diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzkujl" "b/data_all_eng_slimpj/shuffled/split2/finalzzkujl" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzkujl" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction}\n\nBoolean Satisfiability Problem (SAT) is a significant NP-complete problem with numerous practical applications -- product configuration, hardware verification, and software package management, to name a few. There is no single efficient algorithm that solves every SAT problem, but heuristics have been developed that are sufficient for solving many real-life tasks.\n\nNeural networks have demonstrated superior performance on many complex tasks, e.g., image and language processing, protein folding, and playing board and computer games. So we may expect that, in the long run, neural networks could also replace handcrafted heuristics for SAT problems. \n\nThere are attempts for applying neural networks to solve SAT by integrating them in contemporary solvers or replacing them altogether. \\cite{selsam2018learning} proposes training NeuroSAT architecture with single-bit supervision, predicting whether the SAT instance is satisfiable.\nYet, it does not directly produce variable assignments and requires labels for the training that may be troublesome to obtain. \\cite{amizadeh2018learning} introduces unsupervised loss for training a neural network to solve a related Circuit-SAT problem without requiring labels. They use differentiable constraint relaxation to evaluate the network output and penalise the network for violating constraints. That leads to successful training for Circuit-SAT problems.\n\nUsually, neural networks are trained by minimizing the loss which is obtained at the final layer. We ask whether it is beneficial to let the network know the loss it is currently producing and adapting its output correspondingly. That is not possible with the traditional supervised losses since labels are not available at the inference time. Yet, such an approach is sound for an unsupervised loss and, indeed, provides an excellent basis for designing a neural SAT solver. \n\nIn this paper, we introduce a step-wise recurrent neural SAT solver that at each step comes up with a query of variable assignments, evaluates it with an unsupervised loss, and updates its state based on the evaluation results. At every step, it outputs a target solution that may contain information from all the queries performed so far. \n\nWe prove that the proposed query mechanism employing unsupervised loss function as query evaluator can provide the neural network with information about (a) satisfiability status of a solution if queried at integer points, (b) reveal the structure of the problem instance if queried at fractional (real) points. That provides rich information about the structure, meaning of the problem, and its solution.\n\n\\begin{figure}[t]\n \\centering\n \\includegraphics[width=\\columnwidth]{figures\/query_making.pdf}\n \\caption{The proposed query mechanism works by producing a query, evaluating it using an unsupervised loss function, and passing the resulting value back to the neural network for interpretation. It allows the model to obtain the structure and meaning of the solvable instance and information about the expected model output. The same unsupervised loss can be used for evaluating the query and for training.\n }\n \\label{fig:query_overview}\n\\end{figure}\n\nThe proposed architecture is evaluated on several SAT problems, and it outperforms the neural baseline on all of them. We also compare it with classical SAT solvers (Glucose 4 and GSAT) on 3-SAT and SHA-1 preimage attack tasks. Appendix of this paper is available on \\url{https:\/\/github.com\/LUMII-Syslab\/QuerySAT\/blob\/master\/appendix.pdf}.\n\n\\section{Query mechanism}\\label{sec:querysat}\nWe hypothesize that allowing a neural network to make several solution trials at the runtime and getting feedback about them can significantly improve network capabilities in some cases. It is easy to implement almost for any neural network, yet depending on the unsupervised loss function, some networks can be especially suitable for such augmentation.\n\nThe simplest such implementation (depicted in Fig. \\ref{fig:query_overview}) consists of two different neural network layers. The first layer ($NN_A$) is given the current state ($s_{r}$) and it outputs a query ($q$) and some hidden state ($h$). We then use the unsupervised loss function ($loss$) to evaluate the query and obtain evaluation results ($e$). It is important to use a loss function that does not require labels for evaluation, as they may not be available at the test time. The evaluation results ($e$) together with the hidden state ($h$) are then passed as the input for the second layer ($NN_B$) that is responsible for interpreting the evaluated query and producing the next state ($s_{r+1}$) and output logits ($l$). We can calculate the gradient of the evaluation results ($e$) with respect to query ($q$). The gradient can be optionally added as the third input for the second layer to show the network the direction for decreasing the loss \\cite{andrychowicz2016learning}. The query mechanism is conceptually shown in Fig. \\ref{fig:query_overview} and can be defined as follows:\n\\begin{align*}\n q, h &= NN_{A}(s_{r}) \\\\\n e &= loss(q) \\\\\n s_{r+1}, l &= NN_{B}(h,\\; e,\\; \\nabla_q e) \\\\\n \\mathcal{L} &= loss(l).\n\\end{align*}\n\nCalculating output logits and loss $\\mathcal{L}$ at every step is not noteworthy per se but merely demonstrates that the same loss function can be used for both -- query evaluation and model training. The training loss in principle could be evaluated only once -- at the network's final layer and can be different from the loss used for the query evaluation. Nevertheless, it is essential not to merge the loss (and logits) used for query with those used for training. Although constructed similarly, their meaning and functioning are different (see Appendix \\ref{apx:query_use}).\n\n\\section{Query mechanism for SAT}\\label{sec:query_mechanism}\n\nWe use the Boolean Satisfiability Problem (SAT) as the testbed for validating the benefits of the query mechanism. To this end, we propose an unsupervised loss function for SAT problems in conjunctive normal form and theoretically show that the query mechanism is indeed beneficial -- it gives the model access to the problem structure and contributes to the performance of the model. \n\n\\subsection{Boolean Satisfiability Problem}\nBoolean Satisfiability Problem (SAT) questions whether there exists an interpretation (True or False assignments to the variables) that satisfies the given Boolean formula. We undertake a related problem -- finding a set of variable assignments that satisfies the given Boolean formula. Throughout the paper, we stick to common practice \\cite{biere2009handbook} and represent SAT formulas solely in a conjunctive normal form (CNF) -- conjunction of one or more clauses where a clause is a disjunction of literals.\n\nAny SAT formula can be naturally represented as a bipartite variables-clauses graph, likewise known as an SAT factor graph \\cite{biere2009handbook}. In such a graph, the edge between variables and clauses graph exists whenever the variable is present in the clause. Two types of edges are used to distinguish a variable from its negation.\nVariables-clauses graph of SAT formula with $n$ variables and $m$ clauses can be represented as a sparse $n \\times m$ adjacency matrix. In order to perform batching, several SAT instances can be placed into a single factor graph yielding a single adjacency matrix for the whole batch.\n\n\\subsection{Unsupervised SAT loss}\\label{sec:sat_loss}\nA common way to train neural networks is by using a supervised loss, such as cross-entropy, which matches the network outputs with the correct labels. But such an approach does not work for training variable assignment for SAT due to several possible satisfying assignments for a single SAT instance. Also, obtaining labels involve SAT solving, which is time-consuming for large instances. Moreover, as we want to integrate the loss function into the neural network as a query mechanism, it has to be differentiable and cannot rely on labels as they are not available at the test time. \n\nTherefore, we design a differentiable unsupervised loss function that directly optimizes towards finding a satisfiable variable assignment of the Boolean formula without knowing a correct solution. To this end, we relax the Boolean domain to continuous variables in the range $[0, 1]$ where 0 corresponds to False and 1 to True. The vector of all variable values in the assignment is denoted by $x$. The value $V_c$ for a clause $c$ is obtained by multiplying together negations of all literals in the clause and negating the result and the loss value $\\mathcal{L_{\\phi}}$ for a formula $\\phi$ by multiplying all the clause values of $\\phi$ together: \n\\begin{align*}\n V_c(x) &= 1-\\prod_{i \\in c^+}(1-x_i)\\prod_{i \\in c^-}x_i, \\\\\n \\mathcal{L_{\\phi}}(x) &= \\prod_{c \\in \\phi}V_c(x), \\\\\n\\end{align*}\nwhere $x_i$ is the value of $i$-th variable and $c^+$ gives the set of variables that occur in the clause $c$ in the positive form and $c^-$ in the negated form. \nThe values $V_c(x)$ and $\\mathcal{L_{\\phi}}(x)$ are equal to 1 if and only if $x$ is a satisfying assignment of clause $c$ or formula $\\phi$, respectively, and strictly smaller otherwise. So by maximizing $\\mathcal{L}_{\\phi}(x)$ we can hope to find a satisfying assignment to the variables. But this function is not usable in practice since it often yields zero value in machine precision due to multiplying together many clause values, which are all less than one. Therefore, we use the negative logarithm of $\\mathcal{L}_{\\phi}$ and minimize it:\n\\begin{align*}\n \\mathcal{L}_{\\phi}^{\\log}(x) &= -\\log(\\mathcal{L}_{\\phi}(x)) = -\\sum_{c \\in \\phi}\\log(V_c(x)).\n\\end{align*}\nTaking the logarithm of the loss does not change its minimum\/maximum structure, so minimizing $\\mathcal{L}^{\\log}_{\\phi}$ is equivalent of maximizing $\\mathcal{L}_{\\phi}$.