_\\varepsilon(\\Delta)\\le 6}$. \n \\end{enumerate}\n\\end{lemma}\n\\begin{proof}\nLet $X_{\\min}$ and $X_{\\max}$ be positive integers satisfying ${ X_{\\min} 0$}\n \n \n \\Output{an upper bound for ${\\mathcal{C}}_\\varepsilon(\\Delta)$ for all\n discriminants $\\Delta\\in [-X_{\\mathrm{max}},-X_{\\mathrm{min}}]$}\n \\BlankLine\n\n $counter \\leftarrow$ pointer to array of length\n $X_{\\mathrm{max}}-X_{\\mathrm{min}}+1$ initialized to $0$\\;\n\n $bound \\leftarrow 0$\\;\n \n \\For{$c\\leftarrow \\lfloor X_{\\mathrm{min}}^{1\/2}\/2\\rfloor$ \\KwTo\n $\\lfloor X_{\\mathrm{max}}^{1\/2}\\rfloor$}{\n \\For{$a\\leftarrow \\lfloor c\/(1+\\sqrt 3 \\varepsilon+\\varepsilon^2) \\rfloor$ to $c$}{\n \\For{$b\\leftarrow \\lfloor (1-2\\varepsilon)a\\rfloor$ to $a$}{\n $X\\leftarrow 4ac-b^2$\\;\n \\If{$X\\ge X_{\\mathrm{min}}$ and $X \\le X_{\\mathrm{max}}$\n }{$pos \\leftarrow X-X_{\\mathrm{min}}$\\;\n $counter[pos] \\leftarrow counter[pos]+2$\\;\n \\lIf{$counter[pos] > bound$}{$bound\\leftarrow counter[pos]$}\n }\n }\n }}\n \\Return{$bound$}\\;\n \\caption{Compute an upper bound for ${\\mathcal{C}}_{\\epsilon}(\\Delta)$ in the\n range ${\\Delta\\in [-X_{\\mathrm{max}},-X_{\\mathrm{min}}]}$}\\label{algo:CCDelta}\n }%\n \\end{algorithm}\n \n\n\n\\begin{proof}[Proof of Theorem~\\ref{thlowrange}]\n Assume that~$\\alpha$ is a singular unit of discriminant ${\\Delta\\in (-10^{-10}, -3\\cdot10^5]}$.\nLet ${0<\\varepsilon\\le 4\\cdot 10^{-3}}$. We set again~$Y$ as in (\\ref{ey}).\n As in the proof of Theorem~\\ref{thmidrange} we find (\\ref{eq:midrangeYbound}). \n \n \nWe infer that\n \\begin{alignat*}1\n 1 &\\le 3\\frac{{\\mathcal{C}}_\\varepsilon(\\Delta)}{{\\mathcal{C}}(\\Delta) Y} \\log X +\n \\frac{3\\log(\\varepsilon^{-1}) - 10.65}{Y}\\\\\n & \\le \\frac {3{\\mathcal{C}}_\\varepsilon(\\Delta)}{\\pi} \\frac{\\log X}{X^{1\/2}} +\n \\frac{3\\log(\\varepsilon^{-1}) - 10.65}{\\frac{3}{\\sqrt{5}}\\log X - 9.78}\n \\end{alignat*}\nwhere we use ${{\\mathcal{C}}(\\Delta)Y \\ge \\pi X^{1\/2}}$\n and ${Y\\ge \\frac{3}{\\sqrt{5}} \\log X - 9.78}$. \n\n If ${X\\in [10^7,10^{10})}$ then we set ${\\varepsilon = 10^{-3}}$ and use the estimate\n ${{\\mathcal{C}}_{\\varepsilon}(\\Delta) \\le 16}$ from Lemma~\\ref{lem:lowrangeCebound}\\ref{itenten}. \n Recall that ${x\\mapsto (\\log x)\/x^{1\/2}}$\n is decreasing for ${x\\ge e^2}$. So we find\n \\begin{equation*}\n 1 \\le \\frac{3\\cdot 16}{\\pi}\\frac{\\log(10^7)}{10^{7\/2}} +\n \\frac{3\\log(1000) - 10.65}{\\frac{3}{\\sqrt{5}} \\log(10^7)- 9.78}< 0.929,\n \\end{equation*}\n a contradiction.\n\nWhen ${X\\in [3\\cdot10^5,10^7)}$ we set ${\\varepsilon = 4\\cdot10^3}$. Then\n ${{\\mathcal{C}}_{\\varepsilon}(\\Delta)\\le 6}$ by Lem\\-ma~\\ref{lem:lowrangeCebound}\\ref{itenseven}.\nUsing \n ${X\\ge 3\\cdot 10^5}$ we find as before \n \\begin{equation*}\n 1 \\le \\frac{3\\cdot 6}{\\pi}\\frac{\\log(3\\cdot 10^5)}{(3\\cdot 10^{5})^{1\/2}} +\n \\frac{3\\log(250) - 10.65}{\\frac{3}{\\sqrt{5}} \\log(3\\cdot 10^{5})- 9.78}< 0.961,\n \\end{equation*}\n another contradiction which completes this proof. \n\\end{proof}\n\n\n\\section{The extra low-range}\n\n\\label{sfinal}\n\nThe results of the three previous sections reduce the proof of Theorem~\\ref{thmain} to the\nfollowing assertion.\n\n\n\\begin{theorem}\n\\label{thm:habeggerspari}\n Let~$\\Delta$ be the discriminant of a singular unit. Then \n ${|\\Delta|\\ge 3\\cdot 10^5}$. \n\\end{theorem}\n\\begin{proof}\n Let $\\alpha$ be a singular unit of discriminant $\\Delta$. We write\n ${X=|\\Delta|}$. \nWe may assume that ${X\\ge 4}$ because the only singular modulus of discriminant $-3$ is\n${j(\\zeta_3)=0}$, \nwhich is not an algebraic unit.\n\n\nRecall from Section~\\ref{sceps} that the Galois conjugates of~$\\alpha$ are precisely the singular moduli\n$j(\\tau)$, where ${\\tau = \\tau(a,b,c)}$ with $(a,b,c)$ as in~\\eqref{ekuh}. The imaginary part of such~$\\tau$ is\n$X^{1\/2}\/(2a)$ and ${a\\le (X\/3)^{1\/2}}$\nby Lemma~\\ref{lgauss}\\ref{iadelta}. \n Lemma~\\ref{lttsn} implies that\n$$\n|j(\\tau)|\\ge e^{2\\pi X^{1\/2}\/(2a)} -2079\n =e^{\\pi X^{1\/2}\/a}-2079 > 23^{X^{1\/2}\/a}-2079\n$$\n as ${e^\\pi > 23}$. Using Lemmas~\\ref{lanal} and~\\ref{lliouv}, we find that\n$$\n|j(\\tau)| \\ge 42700 \\min \\left\\{ \\frac{\\sqrt3}{4X},4\\cdot 10^{-3}\\right\\}^3.\n$$\nThese\n bounds together show that\n \\begin{equation}\n\\label{eq:finaljlb}\n |j(\\tau)|\\ge \\max\\left\\{ 23^{\\lfloor X^{1\/2}\/a\\rfloor}-2079,\n 42700\\min\\Bigl\\{\\frac2{5X},\\frac1{250}\\Bigr\\}^3\\right\\}. \n \\end{equation}\nBased on this observation, Algorithm~\\ref{algo:exclude} prints a list of discriminants of potential singular\nunits in the range ${[-X_{\\max},-4]}$. For this purpose, it computes a \nrational lower bound~$P$ for the absolute value of the\n$\\Q(\\alpha)\/\\Q$-norm of each singular moduli in this range. Those singular\nmoduli where ${P \\le 1}$ are then flagged as potential singular units.\n\n We have implemented this algorithm as a \\textsf{PARI}\n script\\footnote{A link to our \\textsf{PARI} script \\textsf{algorithm2.gp}\nis on the second-named author's homepage. The running time is about 23 minutes on a regular desktop computer (Intel Xeon CPU E5-1620 v3, 3.50GHz, 32GB RAM). The only floating point operation used approximates $X^{1\/2}$\n which leads to ${n=\\lfloor X^{1\/2}\/a\\rfloor}$.\nTo rule out a rounding error in the floating point arithmetic we\ncompare $(an)^2$ with~$X$ in our implementation.}. \nThe script flags only $-4$, $-7$ and $-8$ as discriminants of potential singular units. The\nsingular moduli of these discriminants are well-known \\cite[(12.20)]{Cox}: they are $12^3$, $-15^3$ and $20^3$, respectively. None of them is a unit, which concludes the proof. \n\\end{proof} \n \n \\begin{algorithm}[t]{\\footnotesize\n \\SetKwInOut{Input}{Input}\\SetKwInOut{Output}{Output}\n \\Input{An integer $X_{\\mathrm{max}}\\ge 1$}\n \\Output{Print a list containing all discriminants\n in $[-X_{\\mathrm{max}},-4]$ that are attached to a potential\n singular unit.}\n\n \\For{$X\\leftarrow 4$ \\KwTo $X_{\\mathrm{max}}$}{\n $\\Delta \\leftarrow -X$\\;\n \n \\lIf{$\\Delta\\equiv 2 \\text{ or }3 \\mod 4$}{\n next $X$\n }\n \n\n\n$P\\leftarrow 1$\\;\n \\For{$a\\leftarrow 1$ \\KwTo $\\lfloor \\sqrt{X\/3}\\rfloor$} {\n\n $n\\leftarrow \\lfloor X^{1\/2}\/a\\rfloor$\\;\n \\For{$b\\leftarrow -a+1$ \\KwTo $a$}{\n \\lIf{$b^2 \\not\\equiv \\Delta \\mod 4a$}{\n next $b$\n }\n $c\\leftarrow (b^2-\\Delta)\/(4a)$\\;\n \\lIf{$a>c$}{next $b$}\n \\lIf{$a=c \\text{ and } b<0$}{next $b$}\n \\lIf{$\\mathrm{gcd}(a,b,c)\\not=1$}{next $b$}\n \n \n \n \n $P\\leftarrow P \\cdot \\max\\{23^{n\n }-2079,42700\\min\\{2\/(5X),1\/250\\}^3\\}$\\;\n}}\n\n\\lIf{$P\\le 1$}{print $\\Delta$}}\n\\caption{Exclude singular units} \\label{algo:exclude}\n}%\n \\end{algorithm}\n \n\n \n\n\nAs indicated in the introduction, Theorem~\\ref{thmain} is the combination of Theorems~\\ref{thhighrange},~\\ref{thmidrange},~\\ref{thlowrange} and~\\ref{thm:habeggerspari}. \n\n\n{\\footnotesize\n\n\\def$'${$'$}\n\\providecommand{\\bysame}{\\leavevmode\\hbox to3em{\\hrulefill}\\thinspace}\n\\providecommand{\\MR}{\\relax\\ifhmode\\unskip\\space\\fi MR }\n\\providecommand{\\MRhref}[2]{%\n \\href{http:\/\/www.ams.org\/mathscinet-getitem?mr=#1}{#2}\n}\n\\providecommand{\\href}[2]{#2}\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\n\\ \\ \\ \\ The time evolution of coherent states (CS) has attracted a great\ndeal of attention since the introduction of Glauber's CS of the harmonic\noscillator {\\normalsize \\cite{Glauber63}}. Of\nparticular interest has been the determination of the Hamiltonian operator\nfor which an initial coherent state remains coherent under time evolution.