\n\nA model trained by minimizing $\\mathcal{L}^{\\log}_{\\phi}$ produces \"soft\" variable assignments in the range $[0,1]$. We experimentally observe that when the network can find a solution, it is close to binary and rounding the values to the nearest integer produces excellent results. \n\nOther works \\cite{Ryan2020CLN2INV:, pmlr-v97-fischer19a}, albeit in different contexts, have also proposed relaxing SAT formulas to continuous truth values similar to $\\mathcal{L}_{\\phi}$ loss (one not in the log-space) but faces the same problems as $\\mathcal{L}_{\\phi}$ if used for learning SAT solutions directly. Therefore, log-space formulation $\\mathcal{L}^{\\log}_{\\phi}$ of the loss trades capabilities of modelling general Boolean formulas for a loss function that gives non-zero loss for large CNF formulas and can be directly applied for learning their solutions. \n\n\\subsection{Power of the query}\nThe proposed loss function, when used in the query mechanism, gives the satisfiability status of a solution if queried at the binary points 0 or 1 and reveals the structure of the formula if queried at the intermediate (real) points. To maximize the amount of information gained by the query and to match the internal structure of Graph Neural Network (GNN), we employ queries that return the loss value $V_c(x)$ for each clause at the query point $x$. The power of such query mechanism is formulated in the Theorem \\ref{thm:int_points} and \\ref{thm:instance_to_be_solved}, respectively. The proofs of theorems are deferred to the Appendix \\ref{apx:int_points} and \\ref{apx:instance_to_be_solved}.\n\n\\begin{restatable}{theorem}{thintpoints}\\label{thm:int_points}\nFor a binary query point $x$ the losses $\\mathcal{L}_{\\phi}(x)$ or $V_c(x)$ are equal to 1 if the formula $\\phi$ or clause $c$ is satisfied and 0 otherwise.\n\\end{restatable}\n\n\\begin{restatable}{theorem}{thsolvedinstances}\\label{thm:instance_to_be_solved}\nA single query that returns the loss $V_c$ for each clause $c$ is sufficient to uniquely identify the SAT formula $\\phi$ to be solved.\n\\end{restatable}\n\nThe proof of theorem \\ref{thm:int_points} follows immediately from the definitions of the losses. The proof of theorem \\ref{thm:instance_to_be_solved} shows how to create a query point $x$ in such a way that the literals making up the clauses can be uniquely decoded solely from the clause losses. The proof assumes that we know the variable and clause count of the formula, but this assumption can be relaxed by padding all the formulas to some fixed maximum size. A stronger result that a single query returning only the total loss $\\mathcal{L}_{\\phi}$ (a single real number) might also be shown by using more intricate reasoning, but we did not pursue this direction since we use per-clause loss. \n\nThe construction used in theorem \\ref{thm:instance_to_be_solved} assumes sufficiently high precision of the numbers, exceeding the limits of standard floating-point arithmetic. Regarding that, in our implementation, we issue several queries in parallel (up to 128) and organize the computation in multiple recurrent steps in which queries are performed repeatedly and one query can depend on the results of the previous one. Such design significantly enhances the power of queries and relieves the need for high precision.\n\n\n\\section{QuerySAT}\nWe validate the theory in practice by building a neural SAT solver that we call QuerySAT and evaluate its performance on a wide range of SAT tasks - k-SAT, 3-SAT, 3-Clique, k-Coloring, and SHA-1 preimage attack. Experimental evaluation is performed on a single machine with 16GB RAM and a single Nvidia T4 GPU (16GB) using AdaBelief optimizer \\cite{zhuang2020adabelief}. Code for reproducing experiments is implemented in TensorFlow and is available at \\url{https:\/\/github.com\/LUMII-Syslab\/QuerySAT}.\n\n\\subsection{Model}\\label{sec:querysat_model}\n\n\\begin{figure*}[ht]\n \\centering\n \\begin{minipage}[t]{0.32\\textwidth}\n \\centering\n \\includegraphics[width=\\textwidth]{figures\/variable_generalization_3sat.pdf}\n \\caption{Part of fully solved 3-SAT instances of the test set depending on the variable count. Models were trained with 32 recurrent steps on formulas with up to 100 variables.}\n \\label{fig:accuracy_on_datasets}\n \\end{minipage}\n \\hfill\n \\begin{minipage}[t]{0.32\\textwidth}\n \\centering\n \\includegraphics[width=\\textwidth]{figures\/generalization_steps_3sat.pdf}\n \\caption{Part of fully solved 3-SAT instances with 400 variables depending on the steps taken at test time. Models were trained with 32 recurrent steps on formulas with up to 100 variables.}\n \\label{fig:accuracy_on_steps}\n \\end{minipage}\n \\hfill\n \\begin{minipage}[t]{0.32\\textwidth}\n \\centering\n \\includegraphics[width=\\textwidth]{figures\/neurosat_vs_querysat_cactus.pdf}\n \\caption{Time vs fully solved instances trained with 32 recurrent steps on 3-SAT instances with 5-100 variables and then tested on instances with 5-405 variables.}\n \\label{fig:neural_solver_cactus}\n \\end{minipage}\n\\end{figure*}\n\n\nQuerySAT is based on a GNN employing the proposed query mechanism and unsupervised SAT loss function. It receives a CNF Boolean formula $\\phi$ in the input represented as two adjacency matrices of variables-clauses graphs - $A_{p} \\in \\{0,1\\}^{n \\times m}$ and $A_{n} \\in \\{0,1\\}^{n \\times m}$, where $n$ and $m$ are the number of variables and clauses, respectively. \n$A_{p}$ represents all positive variable mentions in the clauses, and $A_{n}$ represents all negated variable mentions. The network outputs a vector $out \\in [0,1]^n \\subseteq \\mathbb{R}^n$ -- a variable assignment. \nQuerySAT works in a step-wise manner with recurrent application of the following graph-based recurrent unit: \n\\begin{align*}\\label{formula:querysat}\n \n q_i &= \\sigma (\\text{MLP}_q(v_i,\\; t)) \\\\\n e_i &= \\mathcal{V_{\\phi}}(q_i) \\\\\n c_{i+1} &= \\text{PairNorm}(\\text{MLP}_c(c_i,\\; e_i)) \\\\ \n v_{i+1} &= \\text{PairNorm}(\\text{MLP}_v(v_i,\\; A_{p} c_{i+1},\\; A_{n} c_{i+1},\\; \\nabla_{q_i}e_i)) \\\\\n out &= \\sigma (\\text{MLP}_o(v_{i+1})).\n\\end{align*}\n\nAt the beginning an empty state vector initialized with all ones is allocated for each variable and each clause and the unit is recurrently applied to them $s_{train}$ steps at training and $s_{test}$ steps at evaluation. At each step $i$ starting from $i=0$, QuerySAT produces a query $q_i \\in \\{0,1\\}^{n \\times d}$, where $d$ is the feature map count, by applying a 2-layer multi-layer perceptron $\\text{MLP}_q$ to the variables state $v_i \\in \\mathbb{R}^{n \\times d}$. A random noise vector $t \\in \\mathbb{R}^{n \\times r}$ is also given in the input to this MLP. The query is range-limited by the sigmoid function $\\sigma$ and evaluated by our unsupervised loss. Here we employ per-clause loss defined as $\\mathcal{V}_{\\phi}(q_i)=\\{V_c(q_i))\\; | \\; c \\in \\phi\\}$ which returns the vector of clause losses $e_i$. \n\nWe then obtain the new clauses state $c_{i+1} \\in \\mathbb{R}^{m \\times d}$ by applying a 2-layer perceptron $\\text{MLP}_c$ and PairNorm \\cite{zhao2019pairnorm} to the previous clauses state $c_i$ and the query result $e_i$. Query mechanism thus replaces a variables-to-clauses message passing step that would be present in a classical message-passing architecture.\nThen, information aggregation from clauses-to-variables is performed by message-passing, which for some variable sums all the clause states in which the variable occurs. Positive and negative occurrences are treated separately and are implemented as a sparse matrix multiplication between the clause state $c_{i+1}$ and occurrence matrices $A_{p}$ and $A_{n}$.\n\nThe new variables state $v_{i+1} \\in \\mathbb{R}^{n \\times d}$ is then obtained by applying a 3-layer perceptron $\\text{MLP}_v$ and PairNorm to the variables state $v_i$, aggregated clause messages and the query loss gradient with the respect to query $q_i$. The new variables state is then mapped to the output variable assignment $out \\in [0,1]^n$ using another 2-layer perceptron $\\text{MLP}_o$ and sigmoid function $\\sigma$. In all MLP layers we use LeakyReLU activation. \n\nThe model is trained using the loss function $\\mathcal{L}_{\\phi}^{\\log}$ proposed in section \\ref{sec:sat_loss}. The loss is calculated from variable assignments $out$ at each step, and the sum of all losses is minimized. Using the loss at each time-step has shown performance improvements \\cite{palm2017recurrent, amizadeh2019pdp, ozolins2020matrix} versus a single loss calculation at the end. Also, it enables using many more steps in evaluation than in training. Without impacting the model's performance, we use a conditional exit at any step if the variable assignment satisfies the input formula.\n\nAlthough giving an additional noise parameter $t$ to $\\text{MLP}_q$ seems like a minor detail, it serves several purposes. First, it avoids a blowup in the gradient during training, which may result from applying normalization to zero values. Secondly, it enables the model to learn a randomized algorithm by providing greater diversity between the queries, which is especially needed for obtaining good accuracy in testing with many steps. Thirdly, it allows the model to perform differently in several evaluations (we observed that in the experiments), mimicking restarts in the classical SAT solvers. We find that for our purposes $t \\in \\mathcal{N}(0,1)^{n \\times r}$, where $r=4$, works well.\n\n\\subsection{Training tricks}\\label{sec:tricks}\nIn this section, we describe the model architecture and training tricks that are not per se important for the power of the proposed architecture yet slightly improves model performance.\n\n\\textbf{Gradient scaling.} As we calculate loss and minimize it at each step, large gradient values may accumulate in the backwards pass for the first-time steps. Such uneven gradient distribution slows down training, and to mitigate that, we introduce gradient scaling and apply it to the variables and clauses states. It works by downscaling the gradient in the backward pass at each time-step as follows: $stop\\_gradient(x)\\alpha + x(1-\\alpha)$, where $x$ is value and $\\alpha \\in [0,1]$ is a hyperparameter. We find that for QuerySAT with 32 recurrent steps, $\\alpha = 0.2$ works the best.\n\n\\textbf{Multi-assignment loss.} Instead of returning a single variable assignment, we let the model return several $out \\in [0,1]^{n \\times u}$, where $u \\in \\mathbb{Z}^+$ is hyperparameter. The $u=8$ works best for us. For each assignment, we calculate loss value using the proposed loss function $\\mathcal{L}_{\\phi}^{\\log}$ and then obtain the final step loss as a weighted sum of assignment losses. Weighting is done as follows - we sort assignments losses in descending order and enumerate them from $1$ to $u$. Let's call each such number the loss index. Then we calculate the final loss as the sum of each loss value multiplied with its squared index and then dividing the sum with the sum of squared indices. From our observations, such multi-assignment loss only gives improvements when used with unsupervised loss as the model in the early stages of training is free to explore various outputs. The assignment with the lowest loss value (before weighting) is promoted as the final model assignment. \n\n\\subsection{Evaluation}\\label{sec:querysat_evaluation}\n\n\\begin{table*}[ht]\n\\centering\n\\caption{Mean test accuracy (higher is better) as per cent of fully solved instances from the test set over 3 consecutive runs. Both models were trained with 32 recurrent steps for 500k training iterations and then tested with 32, 512, and 4096 recurrent steps. Value after $\\pm$ indicates the standard error.}\n\\label{tab:sat_tasks_results}\n\\begin{adjustbox}{max width=\\textwidth}\n\\begin{tabular}{lcccccc} \n\\toprule\n\\multirow{2}{*}{Task} & \\multicolumn{3}{c}{QuerySAT} & \\multicolumn{3}{c}{NeuroCore} \\\\ \n\\cmidrule(lr){2-4}\\cmidrule(lr){5-7}\n & $s_{test} = 32$ & $s_{test} = 512 $ & $s_{test}=4096$ & $s_{test} =32$ &$s_{test} = 512$ & $s_{test} = 4096$ \\\\ \n\\toprule\nk-SAT &$72.12 \\pm 0.19$ &$96.61 \\pm 0.78$ &$\\boldsymbol{99.05} \\pm 0.38$ &$21.64 \\pm 0.27$ &$46.85 \\pm 5.02$ &$50.82 \\pm 6.41$ \\\\ \n\\midrule\n3-SAT &$61.89 \\pm 5.19$ &$88.20 \\pm 4.01$ &$\\boldsymbol{93.32} \\pm 3.21$ &$28.38 \\pm 3.24$ &$53.49 \\pm 3.94$ &$57.63 \\pm 4.38$ \\\\ \n\\midrule\n3-Clique &$82.00 \\pm 4.73$ &$93.06 \\pm 4.67$ &$\\boldsymbol{94.74} \\pm 4.62$ &$1.03 \\pm 0.69$ &$1.03 \\pm 0.66$ &$1.04 \\pm 0.66$ \\\\ \n\\midrule\nk-Coloring &$91.70 \\pm 1.01$ &$97.76 \\pm 0.98$ &$\\boldsymbol{98.32} \\pm 0.82$ &$0.0 \\pm 0.0$ &$0.0 \\pm 0.0$ &$0.0 \\pm 0.0$ \\\\ \n\\midrule\nSHA-1 &$33.25 \\pm 4.17$ &$\\boldsymbol{46.57} \\pm 1.16$ &$46.45 \\pm 1.10$ &$0.00 \\pm 0.0$ &$0.27\\pm 0.09$ &$0.24 \\pm 0.09$ \\\\\n\\bottomrule\n\\end{tabular}\n\\end{adjustbox}\n\\end{table*}\n\nWe evaluate the QuerySAT model on several SAT tasks, which are standard benchmarks for classical SAT solvers, and the majority of them are not an easy feat for neural solvers. Namely, we chose k-SAT, 3-SAT, and also 3-Clique, k-Coloring, and SHA-1 preimage attack problems represented as CNF Boolean formulas. All datasets consists only of satisfiable formulas. The QuerySAT architecture is compared to the derivative of NeuroSAT \\cite{selsam2018learning} that was used for predicting unsatisfiable cores by \\cite{selsam2019guiding} but is generally applicable to any variables-wise predictions on SAT factor graph. Further on, we refer to this NeuroSAT derivative as NeuroCore. The results are also compared to the GSAT and Glucose 4 classical solvers. For all tasks, we generate a train set of 100k formulas and validation and test sets of 10k formulas each. The 3-SAT and SHA-1 tasks are an exception as their test sets consist of 40k and 5k formulas, respectively. For most tasks, we use larger formulas in the test set to evaluate the generalization capability of the model.\n\nThe k-SAT task is taken from \\cite{selsam2018learning} -- each clause in the $n$ variable formula is generated by sampling a small integer $k$ (size of the clause), and then randomly without replacement taking $k$ of the $n$ variables. Each variable in the clause is then negated with 50\\% probability. The resulting clauses consist of roughly four variables on average. The train and validation sets consist of formulas with 3 to 100 variables, but the test set -- with 3 to 200 variables.\n\nGraph and 3-SAT tasks are generated using the CNFGen library \\cite{lauria2017cnfgen}, which allows encoding several popular problems as SAT instances. We generate hard 3-SAT instances at the satisfiability boundary where the relationship between the number of clauses ($m$) and variables ($n$) is $m = 4.258n + 58.26n^{-\\frac{2}{3}}$ \\cite{crawford1996experimental}. The train and validation sets of 3-SAT tasks consist of formulas with 5 to 100 variables, but test set -- with 5 to 405 variables. \n\nFor the graph-based tasks, we generate Erd\u0151s\u2013R\u00e9nyi graphs with edge probability $p$. For the 3-Clique task, where the main goal is to find all triangles in the graph, we use $p=3 ^ \\frac{1}{3} \/ (v(2 - 3v + v^2))^ {\\frac{1}{3}}$, where $v$ is the vertex count in the graph. For the k-Coloring task we use $p = \\frac{(1 + 0.2) \\ln v}{v} + 0.05$ and the goal is to color the graph with at least $k$ colours. Such generation produces mostly sparse connected graphs that are colourable using 3 to 5 colours. The graphs are then encoded as SAT instances using the CNFGen library. For both tasks, train and validation sets consist of graphs with 4 to 40 vertices, but the test set -- with 4 to 100 vertices.\n\nWe also experiment with the SHA-1 preimage attack task from the SAT Race 2019 \\cite{skladanivskyy_minimalistic_2019}. The goal of this task is to find the message value given the hash value. The original SAT competition task uses the SHA-1 algorithm with 17 rounds and asks the solver to find the first up to 160 message bits. Such configuration produces SAT instances at the threshold of satisfiability, and it is expected that only a single solution exists. That is a challenging task even for modern SAT solvers; hence, we use a smaller set-up to find the first 2 to 20 bits of the message. But even then, it is still a moderately hard task. We generate instances for this task using the CGen tool \\cite{skladanivskyy2020tailored}.\n\nTo make the QuerySAT and NeuroCore architectures comparable, we train NeuroCore with the same per-step loss and apply the same bag of tricks that we used for QuerySAT, as described in Sections \\ref{sec:querysat_model} and \\ref{sec:tricks}. NeuroCore is also given a chance to return results in any step if the correct solution has been found. The rest of the architecture is left intact.\nFor both models, we use 128 feature maps that produce similarly sized models and train them with a batch size of 20000 nodes (max node count in the input factor graph), 32 recurrent steps and a learning rate $2 \\times 10^{-4}$. Hyperparameters are selected by performing a grid search by hand. On the SHA-1 preimage attack, models are trained for 1M iterations, but on the rest of the tasks for 500k iterations. Afterwards, models are tested with 32, 512, and 4096 recurrent steps and the mean accuracy as a percentage of fully solved instances over 3 consecutive runs is depicted in Table \\ref{tab:sat_tasks_results}. We see that QuerySAT with 4096 steps performs consistently the best for all the tasks. It can solve 3-Clique, k-Coloring and SHA-1 tasks on which NeuroCore produces an accuracy of almost zero.\n\nDetailed comparison of both architectures is conducted on the 3-SAT task, where we check step-wise and formula-wise generalization of both models. To evaluate both properties, we use models trained with 32 recurrent steps on formulas with 3 to 100 variables. Fig. \\ref{fig:accuracy_on_datasets} depicts generalization to harder formulas (up to 400 variables) with 32 and 4096 recurrent steps at the test time. Fig. \\ref{fig:accuracy_on_steps} shows step-wise generalization by changing the number of steps $s_{test}$ from 4 up to 131k when testing on the same formulas with 400 variables. QuerySAT outperforms NeuroCore on both generalization tasks by a wide margin and, with 32 steps, has a similar performance to the NeuroCore with 4096 steps. QuerySAT's performance increases with the step count at the test time, although it is trained only with 32 steps. NeuroCore, on the contrary, plateaus at approximately 16k steps and 40\\% accuracy. In Fig. \\ref{fig:neural_solver_cactus}, the cactus plot is shown comparing NeuroCore, and QuerySAT tested with 4096 recurrent steps on 3-SAT instances with 5 to 405 variables. Cactus plot shows cumulative time spend for solving instances (y-axis) versus cumulative solved instance count (x-axis). More solved instances in less time indicate better model performance. At the same time interval, QuerySAT solves $\\sim 36k$ formulas, while NeuroCore solves only $\\sim 18k$ from the total of 40k 3-SAT formulas in the test set (see Fig. \\ref{fig:neural_solver_cactus}).\n\n\\begin{figure*}[ht]\n \\centering\n \\begin{minipage}[t]{0.48\\textwidth}\n \\centering\n \\includegraphics[width=0.77\\textwidth]{figures\/sha1_querysat_vs_gsat_vs_glucose.pdf}\n \\caption{Cactus plot representing the number of solved SHA-1 preimage attack instances vs. used time for QuerySAT, GSAT and Glucose 4 solvers.}\n \\label{fig:sha1_query_vs_gsat}\n \\end{minipage}\n \\hfill\n \\begin{minipage}[t]{0.48\\textwidth}\n \\centering\n \\includegraphics[width=0.77\\textwidth]{figures\/3sat_querysat_vs_classical.pdf}\n \\caption{Cactus plot representing the number of solved 3-SAT instances vs. used time for QuerySAT, GSAT and Glucose 4 solvers.}\n \\label{fig:scatter_query_vs_gsat}\n \\end{minipage}\n \\begin{minipage}[t]{0.48\\textwidth}\n \\centering\n \\includegraphics[width=0.77\\textwidth]{figures\/training_k_sat.pdf}\n \\caption{Error as per cent of unsolved instances in the validation set depending on training iteration for the k-SAT task. The validation set consists of 10k formulas with 3-100 variables. $s_{test}=64$ is used for validation.}\n \\label{fig:training_k-sat}\n \\end{minipage}\n \\hfill\n \\begin{minipage}[t]{0.48\\textwidth}\n \\centering\n \\includegraphics[width=0.77\\textwidth]{figures\/training_clique.pdf}\n \\caption{Error as per cent of unsolved instances in the validation set depending on training iteration for the 3-Clique task. The validation set consists of 10k graphs with 4-20 vertices. $s_{test}=64$ is used for validation.}\n \\label{fig:training_k-clique}\n \\end{minipage}\n\\end{figure*}\n\nTo show current capabilities of QuerySAT, we also compare it to GSAT \\cite{selman1992a} and Glucose 4 \\cite{audemard2018glucose,een2003extensible} classical solvers. GSAT is a widely known incomplete local-search solver, and Glucose is a contemporary conflict-driven clause learning solver. As the QuerySAT is also an incomplete solver, it's directly comparable to the GSAT algorithm. Comparison to Glucose is not on equal ground since Glucose can certify that some instance is UNSAT while QuerySAT runs indefinitely for such instance. That said, we present their comparison nonetheless to give a rough idea of their relative performance. Note that both - GSAT and Glucose 4 - may not accurately represent the performance of modern state-of-the-art SAT solvers. All solvers are evaluated on the same hardware but note that QuerySAT utilizes a single Nvidia T4 GPU while the others do not. We configure all three solvers to have approximately a 2-second time limit by giving QuerySAT 1024 recurrent steps, GSAT -- 500k steps, and setting a 2-second timeout for Glucose solver. We compare these solvers on previously described test sets of 3-SAT and SHA-1 tasks. The QuerySAT was trained on the forenamed train sets. The results are depicted as cactus plots in the Fig. \\ref{fig:sha1_query_vs_gsat} and \\ref{fig:scatter_query_vs_gsat}. Glucose utilizes the structure of the problem and therefore struggles on random 3-SAT instances yet solves the SHA-1 task in almost no time. Even though GSAT outperforms QuerySAT on 3-SAT instances, QuerySAT seems to utilize the SHA-1 structure and achieves better performance than GSAT. QuerySAT, similarly to other contemporary end-to-end neural solvers \\cite{selsam2018learning, amizadeh2019pdp}, in the general case, is outperformed by classical solvers and requires breakthroughs to allow scaling them to large industrial instances.