\nIt is established that this Hamiltonian has the form of the nonstationary\nbosonic forced oscillator Hamiltonian {\\normalsize \\cite{Glauber66, Mehta66,\n Stoler75, Kano76}: \\ \\ \\ \\ \\ \\ \\ \\ \\ \\ } \\\n\\begin{equation}\nH_{\\mathrm{cs}}=\\omega (t)a^{\\dagger }a+f(t)a^{\\dagger }+f^{\\ast }(t)a+\\beta\n(t), \\label{Hcs}\n\\end{equation}%\nwhere $\\omega (t)$ and $\\beta (t)$ are arbitrary real functions of time $t$,\nand $f(t)$ is arbitrary complex function. \n$H_{\\mathrm{cs}}$ is a particular case of the nonstationary forced oscillator\nHamiltonian for which the exact time evolution of CS has been obtained in\n\\cite{MMT70,Holz70} by first constructing boson ladder operator dynamical %\ninvariants according to the Lewis and Riesenfeld scheme of time\ndependent invariants \\cite{Lewis69}.\n\nOur purpose in the present article is to study the dynamical invariants and\ntime evolution of CS for the {\\it fermionic} forced oscillator (FFO), which %\nin fact is the general (one mode) Hamiltonian.\n\nThe organization of the article is as follows. In Sec. 2 we construct fermionic %\nladder operator dynamical invariants and the corresponding Lewis-Riesenfeld\nHermitian invariant {\\normalsize \\cite{Lewis69}}, following the scheme related %\nto the boson system {\\normalsize \\cite{MMT70}}. Using these invariants, we\nconstruct in Sec. 3 fermionic CS and Fock states of FFO system as eigenstates %\nof the constructed invariant fermionic annihilation operator $B(t)$ and %\n$B^\\dagger(t)B(t)$ correspondingly. These CS can\nrepresent (under appropriate initial conditions) the exact time-evolution of\ninitial canonical fermionic CS. Finally the relation of the invariant ladder %\noperators method {\\normalsize \\cite{MMT70, Holz70}} to the Lewis-Riesenfeld method\n{\\normalsize \\cite{Lewis69}} is briefly described on the example of FFO. The\npaper ends with concluding remarks.\\\n\n\\medskip\n\n\\section{FFO and invariant ladder operators}\n\nWe consider the single nonstationary fermionic forced oscillator (FFO)\ndescribed by the following Hamiltonian,%\n\\begin{equation}\nH_{\\!f}=\\omega (t)b^{\\dagger }b+f(t)b^{\\dagger }+f^{\\ast }(t)b+g(t),\n\\label{Hf}\n\\end{equation}%\nwhere $\\omega (t)$ and $g(t)$ are arbitrary real functions of time, $f(t)$\nis arbitrary complex function, $b$ and $b^{\\dagger }$ are fermion\nannihilation and creation operators respectively, which obey to the fermion\nalgebra:\n\\begin{equation}\n\\{b,b^{\\dagger }\\} = 1,\\text{ \\ }%\nb^{2}=b^{\\dagger }{}^{2}=0, \\label{013}\n\\end{equation}%\nwhere $\\{b,b^{\\dagger }\\}\\equiv bb^{\\dagger }+b^{\\dagger }b$.\nDue to the nilpotency of the fermionic operators $b$, $b^\\dagger$\nthe operator $H_{\\!f}$ represents the most general (one mode) fermionic %\nHamiltonian.\n\nThe Hilbert space $\\mathcal{H}$ of the single-fermion system is spanned by\nthe two eigenstates $\\left\\{ \\left\\vert 0\\right\\rangle ,\\left\\vert\n1\\right\\rangle \\right\\} $ of number operator $b^{\\dagger }b$:\\, %\n$b^{\\dagger }b\\left\\vert n\\right\\rangle =n\\left\\vert n\\right\\rangle ,\\text{ \\ }n=0,1$.\nThe operators $b$ and $b^{\\dagger }$\\ allow transitions between number states,\n\\begin{equation}\nb\\left\\vert 0\\right\\rangle =0,\\text{\\ }b\\left\\vert 1\\right\\rangle\n=\\left\\vert 0\\right\\rangle \\,,\\text{ \\ }b^{\\dagger }\\left\\vert\n1\\right\\rangle =0,\\text{ }b^{\\dagger }|0\\rangle =|1\\rangle .\\,\n\\end{equation}%\nThe form of the Hamiltonian ({\\normalsize \\ref{Hf}}) is a Hermitian linear\ncombination of $b$, $b^{\\dagger }$ and $N = b^{\\dagger }b$. The fermion number\noperator $N$ obey the relation $N^{2}=N$ and the three operators $b$, $%\nb^{\\dagger }$and $N$ close under commutation the algebra:\n\\begin{equation}\n\\left[ b,N\\right] =b,\\text{ }\\left[ b^{\\dagger },N\\right] =-b^{\\dagger },%\n\\text{ }\\left[ b,b^{\\dagger }\\right] =1-2N,\\text{ \\ \\ \\ }\n\\end{equation}%\nLet us note that linear combinations of $b^{\\dagger }$, $b$ and $N$ produce\nthe half-spin operators $J_{i}$,\n\\begin{equation}\nJ_{1}=\\tfrac{1}{2}(b^{\\dagger }+b),\\quad J_{2}=\\tfrac{1}{2\\ri}(b^{\\dagger\n}-b),\\quad J_{3}=b^{\\dagger }b-\\tfrac{1}{2}, \\label{30a}\n\\end{equation}%\nclosing the \\textit{su}(2) algebra: $\\left[ J_{k},J_{l}\\right] =i\\epsilon\n_{klm}J_{m}.$\n\nIt is convenient to use raising and lowering operators $J_{\\pm }=J_{1}\\pm\niJ_{2}$ which satisfy the following commutation relation: $\\left[ J_{+},J_{-}%\n\\right] =2J_{3},\\text{ }\\left[ J_{3},J_{\\pm }\\right] =\\pm J_{\\pm }\\text{\\ }$%\n, where $J_{+}=b^{\\dagger },$ $J_{-}=b.$ So that in terms of these half spin\noperators the Hamiltonian ({\\normalsize \\ref{Hf}}) takes the form\n\\begin{equation}\nH_{\\!f}=\\omega (t)J_{3}+f(t)J_{+}+f^{\\ast }(t)J_{-}+g(t)+\\tfrac{\\omega (t)}{2%\n}. \\label{Hf 2}\n\\end{equation}%\n\nOur task is the construction of the time-dependent invariants for the system\n{\\normalsize (\\ref{Hf})}, {\\normalsize (\\ref{Hf 2})}. The defining equation\nof the invariant operator $B(t)$ for a quantum system with Hamiltonian $H(t)$\nis%\n\\begin{equation}\n\\frac{\\partial }{\\partial t}B(t)-i[B(t),H]=0 \\label{19}\n\\end{equation}\nFormal solutions to Eq. {\\normalsize (\\ref{19})} are operators $%\nB(t)=U(t)B(0)U^{\\dagger }(t)$, where $U(t)$ is the evolution operator of the\nsystem, $U =T\\exp [-\\ri\\tint_{0}^{t}H(t^{^{\\prime }}){\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e} t^{^{\\prime }}]$. \nIn our case of FFO {\\normalsize (\\ref{Hf})}, {\\normalsize (\\ref{Hf 2})} we look\nfor the non-Hermitian invariants $B(t)$, $B^\\dagger(t)$ of the form of linear %\ncombination of the $SU(2)$ generators {\\normalsize (\\ref{30a})}, \\\n\\begin{equation}\n\\begin{aligned} B = \\nu_{-}(t)J_{-} + \\nu_{+}(t) J_{+} + \\nu_3 (t) J_{3}\n,\\\\ B^\\dagger = \\nu_{-}^* (t)J_{+} + \\nu_{+}^*(t) J_{-} + \\nu_3^* (t) J_{3},\n\\end{aligned} \\label{B}\n\\end{equation}%\nwhere $\\nu _{\\pm }(t),\\,\\,\\nu _{3}(t)$ may be complex functions of the time.\nHermitian invariants then can be easily built up as Hermitian combinations\nof $B$ and $B^{\\dagger }$. In particular if $B$ is a non-Hermitian invariant\nthe operator $I=B^{\\dagger }B- 1\/2$ \nis a Hermitian invariant, the fermion analog of the Lewis-Riesenfeld quadratic\ninvariant {\\normalsize \\cite{Lewis69}}.\n\nLet us note at this point that we look for FFO invariants as elements of the\nsame algebra $su(2)$ to which the Hamiltonian belongs. Similar is the approach\nused in \\cite{Chumakov86} in construction of invariants for\nthe nonstationary singular oscillator, where the related algebra is $su(1,1)$.\nThis is to be compared with the case of nonsingular oscillator, for which invariant\nladder operators have been built up as elements of the Heisenberg-Weyl algebra %\n$h_w$ (i.e. as linear combinations of coordinate and momentum operators $x$ and %\n$p$ \\cite{MMT70,Holz70}), while the related nonstationary Hamiltonian belongs to \n$su(1,1)$. %\nFor {\\it forced} boson oscillator the Hamiltonian belongs to the large algebra of \nsemi-direct sum $su(1,1)\\dot{+}h_w$ but the ladder operator invariants %\n are again elements of the invariant subalgebra $h_w$ %\n\\cite{Prants86,Dattoli86,Trif93}.\n\n\nTo proceed with construction of FFO invariants we substitute {\\normalsize (\\ref{B}) and \n(\\ref{Hf 2}) } into {\\normalsize (\\ref{19})}, and find the following system of\ndifferential equations for the parameter functions $\\nu_\\pm,\\,\\nu_3$: \\\n\\begin{eqnarray}\n\\dot{\\nu}_{3} &=&2\\ri(\\nu _{+}f^{\\ast }-\\nu _{-}f), \\label{(a)} \\\\\n\\dot{\\nu}_{+} &=&\\ri(\\nu _{3}f-\\nu _{+}\\omega ), \\label{(b)} \\\\\n\\dot{\\nu}_{-} &=&\\ri(\\nu _{-}\\omega -\\nu _{3}f^{\\ast }). \\label{(c)}\n\\end{eqnarray}%\nSolutions to the above linear system of first order equations are\nuniquely determined by the initial conditions $\\nu _{\\pm }(0)=\\nu _{0,\\pm }$%\n, $\\nu _{3}(0)=\\nu _{0,3}$. If we want the invariants $B(t)$ and $B^{\\dagger\n}(t)$ be again fermion ladder operators, i.e. to obey the conditions\n\\begin{equation}\nB^{2}=0, \\quad \\{B,B^{\\dagger}=1, \\label{constr}\n\\end{equation}%\nwe have to take $\\nu _{0,\\pm }$ and $\\nu _{0,3}$ satisfying\n\\begin{equation}\n \\nu _{0,3}^{2} = -4\\nu _{0,+}\\nu _{0,-},\\quad\n |\\nu _{0,-}|+|\\nu _{0,+}| = 1\\,. \\label{0}\n\\end{equation}%\nIndeed, for $B^{2}(t)$ and $\\{B(t),B^{\\dagger }(t)\\}$ we find\n\\begin{equation}\n\\begin{tabular}{l}\n$\\displaystyle B^{2}=\\nu _{+}\\nu _{-}+\\tfrac{1}{4}\\nu _{3}^{2}\\equiv\n\\lambda _{1}$, \\\\[2mm]\n$\\displaystyle\\{B,B^{\\dagger}\\}=|\\nu _{-}|^{2}+|\\nu _{+}|^{2}+\\tfrac{1%\n}{2}|\\nu _{3}|^{2}\\equiv \\lambda _{2}$.