\n\n\n\n\n\\section{Evaluating the query mechanism}\\label{sec:eval_query_mechanism}\n\n\n\\begin{table*}[ht]\n\\centering\n\\caption{Mean accuracy of NeuroCore and its augmentation with query and gradient and as per cent of fully solved instances from the test set over 3 consecutive runs. All models are trained with 32 recurrent steps and evaluated with 32, 512, and 4096 steps. Value after $\\pm$ indicates the standard error. Note that the training, validation and test instances are smaller than in Table \\ref{tab:sat_tasks_results}. \n}\n\\label{tab:query_generalization}\n\\begin{adjustbox}{max width=\\textwidth}\n\\begin{tabular}{lcccccc} \n\\toprule\n\\multirow{2}{*}{} & \\multicolumn{3}{c}{k-SAT } & \\multicolumn{3}{c}{3-Clique} \\\\ \n\\cmidrule(lr){2-4}\\cmidrule(lr){5-7}\n & \\multicolumn{1}{c}{$s_{test} = 32$} & \\multicolumn{1}{c}{$s_{test} = 512$} & \\multicolumn{1}{c}{$s_{test} = 4096$} & \\multicolumn{1}{c}{$s_{test} = 32$} & \\multicolumn{1}{c}{$s_{test} = 512$} & \\multicolumn{1}{c}{$s_{test} = 4096$} \\\\ \n\\toprule\n\\textit{NeuroCore}\n&\\begin{tabular}{@{}c@{}}$38.86 \\pm 1.04$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$56.14 \\pm 3.48$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$60.97 \\pm 6.19$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$65.38 \\pm 5.09$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$67.12 \\pm 3.25$\\end{tabular} \n&\\begin{tabular}{@{}c@{}}$70.39 \\pm 2.68$\\end{tabular} \\\\\n\\midrule\n\\textit{\\, + Query}\n&\\begin{tabular}{@{}c@{}}$41.09 \\pm 2.33$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$61.21 \\pm 9.89$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$67.21 \\pm 13.54$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$62 .90 \\pm 2.55$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$78.19 \\pm 2.67$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$84.55 \\pm 2.93$\\end{tabular} \\\\\n\\midrule\n\\textit{\\, + Query + G}\n&\\begin{tabular}{@{}c@{}}$\\boldsymbol{47.92 \\pm 1.66}$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$\\boldsymbol{71.95 \\pm 6.59}$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$\\boldsymbol{75.50 \\pm 7.15}$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$\\boldsymbol{86.13 \\pm 4.83}$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$\\boldsymbol{93.96\\pm 2.59}$\\end{tabular}\n&\\begin{tabular}{@{}c@{}}$\\boldsymbol{95.50\\pm 1.95}$\\end{tabular} \\\\\n\\bottomrule\n\\end{tabular}\n\\end{adjustbox}\n\\end{table*}\n\n\nWe evaluate the impact of the query mechanism by augmenting NeuroCore architecture with it and measuring the improvement. It is straightforward to do since NeuroCore is similar to QuerySAT. NeuroCore uses literals-to-clauses and clauses-to-literals message-passing to update the internal states for literals and clauses. Therefore, we add a query mechanism alongside the literals-to-clauses message-passing and also give the gradient of the evaluated query to the MLP that updates the literals state. We experiment with two variants of the query mechanism: with both query and gradient (\\textit{+ Query + G}) and only with query (\\textit{+ Query}). Both versions are compared with the standard NeuroCore architecture (\\textit{NeuroCore}).\n\nWe chose k-SAT and 3-Clique tasks (see section \\ref{sec:querysat_evaluation}) for evaluation. We use the same dataset split as previously. For the k-SAT task, we use the same train and validation set as previously, but for the test set generate formulas with 100 to 200 variables. On the same note, train and validation sets for the 3-Clique task consists of graphs with 4 to 20 vertices but the test set of 20 to 40 vertices. The test set consists of harder formulas to see how various variants of query mechanism impacts generalization. \n\n\nAll three versions are trained for 100k iterations with the same training and network configuration as described in Section \\ref{sec:querysat_evaluation}. The trained model is then tested with 32, 512, and 4096 recurrent steps. The mean results of 3 consecutive runs are depicted in Table \\ref{tab:query_generalization}. We also include validation error in the train time for each version using 64 recurrent steps ($s_{test}$), it is depicted in the Fig. \\ref{fig:training_k-sat} and \\ref{fig:training_k-clique}.\n\nIn Table \\ref{tab:query_generalization}, we can see that both versions with a query mechanism outperform the NeuroCore baseline. The version with a query mechanism and gradient trains faster and achieves better accuracy. Interestingly, adding a gradient significantly improves the model's accuracy of the 3-Clique task. \nWe reason that gradient is helpful for the tasks that represented as a SAT instance has a distinct structure. Such structure is very expressive for the 3-Clique task, yet, on the contrary, k-SAT doesn't have any distinct structure as it is sampled from a uniform distribution \\cite{yolcu2019learning}. \n\n\n\n\\section{Related work}\nNeural networks have been proposed as an effective alternative for automatically developing heuristic algorithms for NP-hard problems \\cite{bengio2020machine, cappart2021combinatorial}. Two main research directions are replacing handcrafted heuristics with a neural network in a classical solver and building end-to-end neural solvers.\n\n\\cite{selsam2018learning} proposed a Graph Neural Network (GNN) architecture called NeuroSAT that is trained using single-bit supervision to predict whether the Boolean formula is satisfiable or not. The variable assignment is obtained from the last layer by performing 2-clustering on the output, but the network is never optimized for producing the assignment explicitly. Hence, it is not clear why such an approach should work. In a later work, \\cite{selsam2019guiding} simplified NeuroSAT architecture (they call it NeuroCore) and used it for guiding high-performance solvers (e.g., MiniSAT, Glucose) by predicting how likely each variable is in the unsatisfiable core. \\cite{kurin2019improving} use Q-learning to train similar GNN for predicting branching heuristics in MiniSAT solver. Others \\cite{azar2020nngsat, kurin2019improving} similarly augment classical solvers with neural networks to solve 2-Quantified Boolean Formulas and logic locked circuits. On the same trend, \\cite{yolcu2019learning} trained a GNN using reinforcement learning to learn a local-search heuristic that finds a solution similar to the GSAT \/ WalkSAT algorithm by flipping a single variable in each step. It is also theoretically shown that GNN can learn to mimic the WalkSAT algorithm \\cite{chen2019graph}. \n\nSupervised and reinforcement learning are not the best options for solving SAT, as the variable assignment is time-consuming to obtain, many possible solutions exist, and training may be slow. \\cite{amizadeh2018learning} proposed a differentiable unsupervised approach for directly solving Circuit-SAT. For this purpose, they represent variables as real numbers in the range $[0,1]$ and use GNN for producing variable predictions on the Circuit-SAT graph. The predicted variables are evaluated as Circuit-SAT, but the AND and OR functions are substituted with differentiable Softmax and Softmin functions. In a later work \\cite{amizadeh2019pdp}, they have applied the same method for solving SAT problems. To encourage exploration, they introduce a temperature parameter that is annealed in the training time, making training cumbersome. Although the training method is closely related to ours, their loss function is more complex. \\cite{kyrillidis2020fouriersat} lately proposed to represent Boolean functions by multilinear polynomials and find a solution to a single formula by using gradient descent optimization. Even though their loss function is similar to ours, they solve the MAX-SAT problem and do not directly optimize towards finding a solution to the SAT problem.\n\n\n\\section{Conclusions}\nIn this paper, we have proposed a query mechanism that allows the neural network to make several solution trials, obtain the loss of each trial and change its strategy accordingly. To evaluate the impact of the query mechanism, we propose an unsupervised SAT loss and integrate it with queries to form the QuerySAT architecture. We find that QuerySAT outperforms the message-passing neural baseline on all proposed tasks: k-SAT, 3-SAT, 3-Clique, k-Coloring, and SHA-1 preimage attack. Experiments show that query mechanism can significantly increase the performance of message-passing graph neural networks. To give a better insight into the current capabilities of QuerySAT, we also compare it with classical solvers on 3-SAT and SHA-1 preimage attack tasks. Although we have analyzed only SAT solving, we expect a similar benefit from the query mechanism on other neural architectures and tasks. Since QuerySAT employs unsupervised loss, not requiring to know the labels, it provides rich opportunities for integration with classical solvers.\n\n\\section*{Acknowledgment}\n\n We would like to thank the IMCS UL Scientific Cloud for the computing power and Leo Truk\u0161\u0101ns for the technical support. This research is supported by Google Cloud and funded by the Latvian Council of Science, projects No.~lzp\\nobreakdash-2018\/1\\nobreakdash-0327, lzp-2021\/1-0479.\n\n\\bibliographystyle{IEEEtran}\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Motivations and outline}\nGibbs measures are familiar objects in various areas of applied\nprobability. Originally they have been devised in the framework of \nlattice spin systems to describe thermodynamic equilibrium states. \nObtaining these probability measures involves two basic steps. \nFirst a family of Gibbs measures for a finite number of random\nvariables is constructed as a modification of a reference measure \nusually describing the independent random field. The density is the\nexponential of an additive functional dependent on an interaction \nfunction. Then one takes a weak limit of these measures by increasing \nthe number of variables to infinity. This procedure copes with the\nfact that in usual model systems the interaction diverges in this\nlimit, thus the limit measure cannot be directly defined. \n\nThe context in which the class of Gibbs probability measures proved \nto be useful has substantially widened over the years, reaching the \nrealm of Brownian motion. \nOne natural way leading to Gibbs measures on path space is the \napplication of the Feynman-Kac formula and its extensions. Suppose \n$H$ is a self-adjoint operator acting on a Hilbert space $\\cal H$, and \nlet $f \\in {\\cal H}$. Then for a variety of specific choices of $H$ \nan equality of the type \n\\begin{equation}\n(e^{-tH}f)(x) = \\int_{\\mathbb{R}^d} dy \\int f(X_t) d\\mu_{[0,t]}(X|x,y)\n\\label{fk}\n\\end{equation}\ncan be derived, where $\\mathbb{R} \\ni t\\mapsto X_t \\in \\mathbb{R}^d$ is Brownian \nmotion and\n\\begin{equation}\nd\\mu_{[0,t]}(X|x,y) = e^{-U_t(X|x,y)} d\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x,y}_{[0,t]}(X).\n\\label{gibb}\n\\end{equation}\nHere $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x,y}_{[0,t]}$ is Brownian bridge over the bounded time \ninterval $[0,t]$, starting at $X_0 = x$ and ending at $X_t = y$, and \n$U_t$ is a functional of Brownian paths derived from $H$. After \nnormalizing to 1, the measure $\\mu_{[0,t]}(X|x,y)$ can be viewn as a \nGibbs measure on path space for ``finite volume'' $[0,t]$, \n``interaction'' $U$, ``boundary condition'' $x,y$, and ``reference \nmeasure'' $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x,y}$. \n\nIn this paper we are interested in the question whether an extension \nin a suitable sense over the full time line $\\mathbb{R}$ of $\\mu_{[0,t]}$ \nexists. The details of the proof strongly depend on the choice of $U$, \nwhich for our purposes here will be specified below. These choices are \nmotivated by particular applications covered by the following classes\nof models, of which we talk only in this introduction. \n\n\\medskip\n\\vspace{0.1cm}\n\\noindent\nI. Densities dependent on the local time of Brownian motion\n\\begin{enumerate}\n\\item[(1)]\n{\\it $P(\\phi)_1$-process (It\\^o diffusion)} \\quad \nThis includes the familiar case of Schr\\\"odinger operators $H = \n(-1\/2)\\Delta + V$, ${\\cal H} = C_0^\\infty(\\mathbb{R}^d)$, with a potential \n$V(x)$ that can be chosen fairly generally (Kato-class). The \nwell-known result is \n$$\nU_t(X) = \\int_0^t V(X_s) ds.\n$$\n$\\mu$ describes in this case the path measure of the Markov process \ngiven by \n$$\ndX_t = dB_t + (\\nabla\\log\\Psi)(X_t)dt,\n$$\nwhere $B_t$ is $\\mathbb{R}^d$-valued Brownian motion and $\\Psi$ is the\neigenfunction of $H$ lying at the bottom of its spectrum (ground\nstate). It\\^o diffusions can be defined also on infinite dimensional \nspaces related to other SPDE. For further details on these Gibbs \nmeasures for bounded intervals and their extensions to $\\mathbb{R}$ we refer \nto \\cite{Sim75,Sim82,BL}.\n\n\n\\item[(2)] \n{\\it Nelson's model} \\quad \nThis is a scalar quantum field model describing the interaction \nof an electrically charged spinless particle with a boson field. In\nthis case $H$ is written as the sum of the free particle Hamiltonian\n$(-1\/2)\\Delta + V$, the free field Hamiltonian $\\int |k|a(k)^*a(k)dk$\nwith the usual boson creation and annihilation operators $a^*$ and $a$,\nand the interaction Hamiltonian $\\int(\\widehat\\rho(k)\/{\\sqrt{2|k|}})\n(e^{ik\\cdot x}a(k) + e^{-ik\\cdot x}a^*(k))dk$, with charge\ndistribution function $\\rho$. Moreover, ${\\cal H} = L^2(\\mathbb{R}^d,dx)\n\\otimes \\mathop{\\mathrm{{\\cal F}}}\\nolimits$, where $\\mathop{\\mathrm{{\\cal F}}}\\nolimits$ is Fock space, and a Feynman-Kac-type formula \nas (\\ref{fk}) above can be obtained by mapping $\\cal H$ into a space \nof continuous functions through a joint use of the so called ground \nstate transform and Wiener-It\\^o isomorphism yielding\n$$\nU_t(X) = \\int_0^t V(X_s) ds + \\int_0^t \\int_0^t W^\\rho(X_s-X_r,s-r) \ndsdr,\n$$\nwith \n$$\nW^\\rho(x,s) = -\\frac{1}{4}\\int_{\\mathbb{R}^d} \\frac{|\\widehat \\rho(k)|^2}{|k|} \ne^{-i k\\cdot x - |k||s|} dk.\n$$\nExtending this Gibbs measure from $[0,t]$ to $\\mathbb{R}$ \nis in this case of special interest since it allows a direct \nexpression of the ground state of $H$ in terms of its Radon-Nikodym \nderivative with respect to an underlying product measure, which makes \na rigorous derivation and proof of ground state properties possible. \nFor the case of translation invariant models (involving $V \\equiv 0$) \nthere are only few results available. For details see \n\\cite{LM,LMS1,LMS2,BHLMS}. \n\n\\item[(3)]\n{\\it Polaron and bipolaron models}\\quad\nThe polaron is a ``dressed'' electron (i.e., embedded into an energy\ncloud) interacting with a phonon field (i.e., quantum particles\ncarrying the vibrational energy of an ionic crystal). In this case we\nhave similar operators acting on the same Hilbert space as above \nexcept that the dispersion relation $|k|$ is replaced by 1 in the\nfree field Hamiltonian, and $\\widehat\\rho\/\\sqrt{2|k|}$ in the \ninteraction term by $1\/|k|$. This leads to the same $U_t$ as in the \ncase of Nelson's model with \n$$\nW^{\\rm\\tiny{pol}}(x,s) = -\\frac{1}{4|x|}\\,e^{-|s|}.\n$$\nThe bipolaron differs by the fact that it consists of two dressed\nelectrons coupled to the same phonon field, which are repelling each\nother by Coulomb interaction. In this case \n$$\nU_t(X) = \\alpha^2 \\int_0^t\\int_0^t {\\cal E}^W(X_s,Y_r,s-r) dsdr\n- g\\int_0^t \\frac{ds}{|X_s-Y_s|},\n$$\nwhere \n$$\n{\\cal E}^W(X_s,Y_r,u) = W(X_s-X_r,u) + 2W(X_s-Y_r,u) + W(Y_s-Y_r,u)\n$$\nwith $W = W^{\\rm\\tiny{pol}}$, $\\alpha < 0$ being the polaron-phonon\ncoupling paramater and $g>0$ the strength of the Coulomb repulsion \nbetween the two polarons. Moreover, in this case the reference measure\nis a product of two independent Wiener measures. For literature see\n\\cite{DV,S,LM06}.\n\n\\item[(4)]\n{\\it Intersection local time (weakly self-avoiding polymer)}\\quad\nFormally, the densities are given by\n$$ \nU_t(X) = \\int_0^t \\int_0^t \\delta(X_s-X_r) ds dr, \n$$ \nmeant to describe a polymer model with short-range ``soft-core'' \ninteraction encouraging to avoid self-intersections. For $d=2, 3$ \nsee \\cite{MR853758,MR667754,MR609228,MR573702,MR1240717}. In \n\\cite{Symanzik:1969lr} it was proved that in $d=2$ the model can be \nrigorously defined after an additive renormalization and the so \nobtained measure is absolutely continuous with respect to Wiener measure \nin 2 dimensions. In $d=3$ the singularity of the energy $U_t$ is more \nserious but an additive renormalization still suffices; Westwater proved \nexistence of the Gibbs measure which, however, in this case is not \nabsolutely continuous with respect to Wiener measure. Other works \ninclude \\cite{MR729794,MR1329112,MR889758,MR942042}.\n\n\\end{enumerate}\n\n\\vspace{0.1cm}\n\\noindent\nII. Stochastic currents\n\\begin{enumerate}\n\\item[(1)]\n{\\it Nelson's model in point charge limit and Pauli-Fierz model} \\quad\nThe point charge limit of Nelson's model corresponds to the case of \nreplacing $\\rho$ above with a delta-function, while the Pauli-Fierz \nmodel is obtained by replacing the scalar boson field with a quantized\nMaxwell (vector) field. The function in the density in both cases \nformally becomes \n\\begin{equation}\nU_t(X) = \\int_0^t V(X_s) ds + \\int_0^t \\int_0^t W(X_s-X_r,s-r) \ndX_s \\cdot dX_r,\n\\label{doubleito}\n\\end{equation}\nwith a $W=W^\\delta$ and $W=W^{\\rm PF}$ we do not write explicitly \ndown here. The difference from the case above is that instead of \ndouble Riemann we have to deal with double It\\^o integrals. A\nsubstantial difficulty to solve in this case is the proper definition \nof the double integrals in the first place, which will be done below. \nFor the specific model applications see \\cite{GL} in which we perform \nan ultraviolet renormalization of Nelson's Hamiltonian by using path\nmeasures whose densities are of the above form, and \\cite{H,HL2} for the\nPauli-Fierz model. The similar point charge limit in the latter model \nis an open problem.\n\n\\item[(2)]\n{\\it Turbulent fluids} \\quad\nIn fluid dynamics the understanding prevails that fully developed \nturbulence should be described by a suitable measure over\ndivergence-free velocity fields $u(x)$. One way of modelling it starts \nfrom the assumption that the vorticity field $\\nabla \\wedge u(x)$ is \nconcentrated along Brownian curves $X_t \\in \\mathbb{R}^3$. Under the \nEulerian incompressible flow, the kinetic energy $(1\/2) \\int u(x)^2 \ndx$ is conserved. The formal expression of the total energy is\n\\begin{equation}\\label{3.11}\nU_t(X)= \\int_0^t \\int_0^t \\frac{1}{|X_t - X_s|} \\; dX_t \\cdot dX_s.\n\\end{equation}\nIn order to have $e^{-U_t(X)}$ as a well-defined random variable at\nall, \\cite{Ffil} imposed the condition that the Coulomb potential is \nmollified so that the fluid has finite static energy. For details we\nrefer to~\\cite{statvort,FGub,Nua,evolution,Ascona}. \n\\end{enumerate}\n\n\n\n\\vspace{0.1cm}\n\\noindent\n{III. Processes with jumps} \\quad \n\\begin{enumerate}\n\\item[]\nApplications include Hamiltonians with spin. Since spin is a discrete \nvariable, this case goes beyond stochastic integration with respect to\nBrownian motion alone. The spin paths $\\sigma_t$ are described via a \nPoisson process $N_t$, and we obtain\n$$\nU_t(X) = \\int_0^t V(X_s) ds + \\int_0^t a(X_t) \\circ dX_t - \n\\int_0^t S(X_t,\\sigma_t) dt +\\int_0^{t^+} \\Phi(X_t,-\\sigma_t) dN_t.\n$$\nFor more details and explicit formulas of the terms above see\n\\cite{Jona,HL1}.