%\n\\end{tabular}\n\\label{lam_i}\n\\end{equation}%\nThe quantities $\\lambda _{1}(\\nu _{\\pm },\\nu _{3})$, $\\lambda _{2}(\\nu _{\\pm\n},\\nu _{3})$ turned out to be two different '\\textit{constants of motion}'\nfor the system {\\normalsize (\\ref{(a)})}-{\\normalsize (\\ref{(c)})}, their\ntime derivatives being vanishing:\n\\begin{equation}\n\\begin{tabular}{l}\n$\\displaystyle \n\\frac{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e}}{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e} t}\\lambda _{1} = \\frac{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e}}{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e} t}\\left( \\nu _{+}\\nu _{-}+\\tfrac{1%\n}{4}\\nu _{3}^{2}\\right) =0$, \\\\[3mm]\n$\\displaystyle\n\\frac{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e}}{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e} t}\\lambda _{2} = \\frac{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e}}{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e} t}\\left( |\\nu _{-}|^{2}+|\\nu\n_{+}|^{2}+\\tfrac{1}{2}|\\nu _{3}|^{2}\\right) =0$.%\n\\end{tabular}\n\\label{dotlam_i}\n\\end{equation}%\nTherefore we can fix the values of these constants as $\\lambda _{1}=0$, $%\n\\lambda _{2}=1$, i.e.\n\\begin{equation}\n\\begin{tabular}{l}\n$ \\displaystyle \\nu _{+}\\nu _{-}+\\tfrac{1}{4}\\nu _{3}^{2}=0,$\\\\[3mm]\n$\\displaystyle |\\nu _{-}|^{2}+|\\nu_{+}|^{2}+\\tfrac{1}{2}|\\nu _{3}|^{2}=1$, \n\\end{tabular}\n\\label{lam 0}\n\\end{equation}%\nand satisfy the conditions {\\normalsize (\\ref{constr})}. If furthermore the initial\nconditions are taken as\n\\begin{equation}\n\\nu _{-}(0)=1,\\,\\,\\,\\nu _{+}(0)=0=\\nu _{3}(0), \\label{nu_i0}\n\\end{equation}%\nthen $B(0)=b$. Later on we work with these fermionic ladder operator invariants, %\ni.e. we consider conditions (\\ref{lam 0}) satisfied.\n\nLet us now recall that in the case boson nonstationary oscillator the ladder %\noperator invariants, constructed first in \\cite{MMT70, Holz70} (see also %\n\\cite{Chumakov86, Prants86}), are expressed in terms of one only parameter function %\n$\\epsilon(t)$, which obeys a simple second order equation, namely that of the %\nclassical oscillator with varying frequency. It turned out that this can be done %\nin the case of fermionic oscillator as well.\nIn this aim we first express all the three parameter functions $\\nu _{\\pm }(t)$,%\n $\\nu _{3}(t)$ in terms of one of them, which has to obey a second order %\ndifferential equation.\nLet for example, express $\\nu _{3}(t)$ and $\\nu _{-}(t)$ in terms of $\\nu _{+}(t)$%\n and its derivatives. We have %\n\\begin{equation} \\label{nu3}\n\\nu _{3} = - \\tfrac{\\ri}{f}(\\dot{\\nu}_{+} + \\ri \\nu _{+}\\omega ),\n\\end{equation}%\n\\begin{equation}\n\\nu _{-}=\\tfrac{1}{2f^{2}}\\left[ \\ddot{\\nu}_{+}+\\left( \\ri \\omega -\\tfrac{\\dot{f%\n}}{f}\\right) \\dot{\\nu}_{+}+\\left( 2ff^{\\ast }+\\ri\\dot{\\omega}-\\ri \\frac{\\omega }{f%\n}\\dot{f}\\right) \\nu _{+}\\right] . \\label{nu-}\n\\end{equation}%\nSubstituting these expressions into the expression of $\\lambda_1$ in terms of $\\nu_\\pm,\n\\,\\nu_3$ and taking into account that $\\lambda_1$ is fixed to $0$ we find that $\\nu_+$ %\nshould satisfy the following second order equation,\n\n\\begin{equation}\n2\\nu _{+}\\ddot{\\nu}_{+}-\\dot{\\nu}_{+}^{2} - 2\\nu_{+}\\dot{\\nu}_{+}\\tfrac{\\dot{f}}{f} +\n4\\nu _{+}^{2}\\left(\n|f|^{2}+\\tfrac{\\omega ^{2}}{4}+ \\ri\\tfrac{\\dot{\\omega}}{2}-\\ri \\tfrac{\\omega %\n\\dot{f }}{2f}\\right) = 0 . \\label{lam}\n\\end{equation}%\n Using this, and supposing that $\\nu_{+}\\neq 0$, we obtain for $\\nu _{-}$ a more %\ncompact expression in terms of $\\nu _{+}$ and $\\dot{\\nu}_{+}$, %\n\\begin{equation}\n\\nu _{-}= -\\nu _{3}^{2}\/4\\nu _{+}, \\label{nu- 2}\n\\end{equation}%\nwhere $\\nu_3$ is given again by eq. (\\ref{nu3}).\n\nThus the operators $B(t),\\,B^{\\dagger}(t)$, eq. (\\ref{B}), are fermionic %\nladder operator invariants for the forced oscillator (\\ref{Hf}), (\\ref{Hf 2})%\nif $\\nu _{3}$ and $\\nu _{-}$ are given by eqs. (\\ref{nu3}) and (\\ref{nu- 2}),%\nand $\\nu _{+}(t)$ is a nonvanishing solution of the second order equation (\\ref{lam}).\n\nNext we try to linearize the auxiliary eq. (\\ref{lam}). In this purpose we put\n\\begin{equation}\n\\nu _{+}(t) = \\tfrac{1}{2}{\\epsilon'} ^{2}(t) \\label{nu+ eps'}\n\\end{equation}%\nand obtain that $\\epsilon' (t)$ satisfies the linear equation\n\\begin{equation} \\label{eps' eq}\n\\ddot{\\epsilon}' - \\tfrac{\\dot{f}}{f}\\dot{\\epsilon}'+\\Omega' (t)\\epsilon' =0,\n\\end{equation}%\nwhere\n\\begin{equation}\n\\Omega' (t)=|f(t)|^{2}+\\tfrac{1}{4}\\omega ^{2}(t)+\\tfrac{\\ri}{2}\\dot{\\omega}-%\n\\tfrac{\\ri}{2}\\omega \\tfrac{\\dot{f}}{f}.\n\\end{equation}\n\nIn terms of $\\epsilon'$ the formulas (\\ref{nu-}) and (\\ref{nu3} for $\\nu_-,\\, \\nu_3$ read\n\\begin{equation}\n\\begin{tabular}{l}\n$\\displaystyle\\nu _{-}=-\\frac{1}{2\\epsilon'}\\nu_3^{2}$, \\\\[2mm]\n$\\displaystyle\\nu _{3}=\\frac{1}{f}\\left( \\tfrac{\\omega }{2}{\\epsilon'}\n^{2}-\\ri\\epsilon' \\dot{\\epsilon}'\\right) $.%\n\\end{tabular}%\n\\end{equation}%\nThe term in (\\ref{eps' eq}) proportional to the first derivative can be eliminated %\nby the substitution %\n\\begin{equation}\\label{eps}\n\\epsilon' = \\epsilon \\exp \\left( \\tfrac{1}{2}\\tint_{0}^{t}{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e}\\tau \\,\\dot{f}%\n(\\tau )\/f(\\tau )\\right) .\n\\end{equation}%\nThis leads to the desired simple equation for $\\epsilon$,\n\\begin{equation}\\label{eps eq}\n\\ddot{\\epsilon}+\\Omega (t)\\epsilon = 0,\n\\end{equation}%\nwhere $ \\Omega(t)= \\Omega^{\\prime }(t) + \\ddot{f}\/2f - 3\\dot{f}\\,^{2}\/4f^{2}$. %\nEquation (\\ref{eps eq}) is of the same type, as the auxiliary equation used in the %\ncase of nonstationary boson oscillator \\cite{MMT70, Holz70}. Here the %\n'squared frequency' $\\Omega$ however is complex and depends in a different manner %\non the corresponding Hamiltonian parameters. And the solutions are subject to %\ndifferent constraints, stemming from the different commutation relations: \nin terms of our $\\epsilon$, eq. (\\ref{eps eq}), the constraint $\\lambda_2=1$ reads %\n( $\\epsilon'$ is related to $\\epsilon$ according to (\\ref{eps})),\n\\begin{equation}\n\\tfrac{|\\epsilon' |^{4}}{4}\\left( 1+\\tfrac{2}{|f|^{2}}\\left\\vert\n\\tfrac{\\omega }{2}\\epsilon' -\\ri\\dot{\\epsilon}'\\right\\vert ^{2}+\\tfrac{1}{|f|^{4}%\n}\\left\\vert \\tfrac{\\omega }{2}\\epsilon' -\\ri\\dot{\\epsilon}'\\right\\vert^{4}\\right) =1 ,\n\\end{equation}%\n while in the boson case the constraint is ${\\rm Im}\\left({\\epsilon}^*\\dot{\\epsilon}\\right) = 1$ \n\\cite{MMT70, Holz70}.\n\nTo finalize this section let note that in the particular case of the %\n\\textit{free} fermion oscillator, $f(t)\\equiv 0$, the explicit solutions of the problem can %\n be easily found in the form %\n\\begin{equation}\n\\begin{tabular}{l}\n$\\displaystyle \n\\nu _{\\pm }(t)=\\nu _{0,\\pm }\\re^{\\pm \\ri\\tint^{t}\\omega (\\tau ){\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e}\\tau },$\\\\\n$ \\displaystyle \n\\nu_{3}=\\nu _{0,3},$\n\\end{tabular}\n\\label{solutions1}\n\\end{equation}%\nwhere $\\nu _{0,\\pm }$, $\\nu _{0,3}$ are constants. To ensure the fermionic %\ncommutation relations of $B(t),\\,B^\\dagger(t)$ they have to obey the relations\n$\\nu _{0,-}\\,\\nu _{0,+} + \\nu _{0,3}^{2}\/4 =0$ \\,\\, and \\,\\, $ |\\nu\n_{0,-}|^{2} + |\\nu _{0,+}|^{2} + |\\nu _{0,3}|^{2}\/2 = 1.