\n\\end{enumerate}\n\n\\medskip\nIn this paper we address the purely mathematical problem of existence\nand characterization of Gibbs measures with densities of the type \n(\\ref{doubleito}). An accompanying paper \\cite{GL} takes this further \nto an application to quantum field theory. \n\nThe problem can be formulated in more generality by adding a term \nto $U_t$ in (\\ref{doubleito}) taking into account the dependence on \noutside paths of paths run within bounded time intervals. Choose\nwithout loss a bounded interval in the form $[-T,T] \\subset \\mathbb{R}$ and \nset up Wiener measure $\\mathop{\\mathrm{{\\cal W}}}\\nolimits_T$ on it. Consider the energy functions \ncorresponding to $[-T,T]$ as follows: \n\\begin{eqnarray*}\n&&\nV_T(X) := \\int_{-T}^T V(X_s) ds \\qquad \\mbox{for external \npotential $V$}\\\\\n\\bigskip\n&&\nW_T(X) := \\int_{-T}^T \\int_{-T}^T W(X_t-X_s,t-s) dX_t \\cdot dX_s \n\\qquad \\mbox{internal energy} \\\\ \n\\medskip\\bigskip\n&&\nW_T(X|Y):= W^+_T(X|Y) + W^{-}_T(X|Y) \\qquad \n\\mbox{interaction energy}\n\\end{eqnarray*}\nwhere the interaction energies come from a pair interaction \npotential $W$, and $W^+_T(X|Y) := \\int_T^\\infty dY_t \\int_{-T}^T \ndX_s W(X_s-Y_t,t-s)$ is a term accounting for the interaction of \npaths inside $[-T,T]$ with paths in $[T,\\infty)$, and $W^-_T(X|Y) \n:= \\int_{-\\infty}^{-T} dY_t \\int_{-T}^T dX_s W(X_s-Y_t,t-s)$ for \nexternal paths running in $(-\\infty,-T]$. As in (\\ref{gibb}), \nthese energies give rise to the Gibbs measure for $[-T,T]$\n\\begin{equation}\n\\label{ea:gibbs1}\nd\\mu_T(X) := \\frac{e^{-V_T(X)-\\lambda W_T(X)}}{Z_T}\\, d\\mathop{\\mathrm{{\\cal W}}}\\nolimits_T(X)\n\\end{equation}\nwith a parameter $\\lambda \\in \\mathbb{R}$ tuning the strength of the pair \npotential, and $Z_T$ the normalizing factor turning it into a\nprobability measure. The object of interest are the accumulation \npoints $\\mu$ of the family of measures $\\{\\mu_{T}\\}_{T>0}$ in the \ntopology of local weak convergence. In the DLR \n(Dobrushin-Lanford-Ruelle) approach, these limit points can be \ncharacterized by a property of consistency with respect to a \nprescribed family of probability kernels (specification) providing\nthe local conditional probabilities of the limit random fields. These \nkernels are given by \n\\begin{equation}\n\\label{eq:specc}\n\\mu_T(dX|Y) := \\frac{e^{-V_T(X)-\\lambda W_T(X)-\n\\lambda W_T(X|Y)}}{Z_T(Y)} \\, d\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{Y_{-T},Y_{T}}_T(X)\n\\end{equation}\nwith external (or boundary) path $Y$. The possible accumulation \npoints of the family $\\{\\mu_T\\}_T$ should then satisfy the DLR \nequations\n$\n\\int \\mu_T(A|Y) \\mu(dY) = \\mu(A)\n$ \nfor all cylinder sets $A$ in the sub-$\\sigma$-field generated by \nprojections to $[-T,T]$ for all $T>0$.\n\nIn contrast to the case of interactions depending on the local time \nof the process $X$ (given by double Riemann integrals listed above) \nour case encounters two difficulties.\n\\begin{itemize}\n\\item[(1)] \nThe expressions of $W_T(X)$ and $W_T(X|Y)$ are only formal: \nThe double stochastic integrals are not well defined since the \nintegrands are neither forward nor backward adapted with respect \nto the semimartingale $X$ (Brownian bridge under $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x,y}_T$).\n\\item[(2)] \nThe specification (\\ref{eq:specc}) must be defined pathwise for each \n$Y$, however, in general the only information we have on the boundary \npath $Y$ is that it is a continuous path with a Brownian-like\nregularity. This is insufficient for defining the line integrals with \nrespect to $dY$ appearing in $W_T(X|Y)$.\n\\end{itemize}\n\n\\noindent\nOf these two difficulties the first requires a minor amendment, \nwhile the second is far more serious and will urge us to introduce \nthe novel setting of \\emph{Brownian currents} in which we can make\nsense of a formulation of the DLR equations. Roughly speaking, we \nprovide each sample path with sufficient information in order to \ndetermine the work made by test vector fields. Specifically, we \nconsider the family of random variables $\\{X_t\\}_t$ jointly with \nthe random variables \n$\n C_{st}^X( \\varphi) := \\int_s^t \\varphi(u,X_u) dX_u \n$\n(\\emph{stochastic currents})\nand show that in this augmented sample space the specification can \npathwise be defined. (This problem has some analogy with difficulties \nat defining specifications for unbounded spin systems for which too \none has to select a subset of admissible boundary conditions so that \nthe interaction energy makes sense. Our approach here goes further in \nthat we do not only have to control the growth of the boundary paths \nbut must also provide a priori the value of the line integrals against \na sufficiently large set of test vector fields $\\varphi$.) To perform \nthis ``lifting'' procedure we use techniques of rough paths theory. \n\n\\medskip\nHere is an outline of the paper. In Chapter 2 \nwe present our results on rough paths that will be put at use \nsubsequently. In Chapter 3 we introduce the framework of stochastic \ncurrents, in particular Brownian currents. In Chapter 4 we define\nGibbs measures on these currents for bounded intervals of the line. In \nChapter 5 we turn to proving that these Gibbs measures can be extended \nover the whole line provided the interaction term is weakly coupled. \nThis requirement is needed since, \\emph{faute de mieux}, we use cluster \nexpansion in order to construct weak limits of Gibbs measures for \nbounded intervals. The version of cluster expansion we develop here is \ndifferent from the conventional ones since the configurations are \nsegments of Brownian paths rather than spins of compact or real-valued \nstate space, the interactions depend on double stochastic integrals, \nthe reference measure is not a product measure, and the measure we want \nto construct is non-Markovian. However, it has the same spirit of usual \ncluster expansions as it splits off into a part of hands-on analysis of \nenergy bounds and a part of combinatorics. Beside a proof of \nexistence, cluster expansion allows us to study also uniqueness, typical\npath behaviour and mixing properties of Gibbs measures, which will be\ndone in Chapter 6. \n\n\n\n\\label{sec:gibbs}\n\n\\section{Rough paths}\n\\subsection{Definition of double stochastic integrals by rough paths}\n\\label{sec:rough-paths}\n\nIt\\^o's theory of integration gives a meaning to line integrals of the \nform \n$\n\\int_{-T}^T \\varphi(t,X_t) dX_t,\n$\nwith $X$ a semimartingale (with respect to its natural forward\nfiltration $\\mathcal{F}$) and $\\varphi(t,\\xi)$ some function. The way\nof obtaining it is by taking limits of Riemann sums\n$\n\\sum_\\alpha \\varphi(\\tau_\\alpha,X_{\\tau_\\alpha}) \n(X_{\\tau_{\\alpha+1}}-X_{\\tau_\\alpha})\n$\nover a family of partitions $\\{\\tau_\\alpha\\}_\\alpha$ of $[-T,T]$ with \nmesh decreasing to zero. Such limits are known to exist whenever \n$\\varphi(t,X_t)$ is an $\\mathcal{F}$-adapted functional of $X$, i.e.,\nif it only depends on $\\mathop{\\mathrm{{\\cal F}}}\\nolimits_t$ for any $t\\in[-T,T]$ and has some \nintegrability properties. Then \n$\n\\int_{-T}^T \\varphi(t,X_t) dX_t\n$\nis defined as a random variable on the same probability space on which \n$X$ is defined. In our applications $X$ is the coordinate process on the\nBorel space $\\mathcal{X}$, and the stochastic integral defines a Borel map from\n$\\mathcal{X}$ to $\\mathbb{R}$ which is defined on a set of full $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$ measure.\nIt is important to remember the crucial fact that the full-measure set \nin general depends on the integrand: different integrands may give \ndifferent full-measure sets.\n\nIn this section we analyze the regularity of such integrals by using\nthe theory or rough paths as developed in \n\\cite{lyons-98,lyons-qian-02,MR2091358,feyel}. This theory allows to \nproperly understand stochastic integrals from a purely analytic \nperspective. To gain this freedom we need to enlarge the sample space \nover which the measures are defined, however, this will turn out to \nmake no harm in our applications.\n\nRough paths theory has been devised by T.~Lyons~\\cite{lyons-98} to\ngive a meaning to line integrals of the form\n\\begin{equation}\n \\label{eq:basic-rp-integral}\n\\int_0^T \\varphi(X_t) dX_t, \n\\end{equation}\nin case of $X$ being an irregular function of the parameter. A typical\ncase of interest is when $X$ is chosen to be H\\\"older continuous with \nsmall exponent, such as $\\gamma\\in(0,1)$. \nThe natural approach to defining such integrals is that of taking\nRiemann approximations over a finite partition of $[0,T]$ and proving \nthat the sequence converges as the mesh of the partition goes to zero. \nWhen general integrals in the form $\\int_0^T Y_s dX_s$ are considered\nwith bounded $Y$, this can only hold if $X$ is a process of bounded \nvariation leading to the familiar Stieltjes integral. When $Y$ is \nH\\\"older continuous with exponent $\\rho$, Young's work \\cite{young-36} \nmakes sure that a sufficient condition for the convergence of the \nRiemann sums is that $\\gamma+\\rho > 1$. The Young integral is useful \nin studying, for example, fractional Brownian motion of Hurst index \n$H>1\/2$, which is a stochastic process with H\\\"older continuous paths \nof any exponent $\\gamma < H$. In this case (\\ref{eq:basic-rp-integral}) \ncan be defined for any function $\\varphi$ which is at least $C^1$.\n\nIntegrals of the form (\\ref{eq:basic-rp-integral}) when $X$ is a\nsample of Brownian motion are not within the reach of Young's theory \nand indeed it is not difficult to see that different definitions of \nthe Riemann approximations can lead to different limits (or even not \nconverge at all). This difficulty lies at the basis of the existence \nof more than one type of stochastic integral over Brownian motion, \ntwo well known possibilities being the integrals developed by It\\^o \nand Stratonovich. Lyons's approach to the problem was proving that \nunder reasonable conditions the Riemann sums for $\\int_0^T \\varphi(X_t) \ndX_t$ can be modified by adding an extra term in order to make them\nconverge. This compensation is not unique but in many relevant cases \nit can be performed in such a way that the so obtained integral is an \nextension of the classical line integral and Young integral. That is,\nwhenever $X$ is regular enough, the modified integral coincides with \nthe Young one or Riemann-Lebesgue integral if moreover $X$ is almost \nsurely is differentiable.\n\nWe consider only the case when $X$ is H\\\"older continuous with $\\gamma \n> 1\/3$ since this includes the case of Brownian motion and all our \npresent applications. (The theory for more general $\\gamma$ would be \nfar more cumbersome.) We use the notation $X_{st} = X_t-X_s$, and make \nthe following basic\n\n\\begin{definition}\nFor each bounded $I \\subset \\mathbb{R}$ we call the couple $(X,\\mathbb{X})$ \na \\emph{2-step rough path}, where\n\\begin{enumerate}\n\\item[(1)]\n$X \\in C^\\gamma(I,\\mathbb{R}^d)$;\n\n\\item[(2)\nthere exists a function $\\mathbb{X} : I\\times I \\to \\mathbb{R}^{d}\\times \\mathbb{R}^d$ \nsatisfying the \\emph{multiplicative property}\n \\begin{equation}\n \\label{eq:mult}\n \\mathbb{X}^{ij}_{st} - \\mathbb{X}^{ij}_{su} - \\mathbb{X}^{ij}_{ut} = \n X^i_{su} X^j_{ut} \n \\end{equation}\nfor all $s \\le u \\le t \\in I$ and any $i,j =1,\\dots,d$;\n\n\\item[(3)]\nthere is $C \\in [0,\\infty)$ such that\n\\begin{equation}\n \\label{eq:rough-hold-area}\n\\|\\mathbb{X}_{st}\\|_{2\\gamma} := \\sup_{t \\neq s} \n\\frac{|\\mathbb{X}_{st}|}{|t-s|^{2\\gamma}} \\le C.\n\\end{equation}\n\\end{enumerate}\n\\end{definition}\n\\noindent\nWe use the similar norm $\\|{X_{st}}\\|_\\gamma$ correspondingly given\nby the right hand side of (\\ref{eq:rough-hold-area}).