$\n\\medskip\n\n\\section{CS for the fermion forced oscillator}\n\nWe define coherent states (CS) for a given fermion system as eigenstates of\nthe corresponding invariant fermion annihilation (or creation) operator $B(t)\n$. Since the most general fermion one mode Hamiltonian operator is of the\nform of (nonstationary) forced oscillator (\\ref{Hf 2}), the one-mode fermion\nCS are defined as eigenstates of the invariant ladder operator $B(t)$\n(eqs. (\\ref{B}), (\\ref{constr})):\n\\begin{equation}\nB(t)|\\zeta ;t\\rangle =\\zeta |\\zeta ;t\\rangle . \\label{|z;t> 1}\n\\end{equation}%\nSince $B(t)$ is invariant operator, the eigenvalue $\\zeta $ does\nnot depend on time $t$. In terms of the $\\zeta $, $B(t)$, $B^{\\dagger }(t)$\nand the $B(t)$-vacuum $|0;t\\rangle $ we have for $|\\zeta ;t\\rangle $ the\nsame formulas as for the canonical fermion CS $|\\zeta \\rangle $ which are\ndefined {\\normalsize \\cite{Klauder, Abe89, Maam92, Junker98, Cahill99}} as\n\\begin{equation}\n\\left\\vert \\zeta \\right\\rangle =e^{-\\frac{1}{2}\\zeta ^{\\ast }\\zeta }\\left(\n\\left\\vert 0\\right\\rangle -\\text{ }\\zeta \\left\\vert 1\\right\\rangle \\right)\n\\,. \\label{|z>}\n\\end{equation}%\nwhere the eigenvalue $\\zeta $ is a Grassmannian variable: $\\zeta ^{2}=0,\\\n\\zeta \\zeta ^{\\ast }+\\zeta ^{\\ast }\\zeta =0$, $\\left\\vert 0\\right\\rangle $\nis the fermionic vacuum, $b\\left\\vert 0\\right\\rangle =0$, and $\\left\\vert\n1\\right\\rangle $ is the one-fermion state, $\\left\\vert 1\\right\\rangle\n=b^{\\dagger }\\left\\vert 0\\right\\rangle $. In particular\n\\begin{equation}\n|\\zeta ;t\\rangle = \\re^{-\\tfrac{1}{2}\\zeta ^{\\ast }\\zeta }\\left( |0;t\\rangle\n-\\zeta B^{\\dagger }(t)|0;t\\rangle \\right) . \\label{|z;t> 2}\n\\end{equation}%\nIt remains therefore to construct the (normalized) new ground state $%\n|0;t\\rangle $ according to its defining equations\n\\begin{equation}\n\\begin{aligned} B(t) |0;t\\rangle = 0 ,\\quad \n\\ri\\frac{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e}}{dt} |0;t\\rangle = H_{f}|0;t\\rangle. \\end{aligned} \\label{|0;t> 1}\n\\end{equation}%\nWe put\n\\begin{equation}\n|0;t\\rangle =\\alpha _{0}(t)|0\\rangle +\\alpha _{1}(t)|1\\rangle ,\n\\label{|0;t> 2}\n\\end{equation}%\nsubstitute this into (\\ref{|0;t> 1}) and after some tedious calculations find%\n\\begin{eqnarray}\n\\alpha _{1}(t) &=&\\alpha _{0}(t)\\frac{\\nu _{3}^{\\ast }(t)}{2\\nu _{+}^{\\ast\n}(t)}, \\\\\n\\alpha _{0}(t) &=&\\sqrt{|\\nu _{+}(t)|}\\exp \\left[ -\\tfrac{\\ri}{2}\\left(\n\\varphi _{\\nu _{+}}(t)+\\tint_{0}^{t}(2g(\\tau )+\\omega (\\tau )){\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e}\\tau \\right) %\n\\right] ,\n\\end{eqnarray}%\nwhere $\\varphi _{\\nu _{+}}$ is the phase of $\\nu _{+}(t)$. The state $|\\zeta\n;t\\rangle $ will represent the exact time evolution of an initial canonical\nCS $|\\zeta \\rangle $ if the initial conditions (\\ref{nu_i0}) are imposed: %\n$|\\zeta ;0\\rangle = |\\zeta \\rangle $. In this case, the time evolved state %\n$ |\\zeta ;t\\rangle $ could be again an eigenstate of $b$ if the oscillator %\nis not 'forced', i.e. if $f(t)=0$.\nLet us note that the time-dependence of the constructed states is obtained %\nin terms of solutions to the system of auxiliary equations (\\ref{(a)})-(\\ref{(c)}),%\nor equivalently to the 'classical oscillator' equation (\\ref{eps eq}).\n\nOur method of construction of dynamical invariants differs slightly from the\nLewis-Riesenfeld method \\cite{Lewis69} (developed for bosonic oscillators).\nLewis and Riesenfeld used to first construct Hermitian invariant, which then\nis represented as a product of normally ordered ladder operators. To make\nconnection to their approach let us suppose that we first succeeded to\nconstruct the Hermitian invariant $N(t)$ and to find some ladder operators $%\n\\tilde{B}(t)$, $\\tilde{B}^{\\dagger }(t)$ that factorize it: $N(t)=\\tilde{B}%\n^{\\dagger }(t)\\tilde{B}(t)$. It is clear that $\\tilde{B}(t)$ may differ from\nour non-Hermitian invariant $B(t)$ in a phase factor:\\thinspace\\ $\\tilde{B}%\n(t) = {\\rm e}^{\\ri\\varphi (t)}B(t)$.\\newline\nWe can then in a standard way construct normalized eigenstates of $N(t)$,\n\\begin{equation}\nN(t)\\widetilde{|0;t\\rangle }=0,\\quad N(t)\\widetilde{%\n|1;t\\rangle }=\\widetilde{|1;t\\rangle }, \\label{tld 2}\n\\end{equation}%\nand of $\\tilde{B}(t)$, \\\n\\begin{equation}\n\\tilde{B}(t)\\widetilde{|\\zeta ;t\\rangle }=\\zeta \\widetilde{|\\zeta ;t\\rangle },\n\\end{equation}%\n\\begin{equation}\n\\widetilde{|\\zeta ;t\\rangle }=\\left( 1-\\tfrac{1}{2}\\zeta ^{\\ast }\\zeta\n\\right) \\left[ \\widetilde{|0;t\\rangle }-\\zeta \\widetilde{|1;t\\rangle }\\right]\n\\label{tld 3}\n\\end{equation}%\nwhich however do not obey the Schr\\\"odinger equation since, in general $ \\tilde{B}(t)$\nmay not be invariant. To obtain solutions $|n;t\\rangle $ and $ |\\zeta ;t\\rangle $ the\nabove eigenstates $\\widetilde{|n;t\\rangle }$, $n=0,1$, should also be multiplied by\nphase factors,\n\\begin{equation}\n|n;t\\rangle = {\\rm e}^{\\ri\\phi _{n}(t)}\\widetilde{|n;t\\rangle },\\quad n=0,1,\n\\label{tld 4}\n\\end{equation}%\n\\begin{equation}\n|\\zeta ;t\\rangle =\\left( 1-\\tfrac{1}{2}\\zeta ^{\\ast }\\zeta \\right) \\left[\n{\\rm e}^{\\ri\\phi _{0}(t)}\\widetilde{|0;t\\rangle }-\\zeta {\\rm e}^{\\ri\\phi _{1}(t)}%\n\\widetilde{|1;t\\rangle }\\right] \\label{tld 5}\n\\end{equation}%\nwhich should obey the equations \\\n\\begin{equation}\n\\tfrac{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e}}{{\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e} t}\\phi _{n}=\\widetilde{\\langle n;t|}\\ri\\tfrac{\\partial }{\\partial t}%\n-H\\widetilde{|n;t\\rangle }. \\label{tld 6}\n\\end{equation}%\nEvidently the state (\\ref{tld 5}) is an eigenstate of $\\tilde{B}(t)$ with time\ndependent eigenvalue $\\zeta (t)=\\zeta \\exp (\\ri\\varphi (t))$, $\\varphi\n(t)=\\phi _{1}(t)-\\phi _{0}(t)$.\\newline\nThe phase $\\varphi (t)=\\phi _{1}(t)-\\phi _{0}(t)$ consists of two parts -\ngeometrical one $\\varphi ^{G}$, and dynamical one $\\varphi ^{D}=\\varphi\n-\\varphi ^{G}$ \\cite{Maam99},\n\\begin{eqnarray}\n\\varphi ^{G}(t) &=&\\varphi (t)+\\int_{0}^{t}\\left( \\widetilde{\\langle 1;t^{\\prime }|}H%\n\\widetilde{|1;t^{\\prime }\\rangle }-\\widetilde{\\langle 0;t^{\\prime }|}H%\n\\widetilde{|0;t^{\\prime }\\rangle }\\right) {\\rm d}} \\def\\ri{{\\rm i}} \\def\\re{{\\rm e} t^{\\prime }.\n\\end{eqnarray}\n\n\\subsection*{Concluding Remarks}\n\n\\ \\ \\ In this article, we have studied fermionic system of nonstationary\nforced oscillator and we have constructed invariant ladder operators and the\nrelated Fock and coherent states. We succeeded to express these invariants\nand the time evolution of the corresponding states in terms of the same\nclassical equation, that describes the evolution of coherent states of the\nboson nonstationary (forced) oscillator \\cite{MMT70, Holz70}. The relation of\nthe invariant ladder operators method to the Lewis-Riesenfeld method \\cite{Lewis69}\nwas briefly described on the example of nonstationary fermion systems. \\\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{\\label{sec:level1}INTRODUCTION}\n\nIt is well recognized that the $\\alpha$ particle interaction with atomic nuclei is important in astrophysics \\cite{1}. Even if astrophysical reactions involving helium do not proceed through the strong $\\alpha$-cluster states (because of their high excitation energy), these states can provide $\\alpha$ width to the states that are closer to the region of astrophysical interest through configuration mixing.\n\nFor a long time, the surprising alpha cluster structure has been a stimulus for the development of classical shell model approaches (see \\cite{2} for new results). Additionally, work by the authors of Ref. \\cite{3} recently ``strengthened the theoretical motivation for experimental searches of alpha cluster states in alpha-like nuclei\" \\cite{3}. The authors of Ref. \\cite{3} related the nuclear structure in light systems of even and equal numbers of protons and neutrons with the first-order transition at zero temperature from a Bose-condensed gas of alpha particles (\\textsuperscript{4}He nuclei) to a nuclear liquid.