\n\n\\begin{remark}\n\\rm{\nNote that Assumption\n(2) is non-trivial only if $d>1$. When $d=1$ we can take\n$\n\\mathbb{X}_{st} = (X_{st})^2\/2,\n$\nand it can easily be seen that both (\\ref{eq:mult}) and the\nH\\\"older-like bound~(\\ref{eq:rough-hold-area}) are then satisfied. \n}\n\\end{remark}\n\n\nFor our purposes below we need a slight extension since we want \nto be able to integrate functions explicitly dependent on the time\nparameter, such as $\\int_0^T \\varphi(u,X_u) dX_u$. One possibility \nis to consider the couple $(u,X)$ as a new path and construct the \nassociated rough path. This construction, however, has the disadvantage \nof requiring rather much regularity of the function $\\varphi$. A more \neffective approach is to treat the line-integral as a Young integral \nwith respect to the $u$ dependence of the integrand as soon as the \nmap $u\\mapsto \\varphi(u,\\xi)$ has H\\\"older exponent greater than \n$1-\\gamma$ (as we shall see below).\n\nFor the (time-dependent) test vector fields in $C(\\mathbb{R} \\times \\mathbb{R}^d; \\mathbb{R}^d)$\n we use the norm\n$$\n\\|\\varphi\\|_{\\rho,2,s,t} = \\sup_x \\left[ \\sup_{u\\in[s,t]}\n\\left(\\max_{k=0,1,2} |\\nabla^k \\varphi(u,x)|\\right) +\n \\sup_{u,v \\in[s,t]} \\max_{k=0,1}\n\\frac{|\\nabla^k \\varphi(u,x)-\\nabla^k \\varphi(v,x)|}{|u-v|^\\rho} \n\\right]\n$$\nwith $\\rho > 0$ and the convention $\\nabla^0 \\varphi(t,x) = \n\\varphi(t,x)$. \nThen our basic result on the step-2 rough path $(X,\\mathbb{X})$ is\n\n\\begin{theorem}\n\\label{th:rough-main}\nLet $\\varphi \\in C(\\mathbb{R} \\times \\mathbb{R}^d,\\mathbb{R}^d)$ be such that it is $C^2$ \nwith respect to its second variable and H\\\"older continuous with\nexponent $\\rho$ with respect to its first variable, such that \n$\\rho+\\gamma > 1$. Then the sums\n$$\n\\sum_\\alpha \\left(\\varphi_i(\\tau_\\alpha,X_{\\tau_\\alpha}) \nX^i_{\\tau_{\\alpha+1} \\tau_\\alpha}+\\nabla_j \n\\varphi_i(\\tau_\\alpha,X_{\\tau_\\alpha}) \n\\mathbb{X}_{ \\tau_\\alpha \\tau_{\\alpha+1}}^{ij} \\right), \n\\quad i,j=1,...,d\n$$\nconverge as the mesh of the partition $\\{\\tau_\\alpha\\}_{\\alpha}$ of\n$[0,T]$ goes to zero, and defines the integral\n$\n\\int_0^T \\varphi(u,X_u) dX_u \n$ \nMoreover for any $T\\le 1$ we have the bound\n$$\n\\left|\\int_0^T \\varphi(u,X_u) dX_u \\right| \\le C T^\\gamma \n\\|\\varphi\\|_{\\rho,2,0,T} (1+\\|X\\|_\\gamma+\\|\\mathbb{X}^2\\|_{2\\gamma})^3.\n$$\n\\end{theorem}\n\\begin{proof}\nWe prove convergence over the dyadic partition of $[0,1]$, then \nconvergence of a general partition follows then by the arguments \ndeveloped in~\\cite{MR2091358}. Let $\\tau^{(n)}_{\\alpha}=T \\alpha\/2^n$ \nfor $\\alpha = 0,\\dots,2^n$, and let\n$$\nS_n = \\sum_{\\alpha=0}^{2^n-1} \n\\left(\\varphi_i(\\tau^{(n)}_\\alpha,X_{\\tau^{(n)}_\\alpha}) \nX^i_{\\tau^{(n)}_{\\alpha+1}\n\\tau^{(n)}_\\alpha}+\\nabla_j \n\\varphi_i(\\tau^{(n)}_\\alpha,X_{\\tau^{(n)}_\\alpha}) \n\\mathbb{X}_{\\tau^{(n)}_{\\alpha+1}\n \\tau^{(n)}_\\alpha}^{ij}\\right). \n$$\nNote that\n\\begin{equation}\n\\label{eq:rough-split}\n S_{n}-S_{n-1} = \\sum_{\\alpha=0}^{2^n-1} \n(A_\\alpha^1 + A_\\alpha^2 +\n A_\\alpha^3 + A_\\alpha^4),\n\\end{equation}\nwhere\n\\begin{eqnarray*}\n&& \nA^1_\\alpha =\n\\left[\\varphi_i(\\tau^{(n)}_{2\\alpha+1},X_{\\tau^{(n)}_{2\\alpha+2}})-\n\\varphi_i(\\tau^{(n)}_{2\\alpha+1},X_{\\tau^{(n)}_{2\\alpha+1}})-\\nabla_j\n\\varphi_i(\\tau^{(n)}_{2\\alpha+1},X_{\\tau^{(n)}_{2\\alpha+1}}) \nX^j_{\\tau^{(n)}_{2\\alpha+2} \\tau^{(n)}_{2\\alpha+1}} \\right] \nX^i_{\\tau^{(n)}_{2\\alpha+1} \\tau^{(n)}_{2\\alpha}}, \\\\\n&& \nA^2_\\alpha =\n\\left[\\nabla_j\\varphi_i(\\tau^{(n)}_{2\\alpha+1},X_{\\tau^{(n)}_{2\\alpha+2}})- \n\\nabla_j \\varphi_i(\\tau^{(n)}_{2\\alpha+1},X_{\\tau^{(n)}_{2\\alpha+1}})\\right] \n\\mathbb{X}_{\\tau^{(n)}_{2\\alpha+1} \\tau^{(n)}_{2\\alpha}}^{ij},\\\\\n&&\nA^3_\\alpha = \n\\left[\\varphi_i(\\tau^{(n)}_{2\\alpha+2},X_{\\tau^{(n)}_{2\\alpha+2}})-\n\\varphi_i(\\tau^{(n)}_{2\\alpha+1},X_{\\tau^{(n)}_{2\\alpha+2}}) \\right] \nX^i_{\\tau^{(n)}_{2\\alpha+1}\\tau^{(n)}_{2\\alpha}},\\\\\n&& \nA^4_\\alpha = \n(\\nabla_j\\varphi_i(\\tau^{(n)}_{2\\alpha+2},X_{\\tau^{(n)}_{2\\alpha+2}})-\n\\nabla_j \\varphi_i(\\tau^{(n)}_{2\\alpha+1},X_{\\tau^{(n)}_{2\\alpha+2}})) \n\\mathbb{X}_{\\tau^{(n)}_{2\\alpha+1}\\tau^{(n)}_{2\\alpha}}^{ij}.\n\\end{eqnarray*}\nBy using the fact that $\\varphi\\in C^2$, we have the bounds \n$\n|A^1_\\alpha | \\le T^{3\\gamma}\\|\\varphi\\|_{\\rho,2,0,T} \\|X\\|^3_\\gamma\n2^{-3\\gamma n}\n$\nand\n$\n|A^2_\\alpha| \\le T^{3\\gamma} \\|\\varphi\\|_{\\rho,2,0,T} \\|X\\|_\\gamma\n\\|\\mathbb{X}\\|_{2\\gamma} 2^{-3\\gamma n}\n$\non the first two terms.\nFor the last two we have\n$\n|A^3_\\alpha| \\le T^{\\rho+\\gamma} \\|\\varphi\\|_{\\rho,2,0,T}\n\\|X\\|_\\gamma 2^{-n(\\rho+\\gamma)}\n$\nand \n$\n| A^4_\\alpha | \\le T^{\\rho+2\\gamma} \\|\\varphi\\|_{\\rho,2,0,T}\n\\|\\mathbb{X}\\|_{2\\gamma} 2^{-n(\\rho+2\\gamma)}\n$.\nThus\n$$\n|S_{n}-S_{n-1}| \\; \\le \\; C (T^{3\\gamma} +\nT^{\\rho+\\gamma})\\|\\varphi\\|_{\\rho,2,0,T} (2^{(1-3\\gamma) n}\n+ 2^{(1-\\gamma-\\rho)n}) \n$$\nwhere $C$ is a function dependent only on $X,\\mathbb{X}$. Write\n$S_n$ as the telescopic sum \n$\nS_n = S_0 + \\sum_{k=1}^n(S_k-S_{k-1})\n$.\nThen, since\n$$\n\\sum_{k=1}^n |S_k-S_{k-1}| \\le C (T^{3\\gamma}+T^{\\rho+\\gamma}) \n\\|\\varphi\\|_{\\rho,2,0,T} \\sum_{k=1}^n (2^{(1-3\\gamma) n}+2^{(1-\\gamma-\\rho) n})\n$$\nconverges geometrically whenever $\\gamma > 1\/3$ and $\\gamma+\\rho>1$,\nwe obtain absolute convergence of the sequence $S_n$. The bound on the \nintegral thus follows.\n\\end{proof}\n\n\\medskip\nAs this theorem shows, by using the additional information provided by \nthe path $\\mathbb{X}$ we are able to make Riemann sums converge. The \nfollowing consequence is immediate.\n\\begin{corollary}\nThe integral defined in Theorem \\ref{th:rough-main} has\n$\n\\mathbb{X}_{st}^{ij} = \\int_{s}^t X^i_{su} dX^j_u\n$. \n\\end{corollary}\n\\begin{proof}\nWe have\n$$\n\\int_{s}^t X^{i}_{us} dX^j_u = \\int_{s}^t X^i_{u} dX^j_u - X^i_s \n(X^j_u-X^j_s),\n$$ \nas it is easily seen by using the definition. That is, the rough \nintegral has the same linearity property as Riemann integral and\nbehaves the same way with respect to the integration of constants. \nMoreover, the integral in the right hand side corresponds to the \nfunction $\\varphi_k(\\xi) = \\xi^i \\delta_{kj}$ and by using \n(\\ref{eq:rough-split}) we have $S_{n}=S_{n-1}$ on the dyadic partition \nabove. Hence $S_n = S_0$ and \n$$\n\\int_{s}^t X^i_{u} dX^j_u = S_0 = \\sum_k \\varphi_k(X_s) X^k_{st} +\n\\sum_{k,m} \\nabla_m \\varphi_k(X_s) \\mathbb{X}_{st}^{km} = X_s^i\nX^j_{st} + \\mathbb{X}_{st}^{ij},\n$$\nwhich completes the proof.\n\\end{proof}\n\n\\begin{remark}\n\\rm{\nThe function $\\mathbb{X}$ can be identified as giving the value of a \n\\emph{twice iterated integral} over $X$,\n\\begin{equation}\n \\label{eq:rough-ident}\n\\mathbb{X}_{st}^{ij} = \\int_s^t \\left(\\int_s^u dX^i_v\\right) dX^j_u. \n\\end{equation}\nThus Theorem \\ref{th:rough-main} can be alternatively interpreted as \nsaying that the knowledge of the twice iterated integral (in addition \nwith some H\\\"older continuity) is sufficient for determining the value \nof the integral $\\int \\varphi(X_t)dX_t$ for arbitrary $C^2$ function. \n}\n\\end{remark}\n\nProvided $\\gamma > 1\/2$ and (\\ref{eq:rough-ident}) holds, whenever the\nright hand side is defined by using the Young integral, the integral \ndefined in Theorem \\ref{th:rough-main} coincides with the Young\nintegral. This is easy to see due to the estimate \n$$\n|\\mathbb{X}_{st}^{ij}| \\le C (\\|X\\|_{\\infty}+\\|X\\|_\\gamma)^2 \n|t-s|^{2\\gamma}\n$$\nfor the Young integral. Since $2\\gamma > 1 $, the sums\n$\n\\sum_\\alpha \\nabla_j \\varphi_i(X_{\\tau_\\alpha}) \n\\mathbb{X}_{\\tau_{\\alpha+1} \\tau_\\alpha}^{ij} \n$\nvanish as the mesh goes to zero, so the modified and Riemann sums \nconverge to the same limit.\nMoreover, the rough integral is continuous in the natural topology\nassociated to the step-2 rough path $(X,\\mathbb{X})$, i.e., we have\n\\begin{corollary}\nLet $(X^n,\\mathbb{X}^n)$ be a sequence of step-2 rough paths such \nthat\n$\n\\|X^n - X\\|_\\gamma + \\|\\mathbb{X}^n-\\mathbb{X}\\|_{2\\gamma} \\to 0.\n$ \nThen\n$\n\\int \\varphi(X^n) dX^n \\to \\int \\varphi(X) dX,\n$\nfor all $\\varphi \\in C^2$.\n\\end{corollary}\n\nSuppose $X^n$ is a sequence of regular (say, piecewise linear)\napproximations of the H\\\"older continuous path $X$. Then by putting \n$\\mathbb{X}^n = \\int \\int dX^n\\otimes dX^n $, where the integrals \nare Riemann integrals, a sufficient condition for the convergence of \nthe approximate integrals $\\int \\varphi(X^n) dX^n$ to $\\int \\varphi(X) \ndX$ is that the sequence $\\mathbb{X}^n$ converges to $\\mathbb{X}$ in \nthe topology induced by $\\|\\cdot\\|_{2\\gamma}$.\n\n\n\\subsection{Brownian motion as rough path}\n\nLet now $X$ be a sample path of Brownian motion. By Kolmogorov's Lemma \nthe process $t \\mapsto X_t$ has a version that is H\\\"older continuous \nwith exponent $\\gamma \\in(1\/3,1\/2)$ (actually $\\gamma$ can be taken \narbitrarily close to $1\/2$). In the following we will use such a \nversion without each time mentioning explicitly, i.e., consider a \nsubspace $\\mathcal{X}_\\gamma \\subset C(\\mathbb{R},\\mathbb{R}^d)$ such that every $X \\in \n\\mathcal{X}_\\gamma$ is H\\\"older continuous with exponent $\\gamma$.\n \nTo apply the above results to $X$ we need a choice for $\\mathbb{X}$. \nThis candidate is not unique, different choices will lead to different \nintegrals over $X$. For instance, in order to construct a possible \n$\\mathbb{X}$ we can start by setting\n$$\n(\\mathbb{X}_{\\text{It\\^o}})^{ij}_{st}=\\int_s^t \\int_s^u dX^i_v dX^j_u,\n$$ \nwhere the double integral is understood in It\\^o sense. In this way we\nobtain a family of random variables\n$\\{(\\mathbb{X}_{\\text{It\\^o}})^{ij}_{st}: i,j=1,\\dots,d; \\; t,s\n\\in[0,T]\\}$ satisfying the multiplicative property\n\\begin{equation*}\n(\\mathbb{X}_{\\text{It\\^o}})^{ij}_{st} - \n(\\mathbb{X}_{\\text{It\\^o}})^{ij}_{ut} -\n(\\mathbb{X}_{\\text{It\\^o}})^{ij}_{su} \n= X^i_{ut} X^j_{su} \n\\end{equation*}\nalmost surely for any fixed $t,s,u\\in[0,T]$. The next step is to show \nthat this family has a version for which\n\\begin{equation}\n \\label{eq:rough-holder-area}\n\\|\\mathbb{X}^{ij}_{\\text{It\\^o}}\\|_{2\\gamma} < \\infty \\quad \\mbox{a.s.}\n\\end{equation}\nIn order to prove (\\ref{eq:rough-holder-area}), we use the following\nlemma obtained as an extension of a result of Garsia-Rodemich-Rumsey \nin \\cite{MR2091358}.