\n\n\\textsuperscript{20}Ne nucleus presents a famous example of the manifestation of the alpha-cluster structure, and therefore it makes this nucleus a touchstone for \\textit{ab initio} approaches \\cite{4}. The nucleus \\textsuperscript{20}Ne is a benchmark case for the traditional shell model and its extension into algebraic and clustering domains. The well-established effective interaction Hamiltonians such as \\cite{5} not only shows an outstanding agreement with experimental data for \\textit{sd}-shell nuclei but also generates configuration mixing that shows transition to deformation and clustering.\n\nThe remarkable feature of \\textsuperscript{20}Ne nucleus shows up in the fact that almost all the observed states below 10 MeV can be classified into several overlapping rotational-like bands with the first one based on the ground 0$^+_1$ state. There are three other bands based on 0$^+$ levels: on 0$^+_2$ at 6.725 MeV, on a very narrow 0$^+_3$ at 7.191 MeV, on a very broad 0$^+_4$ at $\\sim$ 8.7 MeV \\cite{6} which are of evident cluster structure. The 0$^+_2$ and 0$^+_4$ bands have $\\alpha$+\\textsuperscript{16}O core structure as can be seen from their reduced $\\alpha$ particles widths, and probably the 0$^+_3$ band has predominant \\textsuperscript{12}C+\\textsuperscript{8}Be structure which manifests used itself in the selectivity of the \\textsuperscript{8}Be transfer reactions \\cite{7}. As the ground state band and the 0$^+_2$ band in \\textsuperscript{20}Ne can be related with similar structures in \\textsuperscript{16}O and \\textsuperscript{12}C, the ``additional\" structure of 0$^+_4$ states is not understood \\cite{4}. The cluster approaches \\cite{8} related the 0$^+_4$ band with large $\\alpha$-widths, starting with the so-called ``\\textsuperscript{16}O+$\\alpha$\" higher nodal band, which has one more nodal point in ``\\textsuperscript{16}O+$\\alpha$\" relative wave function than the lower bands have. However it appeared that there are too many bands with a similar structure.\n\nThe $\\alpha$ particle decay threshold in \\textsuperscript{20}Ne is 4.73 MeV, while the threshold for proton decay is at 12.8 MeV (neutron decay threshold is even higher). Therefore resonance $\\alpha$ particle scattering should be considered as an evident way to obtain data on the natural parity levels in \\textsuperscript{20}Ne up to 13 MeV excitation energy. Indeed the majority of adopted data \\cite{6} on level properties in the region in question based on a resonance work and an analysis made in 1960 year. More recently the $\\alpha$+\\textsuperscript{16}O resonance scattering experiment were developed further to backward angles and the data were reanalyzed using \\textit{R} matrix code Multi 6 \\cite{9}. The authors \\cite{9} obtained quite different results from those used in for many levels (see Table 1); in particular the broad 0$^+_4$ and 2$^+_4$ levels appeared even much more broader. The authors \\cite{9} noted difficulties of the fit in the region of 6-8 MeV excitation energies, in a region mainly free from narrow resonances. Evidently strong states of over 1 MeV width should influence a very broad excitation region (see for instance \\cite{10}).\n\nT. Fortune et al. \\cite{11,12} were the first who recognized the importance of the fact that single particle structure of the broad states is in drastic contradiction with the shell model predictions. They \\cite{11,12} also proposed the idea of mixing different configurations to explain the effect. The same idea was used in \\cite{13}. However, the authors of Refs. \\cite{11,12} used old data for the 0$^+_2$ state and used some estimates for the properties of the broad states proposed in Ref. \\cite{14} to support the idea. Later measurements \\cite{6} gave the width of 19 keV for the 6.72 MeV state (which is about 25\\% larger than that used in work \\cite{11}).\n\n\\begin{figure}[!t]\n \\begin{center}\n \\include{fig1}\n \\end{center}\n \\caption{E-T spectrum for the zero degrees detector. Alpha particles dominate; one can see also a weaker proton locus below the alpha particles.}\n\\end{figure}\n\nThe experimental aim of this work is to obtain new information on the structure of \\textsuperscript{20}Ne states, especially the broad 0$^+$, 2$^+$ states. Unlike other experimentalists, we used the Thick Target Inverse Kinematics method (TTIK) (see \\cite{15,16,17,18,19} and references therein) to study the excitation functions for the \\textsuperscript{16}O($\\alpha$, $\\alpha$)\\textsuperscript{16}O elastic scattering in the \\textsuperscript{20}Ne excitation region of 5.5-9.6 MeV and in a broad angular interval. On the theoretical side, we also used multi configuration shell model calculation to understand the limits of this approach in a description of the cluster states.\n\n\\section{\\label{sec:level2}EXPERIMENT}\n\n\\begin{figure}[!t]\n \\begin{center}\n \\include{fig2}\n \\end{center}\n \\caption{The \\textsuperscript{16}O($\\alpha$, $\\alpha$)\\textsuperscript{16}O elastic scattering excitation function at 180$^\\circ$ cm. The excitation energies in \\textsuperscript{20}Ne, E$_x$ in Table 1, are related with cm energy, E$_{cm}$, by expression, E$_x$=E$_{cm}$+4.73 MeV. The bold (red) line is the \\textit{R} matrix fit with the parameters of the present work. The dot (cyan) line is a fit with the 0$^+_4$ excitation energy of 8.3 MeV \\cite{11}, and dot (black) line is a fit with the 0$^+_4$ energy excitation energy of 8.62 MeV and the width of 1.472MeV \\cite{9}.}\n\\end{figure}\n\nThe experiment was performed at the DC-60 cyclotron (Astana) \\cite{17} which can accelerate heavy ions up to the 1.9 MeV\/A energy. While the TTIK method can't compete with a classical approach in terms of energy resolution, the possibility of observing excitation functions at and close to 180 degrees, where the resonance scattering dominates over the potential scattering, enables one to obtain more reliable information on the broad states. In the TTIK technique the inverse kinematics is used; and the incoming ions are slowed in a helium target gas. The light recoils, $\\alpha$ particles, are detected from a scattering event. These recoils emerge from the interaction with the beam ions and hit a Si detector array located at forward angles while the beam ions are stopped in the gas, as $\\alpha$ particles have smaller energy losses than the scattered ions. The TTIK approach provides a continuous excitation function as a result of the slowing down of the beam.\n\nFor the present experiment, the scattering chamber was filled with helium of 99.99\\% purity. The 30 MeV \\textsuperscript{16}O beam entered the scattering chamber through a thin entrance window made of 2.0 $\\mu$m Ti foil. Eight monitor Si detectors were placed in the chamber to detect \\textsuperscript{16}O ions elastically scattered from the Ti foil at 21$^\\circ$ angle. This array monitors the intensity of the beam with precision better than 4\\%. Fifteen 10x10 mm$^2$ Si detectors were placed at a distance of $\\sim$ 500 mm from the entrance window in the forward hemisphere at different laboratory angles starting from zero degrees. The gas pressure was chosen to stop the beam at distance of 40 mm from the zero degrees detector. The detector energy calibration and resolution ($\\sim$ 30 keV) were tested with a \\textsuperscript{226}Ra, \\textsuperscript{222}Rn, \\textsuperscript{218}Po and \\textsuperscript{214}Po $\\alpha$-source. The experimental set up was similar to that used before \\cite{19}, and more details can be found in Ref. \\cite{17,19}. The main errors in the present experimental approach are related to the uncertainties of the beam energy loss in the gas. To test the energy loss, we placed a thin Ti foil (2.0 $\\mu$m) at different distances from the entrance window. This can be used during the experiment without cycling vacuum. We found that the data \\cite{20} for energy loss of \\textsuperscript{16}O in helium are correct. The details of these tests will be published elsewhere. As a result, we estimated that the uncertainties in the absolute cross section are less than 6\\%. This conclusion was tested by comparison with the Rutherford cross sections at low energies. The agreement with the Rutherford scattering is within 5\\% error bars at all angles (see Fig.3).\n\nTogether with the amplitude signal, the Si detectors provided for a fast signal. This signal together with a ``start\" signal from RF of the cyclotron was used for the Time-of-Flight measurements. This E-T combination is used for particle identification in the TTIK approach \\cite{16,17,18}. Of course, only $\\alpha$ particles should be detected as a result of the interaction of \\textsuperscript{16}O with helium at the chosen conditions. However, protons can be created in the Ti window, and protons can appear due to hydrogen admixtures in the gas. Indeed, we have observed a weak proton banana, likely as a result of reactions in the window. These protons were easily identified by TF and separated from the $\\alpha$ particles, as seen in Fig.1.\n\n\\section{\\label{sec:level3}Experimental Results and Discussion}\n\n\\begin{figure}[!t]\n \\begin{center}\n \\include{fig3}\n \\caption{\\textit{R} matrix fit (bold red curve) of the excitation functions for the $\\alpha$+\\textsuperscript{16}O elastic scattering. (a) The dashed (cyan) curve presents the data of the 0$^+$ level at the 8.3 MeV excitation energy \\cite{11} and (e) dot (black) line is a fit with 2$^+$ level at the excitation energy of 9.0 MeV \\cite{6}.}\n \\end{center}\n\\end{figure} \n\n\\begin{table*}[!t]\n\\label{tab:1}\n\\caption{\\textsuperscript{20}Ne levels}\n\\begin{center}\n\\begin{tabular}{lccccccccccclccl}\n\\hline\n\\hline\n\\multirow{1}{*}{N} & \\multicolumn{3}{c}{TUNL data \\cite{6}} & \\multirow{1}{*}{} & %\n \\multicolumn{2}{c}{H. Shen et al. \\cite{9}} & \\multirow{1}{*}{} & %\n \\multicolumn{3}{c}{This work} & \\multirow{1}{*}{} & \\multicolumn{4}{c}{CNCIM} \\\\\n\\cline{2-4}\\cline{6-7}\\cline{9-11}\\cline{13-16}\n & E$_x$ & J$^\\pi$ & $\\Gamma_\\alpha$ & & E$_x$ & $\\Gamma_\\alpha$ & & E$_x$ &\n $\\Gamma_\\alpha$ & $\\gamma_\\alpha$ & & E$_x$ & J$^\\pi$ & SF$_p$ & SF$_\\alpha$\\\\\n & (MeV) & & (keV) & & (MeV) & (keV) & & (MeV) & (keV) & & & (MeV) & & & \\\\\n\\hline\n\n1 &\t0 & 0$^+_1$ & & & - & - & & 0 & & Large & & 0 & 0$^+$ & 0.36 & 0.73\\\\\n2 &\t1.63 & 2$^+_1$ & & & - & - & & 1.63 & & Large & & 2.242 & 2$^+$ & 0.41 & 0.67\\\\\n3 &\t4.25 & 4$^+_1$ & & & - & - & & 4.25 & & Large & & 4.58 & 4$^+$ & & 0.62\\\\\n\n4 &\t5.78 & 1$^-$ & (28$\\pm$3)x10$^{-3}$ & & - & - & & 4.45 & 0.03 & 1.4\\\\\t\t\t\t\n5 &\t6.73 & 0$^+_2$ & 19$\\pm$0.9 & & 6.72 & 11 & & 6.78 & 20.6 & 0.47 & & 6.94 & 0$^+_3$ & 0.55 & 0.46\\\\\n6 &\t7.16 & 3$^-$ & 8.2$\\pm$0.3 & & 7.16 & 10 & & 7.18 & 8.3 & 1.37\\\\\t\t\t\t\n7 &\t7.19 & 0$^+_3$ & 3.4$\\pm$0.2 & & 7.19 & 5 & & 7.20 & 3\t& 0.019 & & 6.27** & 0$^+_2$ & 0.055 & 0.44**\\\\\n8 &\t7.42 & 2$^+_2$ & 15.1$\\pm$0.7 & & 7.43 & 7 & & 7.44 & 14.3 & 0.19 & & 7.39 & 2$^+_3$ & 0.01 & 0.12\\\\\n9 &\t7.83 & 2$^+_3$ & 2 & & 7.83 & 1 & & 7.85 & 3.68 & 0.01 & & 7.15** & 2$^+_2$ & 0.12 & 0.18**\\\\\n\n10 & 8.45 & 5$^-$ & 0.013$\\pm$0.004 & & 8.45 & 0.02 & & 8.45 & 0.013\\\\\t\t\t\t\t\n11 & 8.71 & 1$^-$ & 2.1$\\pm$0.8 & & & & & 8.71 & 3.5\\\\\t\t\t\t\t\n12 & $\\approx$8.7 & 0$^+_4$ & $>$800 & & 8.62 & 1470 & & 8.77$\\pm$0.15 & 750$\\pm$220 & $\\sim$0.25 & & 9.66** & 0$^+_4$ & 0.002 & 0.18**\\\\\n13 & 8.78 & 6$^+_1$ & 0.11$\\pm$0.02 & & - & - & & 8.78 & 0.14 & 0.5 & & 9.49 & 6$^+$ & & 0.51\\\\\n14 & 8.85 & 1$^-$ & 19 & & 8.84 & 27 & & 8.85 & 18.0\\\\\t\t\t\t\t\n15 & 9.00 & 2$^+_4$ & $\\approx$800 & & 8.87 & 1250 & & 8.79$\\pm$0.10 & 695$\\pm$120 & 0.86 & & 8.36** & 2$^+_4$ & 0.02 & 0.02**\\\\\n16 & 9.03 & 4$^+_3$ & 3 & & 9.02 & & & 9.03 & 1.9 & 0.03 & & 9.0 & 4$^+$ & & 0.09\\\\\n\n17 & 9.12 & 3$^-$ & 3.2 & & 9.09 & 4 & & 9.13 & 4.1\\\\\n18 & 9.19 & 2$^+$ & & & - & - & & (9.29) & $\\leq$10\t\\\\\t\t\t\t\n19 & 9.48 & 2$^+$ & 29$\\pm$15 & & 9.48 & 46 & & 9.48 & 65$\\pm$20 & 0.02?\\\\\t\t\t\t\n20 & 9.99 & 4$^+_4$ & 155$\\pm$30 & & 10.02 & 150 & & 9.97 & 157 & 0.38 & & 9.5 & 4$^+$ & & 0.009\\\\\n21* & 10.26 & 5$^-$ & 145$\\pm$40 & & 10.26 & 190 & & 10.26 &\t& 1.9\t\\\\\t\t\t\n22 & 10.41 & 3$^-$ & 80 & & 10.40 & 101 & & 10.41 & &\t\\\\\n\n23 & 10.58 & 2$^+$ & 24 & & 10.56 & 15 & & 10.58 & & & & 10.2 & 2+ & 0.005 & 0.04\\\\\n24 & 10.80 & 4$^+_4$ & 350 & & 10.75 & 400 & & 10.80 & & & & 10.7 & 4$^+$ & & 0.04\\\\\n25 & 10.97 & 0$^+_5$ & 580 & & 10.99 & 700 & & 10.97 & & & & 11.9 & 0$^+$\t\\\\\t\n26 & 11.24 & 1$^-$ & 175 & & 11.19 & 85 & & 11.24 & &\\\\\t\t\t\t\t\n27 & 11.95 & 8$^+$ & (3.5$\\pm$1.0)x10$^{-2}$ & & & & & 11.95 & & 0.35 & & 11.50 & 8$^+$ & & 0.40\\\\\n\\hline\n\\hline\n\\multicolumn{16}{l}{ * For the levels with numbers 21-27 the parameters of the present fit were fixed as in \\cite{6} } \\\\\n\\multicolumn{16}{l}{ ** Calculated in \\textit{psd} space. SF is to the first excited state in \\textsuperscript{16}O; SFs for the ground state in \\textsuperscript{16}O are $\\leq$ 0.1 } \\\\\n\n\\end{tabular}\n\\end{center}\n\\end{table*}\n\nThe experimental excitation functions were analyzed using multilevel multichannel \\textit{R} matrix code \\cite{21}. The calculated curves were convoluted with the experimental energy resolution. The experimental energy resolution was $\\sim$ 30 keV at zero degrees and deteriorated up to $\\sim$100 keV with angles estranging from zero degrees. We did not notice a deterioration of the energy resolution with the energy loss of the beam in the chamber. As it seen in Table 1 the excitation energies of the resonances of the present work agree with the adopted ones \\cite{6} within 10-15 keV. This agreement is an evidence of our overall good energy calibrations and the correct account of the ion energy loss in helium in the present work. Fig.2 and Fig.3 give the experimental excitation functions together with the present \\textit{R} matrix fit. Fig.2 demonstrates the data at 180$^\\circ$ and illustrates the differences in the fits due to different parameters of the broad 0$^+_4$ resonance. The data on the resonances used in the present \\textit{R}-matrix fit are summarized in the Table 1 together with the adopted data \\cite{6}. Data of the last \\textit{R} matrix analysis \\cite{9} are also given in Table 1. The analysis \\cite{9} resulted in level parameters which are often different from the adopted.\n\nOur analysis (Table 1) resulted in small discrepancies with the data \\cite{6} in the detail description of the narrow states (widths less than 10 keV). The \\textit{R}-matrix code \\cite{21} is tuned for the TTIK measurements and for the analyses of states with a width of over 10 keV to accelerate automatic fit calculations. Therefore the small disagreements for the narrow states are not significant. We focused on the broader states and on the part of the excitation function changing slowly with energy and angle.\n\nOur analysis indicates that all strong alpha cluster states at 5$\\sim$6 MeV below or up the investigated excitation energy region can influence the \\textit{R}-matrix fit. Therefore we included in the fit the \\textsuperscript{20}Ne ground state, the first 2$^+$ and 4$^+$ states and 1$^-$ (5.67 MeV) state in the fit, even though below the investigated region (see Table 1). Among these states, only the 1$^-$ state is above the $\\alpha$ particle decay threshold in \\textsuperscript{20}Ne; the reduced width of this state is known, and it is large. Shell model calculations also give very large spectroscopic factors for all members of the ground state band. Indeed, a good agreement needs large values of the corresponding amplitudes (over 0.7). Above the investigated excitation energy region, high spin $\\alpha$-cluster resonances are mainly known. Each of these resonances (see Table 1) considered separately influences the fit, especially at 180$^\\circ$. However, their joint influence is much weaker. This cancellation is due to different parities of the spins. Only the influence of the closest to the investigated region, the 4$^+$ (9.99 MeV) resonance, can be noticed. A somewhat better fit needs the width of this resonance to be slightly larger $\\sim$160 keV (well within the quoted uncertainties, see Table 1). Parameters of all other resonances above the investigated region were fixed according to the data of Ref. \\cite{6}. A good general fit ($\\chi^2$=1.1) was reached in this way without any backward resonance inclusion.\n\n\\begin{table*}[!t]\n \\caption{$\\alpha$+\\textsuperscript{12}C levels in \\textsuperscript{16}O}\n \\centering\n \\begin{tabular}{ccccccccc}\n \\hline\n \\hline \n \\multicolumn{1}{c}{\\textsuperscript{16}O level} & \\multirow{1}{*}{} & \\multicolumn{1}{c}{$\\Gamma_\\alpha$ $_{exp}$ keV \\cite{6} } & \\multirow{1}{*}{} & \\multicolumn{1}{c}{$\\gamma_\\alpha$(1); -V$_0$ MeV} & \\multirow{1}{*}{} & \\multicolumn{1}{c}{$\\gamma_\\alpha$(2); -V$_0$ MeV} & \\multirow{1}{*}{} & \\multicolumn{1}{c}{$\\gamma_\\alpha$(3); -V$_0$ MeV}\\\\ \n \\cline{1-1}\\cline{3-3}\\cline{5-5}\\cline{7-7}\\cline{9-9}\n 1$^-$; 9.58 MeV & & 420$\\pm$20 & & 0.70; 138.2 & & 0.72; 150.0 & & 0.84; 158.5 \\\\\n 4$^+$; 10.36 MeV & & 26$\\pm$3 & & 0.68; 125.3 & & 0.88; 139,6 & & 1.21; 143.2 \\\\\n \\hline\n \\hline \n \\end{tabular}\n \\label{tab:2}\n\\end{table*}\n\nAll resonances are at the maximum at 180$^{\\circ}$. The broad hump at this angle at cm energy of 4 MeV (Fig. 2) is a clear indication for the presence of low spin states, 0$^+$ and 2$^+$. Higher spin states are narrower. A single level (2$^+$) cannot produce the strong peak at different angles, and levels of different parity, such as 2$^+$ and 1$^-$, interfere destructively at 180$^{\\circ}$. The present analysis resulted in two practically degenerate states at $\\sim$ 8.8 MeV with the same width (see Table 1). Our results for the broad 2$^+$ are rather close to the adopted values. However, if this level is moved to 9.0 MeV excitation energy (as in \\cite{6}) then the fit becomes worse, especially in the vicinity of the dip die to the presence of another 2$^+$ level at 9.48 MeV (Fig. 3(e)). The fit becomes even worse with a very broad 2$^+$ level resulting from the parameters of Ref. \\cite{9}.\n\nT. Fortune at al., \\cite{11} observed a broad distribution with a center at 8.3 MeV excitation energy in \\textsuperscript{20}Ne and related it with the 0$^+$ level. Fig. 2 presents \\textit{R} matrix calculations with the 0$^+$ level to be moved to 8.3 MeV excitation energy. This move destroyed the good fit. Fig.2 shows also that a very broad 0$^+$ level of the fit \\cite{9} destroys the agreement.\n\nThe 18$^{th}$ (2$^+$) level of \\cite{6} is in the energy region of the present investigation. It was observed in a single work in a study of \\textsuperscript{20}Na $\\beta^+$ decay. We have not found any reliable evidence for the presence of this state. If it exists, its width should be less than 10 keV. We observe a fluctuation of points which might be associated with a narrow 2$^+$ level at excitation energy of 9.29 MeV.\n\nWe noted that the 19$^{th}$ level, 2$^+$, has the adopted \\cite{6} excitation energy in our fit but a different width of 65$\\pm$20 keV. While the two times difference with Ref. \\cite{6} in the widths is marginally exceeds the error bars, the influence on the better fit is evident. The adopted data \\cite{6} for this level are based on the results of a single older work \\cite{22}. The authors \\cite{22} observed a weak $\\gamma$ decay of this state in the presence of a large background. A broader level than in Ref. \\cite{22} was also found in work \\cite{9} as shown in Table 1.\n\nWe characterized the alpha-cluster properties of the states above the alpha particle decay threshold by SF=$\\gamma_\\alpha$=$\\Gamma_\\alpha$ $_{exp}$\/$\\Gamma_\\alpha$ $_{calc}$, where $\\Gamma_\\alpha$ $_{calc}$ is the single alpha particle width calculated in the $\\alpha$-core potential. To normalize SFs we calculated these values for the well-known alpha-cluster states in \\textsuperscript{16}O.\n\nThe Woods-Saxon potential was used to calculate the limit ($\\Gamma_\\alpha$ $_{cal}$) as the widths of single particle states in the potential. First we tried to fit the widths of known alpha-cluster states, 1$^-$ and 4$^+$, in \\textsuperscript{16}O so that $\\gamma_\\alpha$ = $\\Gamma_\\alpha$ $_{exp}$\/$\\Gamma_\\alpha$ $_{cal}$ $\\sim$ 1.0. The real part of the potential was changed to fit the binding energy of the states. The radius of the potential was chosen to be R = r$_0$ $\\times$12$^{1\/3}$; the Coulomb potential was taken in to account as a charge sphere potential with R$_{coul}$ = R. We made first calculations (1) with r$_0$ = 1.31 fm and the diffuseness a = 0.65 fm, then we set r$_0$ = 1.23 fm (2), and we finally performed the third calculations (3) with r$_0$ = 1.23 fm and a = 0.6 fm. The results are summarized in Table 2. The $\\gamma_\\alpha$ calculations for the \\textsuperscript{20}Ne states were made with the final (3) parameters should be compared with SF given by a theory.\n\n\\section{\\label{sec:level4}Theoretical description of the $\\alpha$ cluster states in \\textsuperscript{20}$\\textbf{Ne}$}\n\nCNCIM \\cite{2} is among the latest developments of the classical shell model approaches towards clustering. This model targets a combination of classical configuration interaction techniques with algebraic methods that emerge in the description of clustering. The ability to construct a fully normalized set of orthogonal cluster channels is at the core of this approach; the overlaps of the shell model states with these channels are associated with spectroscopic factors and compared in Table 1 with the reduced widths obtained earlier. The CNCIM allows us to study clustering features that emerge in models with well-established traditional shell model Hamiltonians. These effective model Hamiltonians are built from fundamental nucleon-nucleon interactions followed by phenomenological adjustments to select observables; thus, they generally describe of a broad scope of experimental data with high accuracy. Apart from using these phenomenological shell model Hamiltonians, our study does not involve any adjustable parameters. In order to fully explore the problem, we considered several different model spaces and corresponding Hamiltonians: the \\textit{sd} model space with USDB interaction \\cite{5}; unrestricted \\textit{p-sd} shell model Hamiltonian \\cite{23}, the same Hamiltonian has been used in Refs. \\cite{2}; WBP Hamiltonian \\cite{24} allowing 0$\\hbar\\omega$, 1$\\hbar\\omega$ and 2$\\hbar\\omega$ excitations in \\textit{p-sd-pf} valence space; and the \\textit{sd-pf} Hamiltonian \\cite{25}. This sequence of Hamiltonians represents and expansions of the valence space from \\textit{sd} to \\textit{p-sd}, to \\textit{p-sd-pf}. All models are in good agreement for the \\textit{sd}-states; the low-lying negative parity states as well as positive 2ph excitations are dominated by the \\textit{p-sd} configurations. Thus, in Table 1 we only include the results from the \\textit{p-sd} Hamiltonian which turned out to be most representative although the following discussion and conclusions are largely based on comparisons. The lowest states associated with significant \\textit{fp} shell component appear at excitation energies above 15 MeV.\n\nShell model calculations for \\textsuperscript{20}Ne with open \\textit{2s-1d (sd)} shells predict well the ground state band. The structure of this band is based on the dominating SU(3) configuration (about 75\\%) with quantum numbers (8,0). The model predicts (Table 1) large SF for all members of the band based on the ground state in \\textsuperscript{20}Ne. The 0$^+$, 2$^+$, and 4$^+$ members of the band are below the $\\alpha$ particle decay threshold and do not have observable $\\alpha$ particle widths. However, large $\\alpha$ cluster SFs for these states provide for a better \\textit{R} matrix fit. While uncertainties for the \\textit{R} matrix amplitudes for these states are large, the fit (Fig. 2 and 3) requires these amplitudes to be close to those of the negative parity states with the known extreme $\\alpha$ cluster structure. The $\\alpha$ particle widths of the highest 6$^+$ and 8$^+$ members of this band are known. There is a long history of attempts and ideas to describe these widths using shell model approaches (see, for instance \\cite{5}). The most calculations predicted large clustering for the band but could not explain the decrease of the reduced width for the 6$^+$ and 8$^+$ members. As one sees in Table 1, the CNCIM calculations are in fine agreement with the experimental data for these states. All members of this band have significant clustering that diminishes at higher energies due to configuration mixing. The second 0$^+$ state within \\textit{sd} space appears at around 6.7 MeV of excitation (in \\textit{psd} model in Table 1, this is a third 0$^+$ state at 6.9 MeV). This level also has a substantial clustering component and absorbs nearly all 15\\% of the remaining strength of the SU(3) (8,0) component. The following 2$^+_2$ (in experiment and in \\textit{sd} model, but third in \\textit{p-sd} model as discussed in what follows) at about 7.4 MeV of excitation energy being a member of the 0$^+_2$ band can be described in a similar way.\n\nA defrost of the \\textit{1p} shell (which is filled in \\textsuperscript{16}O) results in the doubling of the levels, as it is observed and is shown in Table 1 new levels 0$^+$ and 2$^+$, marked with asterisks, appear. In our calculations, the ordering in energy is reversed for both doublets. Structurally the members of each doublet are very different which allows them to be so close in energy and inhibits configuration mixing and Wigner repulsion. One of the doublet levels (the lower 0$^+$ and 2$^+$ levels) has a large $\\alpha$ cluster SF relative to the first excited state in \\textsuperscript{16}O and much smaller SF relative to the ground state in \\textsuperscript{16}O (pay attention that the theory gives the wrong order for the levels in question, see Table 1). Indeed the predicted difference in the structure is supported by population selectivity in different nuclear reactions. The 6.72 MeV 0$^+$ and 7.42 MeV 2$^+$ are populated much stronger than neighboring 7.20 and 7.83 MeV levels in the \\textsuperscript{16}O(\\textsuperscript{6}Li, d) reaction \\cite{26}. The opposite is the case in the \\textsuperscript{12}C(\\textsuperscript{9}Be, n) or \\textsuperscript{12}C(\\textsuperscript{12}C,$\\alpha$) \\cite{27} reactions. \n\nThe theory gives large single particle spectroscopic factors for the \\textit{sd} states and smaller for the 7.20 and 7.83 MeV states. Indeed, one expects that states in \\textsuperscript{20}Ne with a hole in the \\textit{p1\/2} shell will be weakly excited in the single nucleon transfer, \\textsuperscript{19}F(\\textsuperscript{3}He, d) reaction in accordance with the experimental data \\cite{28}. The experiment \\cite{11,12,28} supports also detailed single particle SF calculations giving SF for the ground state smaller than for the 0$^+$ 6.73 MeV state and much higher SF for the first excited 2$^+$ than for 2$^+$ member of the band based on the 6.73 MeV state.\n\nIn the \\textit{sd} valence space (USDB) the third 0$^+$ state appears only at 11.9 MeV, and it has a relatively small alpha spectroscopic factor; opening of the \\textit{p}-shall in addition to 0$^+$ at 6.27 MeV, leads to 0$^+$ state at 9.7 MeV. However, the predicted state has a low proton spectroscopic factor which is not as well supported by observations. A similar serious discrepancy is observed with a broad 4$^{th}$ 2$^+$ state at around 9 MeV, both \\textit{sd} and \\textit{p-sd} shell models produce candidates but with very low alpha SF. As evident from our studies the only strong coupling to alpha channels could come from \\textit{fp} shell and higher shells. The lowest two bands saturate the alpha strength within \\textit{sd} configurations, holes in the \\textit{p}-shell do not lead to a significant contribution due to low level of core excitation in the ground state of \\textsuperscript{16}O. While our models predict high excitation energies of states with significant \\textit{fp} components, we can speculate that strong configuration mixing, collective effects such as deformation, and coupling to continuum via super radiance mechanism \\cite{29,30} can enhance admixture needed to reproduce the broad resonances observed. There is a similar problem with $\\alpha$ cluster negative parity states 1$^-$ and 3$^-$, the \\textit{p-sd} Hamiltonian produces an acceptable spectrum but the alpha spectroscopic factors are low (see also \\cite{31}).\n\n\\section{\\label{sec:level5}Conclusions}\n\nIn this work we study $\\alpha$-clustering in \\textsuperscript{20}Ne. This nucleus is a benchmark example of many theoretical techniques targeting clustering in light nuclei. Our \\textit{R}-matrix analysis of TTIK experimental data confirms previously known results and establishes new constraints for the positions and widths of the resonances. We compared our findings with those obtained theoretically using cluster-nucleon configuration interaction approach developed in several previous works \\cite{2} and references therein. There is good overall agreement between theoretically predicted and observed spectra. Our theoretical approach describes very well the ground state band and the band built on the first 0$^+$ state. Allowing cross shell excitations from the \\textit{p}-shell, it was possible to reproduce the band built on the second 0$^+$ state. For these states, all spectroscopic factors for alpha transitions to the ground state of \\textsuperscript{16}O and to the first excited state in \\textsuperscript{16}O as well as proton spectroscopic factors to the ground state of \\textsuperscript{19}F are well reproduced. The situation is not as good when it comes to resonances 1$^-$ 3$^-$ and 4$^{th}$ 0$^+$ and 4$^{th}$ 2$^+$, all of these states are broad and have exceptionally large alpha spectroscopic factors. \n\nIn order to describe strong clustering features these states must include configurations from \\textit{fp} shell and from higher oscillator shells, however Hamiltonians that we explored predict these contributions to be negligible below 15 MeV of excitation. Thus, inability of theoretical models to describe broad states exclusively while working well elsewhere suggests an additional coupling mechanism unaccounted for in the traditional shell model Hamiltonians. The super radiance suggested in Refs. \\cite{29,30} could provide this mechanism. Alternatively, the problem could be associated with relatively unknown cross shell interactions. Therefore, work shows that the experimental study of alpha clustering represents an outstanding tool for exploring cross shell excitations especially those of multi-particle multi-hole nature.\n\n\\begin{acknowledgments}\nThis work was supported by Ministry of Education and Science of the Republic of Kazakhstan (grant number \\#0115\u0420\u041a03029 ``NU-Berkeley\", 2014-2018; grant number \\#0115\u0420\u041a022465, 2015-2017). This material is also based upon work supported by the U.S. Department of Energy Office of Science, Office of Nuclear Physics under Grants No DE-FG02-93ER40773 and No. DE-SC0009883. G.V.R is also grateful to the Welch Foundation (Grant No. A-1853).\n\\end{acknowledgments}\n\n\\nocite{*}\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}}