\n\\begin{lemma}\n\\label{lemma:besov}\nFor any $\\theta > 0$ and $p \\ge 1$ there exists a constant $C$ such\nthat for any $R \\in C([0,T]^2,{\\cal B})$, where $({\\cal B},|\\cdot|)$ \nis a Banach space, we have \n\\begin{equation}\n\\label{eq:generalboundxx}\n\\|R\\|_{\\theta} \\le C \\left(U_{\\theta+2\/p,p}(R)+V_\\theta(R)\\right),\n\\end{equation}\nwith\n\\begin{equation*}\n U_{\\theta,p}(R) = \\left[ \\int_{[0,T]^2}\n \\left(\\frac{|R_{t s}|}{|t-s|^\\theta}\\right)^p dt ds \\right]^{1\/p}\n\\end{equation*}\nand\n\\begin{equation*}\n V_\\theta(R) = \\inf_{\\theta_1 \\in(0,\\theta)} \\sup_{t \\neq u \\neq s} \n\\frac{|R_{st}-R_{ut}-R_{su}|}{|t-u|^{\\theta_1}|u-s|^{\\theta-\\theta_1}}. \n\\end{equation*}\n\\end{lemma}\n\n\\begin{corollary}\nWe have $\\|\\mathbb{X}_{{\\text{It\\^o}}}\\|_{2\\gamma} < \\infty$ $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$-almost \nsurely. \n\\end{corollary}\n\\begin{proof}\nConsider $V_{2\\gamma}(\\mathbb{X}_{\\text{It\\^o}})$. By using the \nmultiplicative property we have\n\\begin{eqnarray*}\nV_{2\\gamma}(\\mathbb{X}_{\\text{It\\^o}}) \n&\\le& \n\\inf_{\\theta_1 \\in (0,2\\gamma)} \\sup_{t \\neq u \\neq s} \n\\frac{|X_{tu}||X_{us}|}{|t-u|^{\\theta_1}|u-s|^{2\\gamma-\\theta_1}}\\\\ \n&\\le& \n\\inf_{\\theta_1 \\in(0,2\\gamma)} \\sup_{t \\neq u} \\frac{|X_{tu}|}\n{|t-u|^{\\theta_1}}\\sup_{u \\neq s}\n\\frac{|X_{us}|}{|u-s|^{2\\gamma-\\theta_1}}\\\\\n&\\le& \n\\left(\\sup_{t \\neq u} \\frac{|X_{tu}|}{|t-u|^{\\gamma}}\\right)^2\\\\\n& = &\n\\|X\\|_\\gamma^2.\n\\end{eqnarray*}\nMoreover,\n\\begin{eqnarray*}\n\\mathbb{E} [U_{2\\gamma,p}(\\mathbb{X}_{\\text{It\\^o}})^p] \n&=& \n\\mathbb{E} \\left[\\int_{[0,T]^2}\\left(\\frac{|(\\mathbb{X}_{\\text{It\\^o}})_{t s}|}\n{|t-s|^{2\\gamma}}\\right)^p dt ds \\right]\\\\\n&=& \n\\int_{[0,T]^2}\\frac{ \\mathbb{E} [|(\\mathbb{X}_{\\text{It\\^o}})_{t s}|^p]}\n{|t-s|^{2\\gamma p}} dt ds, \n\\end{eqnarray*}\nwith expectation with respect to Wiener measure. An application of the \nBurkholder-Davis-Gundy inequality~\\cite{Yor} allows to estimate the \n$p$-moment of the double stochastic integral as\n\\begin{eqnarray*}\n\\mathbb{E} [|(\\mathbb{X}_{\\text{It\\^o}})_{t s}|^p] \n& \\le&\nc_p \\mathbb{E} \\left[\\int_s^t |X_{us}|^2 ds\\right]^{p\/2} \\\\\n&\\le&\nc_p |t-s|^{p\/2-1} \\mathbb{E} \\left[\\int_s^t |X_{us}|^p ds\\right] \\\\ \n& \\le &\nc_p' |t-s|^{p\/2-1}\\int_s^t |u-s|^{p\/2} ds \\\\\n&\\le &\nc_p'' |t-s|^{p}, \n\\end{eqnarray*}\nfor all $p > 1$ and some $c_p, c_p', c_p'' > 0$. Thus\n\\begin{eqnarray*}\n\\mathbb{E} [U_{2\\gamma,p}(\\mathbb{X}_{\\text{It\\^o}})^p] \n&\\le& \nc_p'' \\int_{[0,T]^2} \\frac{ 1}{|t-s|^{(2\\gamma-1) p}} dt ds < \\infty,\n\\end{eqnarray*}\nfor any $\\gamma < 1\/2$, by choosing $p$ large enough ($p > 1\/(1-2\\gamma)$).\n\\end{proof}\n \nThis last result shows that there exists a version of the stochastic process\n$(X,\\mathbb{X}_{\\text{It\\^o}})$ that is a step-2 rough path; from now on we \ndenote by $(X,\\mathbb{X}_{\\text{It\\^o}})$ this particular version. Then \nintegrals can be defined by applying Theorem \\ref{th:rough-main}. We\ncall such an integral \\emph{rough integral} over $(X,\\mathbb{X}_{\\text{It\\^o}})$.\n\nThe relationship between the rough integral and the It\\^o integral is made\nclear by \n\n\\begin{lemma}\nThe rough integral over the couple $(X,\\mathbb{X}_{\\text{It\\^o}})$\ncoincides with the It\\^o integral for any $\\varphi \\in C^2$. \n\\end{lemma}\n\\begin{proof}\nBy It\\^o theory the sums\n$\n\\sum_\\alpha \\varphi_i(X_{\\tau_\\alpha}) X^i_{\\tau_{\\alpha+1} \\tau_\\alpha}\n$ \nconverge in probability to the It\\^o integral $\\int_0^T \\varphi(X_u) dX_u$.\nHence it suffices to show that the sums\n$\n\\sum_\\alpha \\nabla_j \\varphi_i(X_{\\tau_\\alpha}) \\mathbb{X}_{\\tau_{\\alpha+1}\n \\tau_\\alpha}^{ij} \n$\nconverge to zero in $L^2$ sense as then it follows that the two\nintegrals almost surely coincide. A simple computation shows that\n\\begin{equation}\n\\mathbb{E}\\left[ \\sum_\\alpha \\nabla_j \\varphi_i(X_{\\tau_\\alpha}) \n\\mathbb{X}_{\\tau_{\\alpha+1} \\tau_\\alpha}^{ij} \\right]^2 \n= \n\\sum_\\alpha \\mathbb{E} \\left[ \\nabla_j \\varphi_i(X_{\\tau_\\alpha}) \n\\mathbb{X}_{\\tau_{\\alpha+1} \\tau_\\alpha}^{ij} \\right]^2,\n\\nonumber\n\\end{equation}\nsince the cross terms are all zero in the mean by independence of the\nincrements of the Brownian motion. Hence \n\\begin{equation*}\n \\begin{split}\n& \\mathbb{E}\\left[ \\sum_\\alpha \\nabla_j \\varphi_i(X_{\\tau_\\alpha}) \n \\mathbb{X}_{\\tau_{\\alpha+1} \\tau_\\alpha}^{ij} \\right]^2\n \\le \\|\\varphi\\|_1 \\sum_\\alpha \\mathbb{E} \\left[ \\mathbb{X}_{\\tau_{\\alpha+1}\n \\tau_\\alpha}^{ij} \\right]^2\n \\le C \\|\\varphi\\|_1 \\sum_\\alpha |\\tau_{\\alpha+1}\n - \\tau_\\alpha|^{2}\n \\end{split}\n\\end{equation*}\nThe last sum vanishes as the mesh of the partition shrinks to zero, thus\nthe claim follows.\n\\end{proof}\n\nIn this construction choosing the It\\^o version for the double integral\n$\\mathbb{X}$ was arbitrary. Alternatively we could have considered other\ndefinitions, e.g. Stratonovich integral and let\n$$\n(\\mathbb{X}_{\\text{Stra}})_{st}^{ij} = \\int_s^t \\int_s^u \\circ dX^i_v\n\\circ dX^j_u\n$$\nwhere $\\circ dX$ stands for Stratonovich (or symmetric) integration. By \nthe same procedure we find a regular version of $\\mathbb{X}_{\\text{Stra}}$ \nsuch that $\\|\\mathbb{X}_{\\text{Stra}}\\|_{2\\gamma} < \\infty$ and can \nconstruct the rough integral over the couple $(X,\\mathbb{X}_{\\text{Stra}})$ \nwhich we denote again $\\circ dX$. It is not difficult to prove that for\n$\\varphi \\in C^2$ it coincides with the familiar Stratonovich integral.\n\nThe relationship between the rough integrals based on the It\\^o and \nStratonovich constructions follows from the identity\n$$\n(\\mathbb{X}_{\\text{Stra}})_{st}^{ij} =\n(\\mathbb{X}_{\\text{It\\^o}})_{st}^{ij} + \\frac{1}{2} \\delta_{ij} (t-s)\n$$\nbetween It\\^o and Stratonovich stochastic iterated integrals. The\ncorrection is given by the increment of the function $t \\mapsto\n\\delta_{ij} t\/2$. Thus the two rough integrals are related by the \nfamiliar formula\n$$\n\\int_0^T \\varphi(X_u) \\circ dX_u = \\int_0^T \\varphi(X_u) dX_u +\n\\frac{1}{2} \\int_0^T \\text{div } \\varphi(X_u) du.\n$$\nThis is obtained directly from the definitions with the modified\nRiemann sums. \n\n\\begin{remark}\n\\rm{\nDue to the multiplicative property the possible choices for \n$\\mathbb{X}$ differ only by the increment of a function, i.e., \nif $\\mathbb{X}_1$ and $\\mathbb{X}_2$ both satisfy the \nmultiplicative property with respect to $X$, then\n$\n(\\mathbb{X}_1)_{st}^{ij}- (\\mathbb{X}_2)_{st}^{ij} = \nf^{ij}_t-f^{ij}_s\n$ \nwith a function $f \\in C([0,T],\\mathbb{R}^d\\times \\mathbb{R}^d)$. \n}\n\\end{remark}\n\nFinally, it is not difficult to see the following regularity result.\n\\begin{corollary}\nThe map $\\varphi \\mapsto \\int_0^1 \\varphi(X_u) dX_u$ is continuous\nfrom $C^2(\\mathbb{R}^d,\\mathbb{R}^d)$ to $\\mathbb{R}$. \n\\end{corollary}\n\n\\section{Stochastic currents}\n\\subsection{Lifting Wiener measure to the space of currents}\nLet $\\mathcal{X} = C(\\mathbb{R}, \\mathbb{R}^d)$ be path space, i.e., the space of continuous\nfunctions from $\\mathbb{R}$ to $\\mathbb{R}^d$. The $\\sigma$-field $\\mathop{\\mathrm{{\\cal F}}}\\nolimits$ is generated by \nthe coordinate process $\\mathcal{X} \\ni X \\mapsto X_t \\in \\mathbb{R}^d$. For $I \\subset \n\\mathbb{R}$ we denote by $\\mathop{\\mathrm{{\\cal F}}}\\nolimits_I$ the $\\sigma$-field over $\\mathcal{X}$ generated by the \nevaluations for points in $I$, and write $\\mathop{\\mathrm{{\\cal F}}}\\nolimits_T$ when $I =\n[-T,T]$. Also, we put for a shorthand $I^c = \\mathbb{R} \\setminus I$. \nFor $s \\leq t$, the set $\\{\\mathcal{F}_{st} = \n\\sigma(X_u : s \\le u \\le t)\\}_{s \\leq t}$ is the forward filtration \nstarting at $s \\in \\mathbb{R}$.\n\nDenote as before by $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$ Wiener measure defined on $\\mathcal{X}$ (instead of \n$C([0,\\infty),\\mathbb{R}^d)$ as more usual), and write $H_0 = -(1\/2)\\Delta$. \nFor any finite division $t_1 < t_2 < ... < t_n \\in \\mathbb{R}$ we have\n\\begin{equation} \\label{wiener}\n\\mathop{\\mathrm{{\\cal W}}}\\nolimits(F) = (f_1,e^{-(t_2-t_1)H_0} f_2 \\ldots \ne^{-(t_n-t_{n-1})H_0} f_n)_{L^2(\\mathbb{R}^d,dx)}\n\\end{equation}\nwhere $F = f_1(X_{t_1})\\cdots f_n(X_{t_n})$. Here it is understood\nthat $f_2, \\ldots, f_{n-1}$ act as multiplication operators for which \nwe use the same symbol as for the corresponding functions. The\noperator $e^{-tH_0}$ has the integral kernel \n\\begin{equation} \\label{(2.2)}\n\\Pi_t(x,y) = \\frac{1}{(2 \\pi t)^{d\/2}} \n\\exp \\left( - \\frac{1}{2t} (x-y)^2 \\right), \\quad x,y \\in \\mathbb{R}^d.\n\\end{equation}\nWe denote by $\\mathop{\\mathrm{{\\cal W}}}\\nolimits_I$ the measure $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$ restricted to $\\mathcal{X}_I = C(I,\\mathbb{R}^d)$.\nSimilarly, with given $T>0$, $\\xi, \\eta \\in \\mathbb{R}^d$ write \n\\begin{equation} \\label{(2.3)}\n\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{\\xi,\\eta}_I(F) = \\left( e^{-(-T-t_1)H_0} f_1 e^{-(t_2-t_1)H_0}f_2 \n\\ldots e^{-(t_n-t_{n-1})H_0} \\left( f_n e^{-(T-t_n)H_0}(\\cdot,\\eta) \n\\right) \\right) (\\xi)\n\\end{equation}\nwith $I=[-T,T]$ and let $\\mathop{\\mathrm{\\widehat{\\cal W}}}\\nolimits^{\\xi,\\eta}_I(F) = \\mathop{\\mathrm{{\\cal W}}}\\nolimits^{\\xi,\\eta}_I(F)\/ \n\\Pi_{2T}(\\xi,\\eta)$ be the Brownian bridge starting in $\\xi$ at $-T$ \nand ending in $\\eta$ at $T$. Under $\\mathop{\\mathrm{\\widehat{\\cal W}}}\\nolimits^{x,y}_{[-T,T]}$ the process \n$X_t$ is a Gaussian semimartingale (Brownian bridge) satisfying the SDE\n$$\ndX_u = - \\frac{y-X_u}{T-u} du + dB_u\n$$\nwhere $(B_u)_{u\\in[-T,T]}$ is a $\\mathop{\\mathrm{\\widehat{\\cal W}}}\\nolimits^{x,y}_{[-T,T]}$-Browian motion. \n\n\\medskip\nNext, let $\\D$ be a Banach space of functions from $\\mathbb{R} \\times \\mathbb{R}^d$\nto $\\mathbb{R}^d$ containing the space of smooth functions \n$C^\\infty(\\mathbb{R} \\times \\mathbb{R}^d; \\mathbb{R}^d)$. Denote with $\\|\\cdot\\|_\\D$ the\nBanach norm. Let $\\D'$ the topological dual of $\\D$, $\\Delta_I =\n\\{(s,t)\\in I^2 | s\\le t\\}$, and write $\\Delta = \\Delta_\\mathbb{R}$. \n\\begin{definition}\n$C \\in C(\\Delta ; \\D')$ is a \\emph{stochastic current} if it \nsatisfies the following properties:\n\\begin{enumerate}\n\\vspace{0.1cm}\n\\item[(1)]\n$\nC_{tt}(\\varphi) = 0, \\;\\; C_{su}(\\varphi) + C_{ut}(\\varphi) = \nC_{st}(\\varphi), \\;\\;\n$\nfor any $s\\le u\\le t$ and any $\\varphi \\in \\D$;\n\\vspace{0.1cm}\n\\item[(2)]\n\\emph{locality property:} \\quad $C_{st}(\\varphi)=0 \\;$ \nwhenever $\\;\\varphi(u,x)=0$ for all $u\\in[s,t]$, $x \\in \\mathbb{R}^d$.\n\\end{enumerate}\nWe denote $\\C \\subset C(\\Delta ; \\D')$ the space of stochastic\ncurrents. \n\\end{definition}\nSet $\\Xi = \\mathcal{X} \\times \\C$ endowed with the product topology and \nwith the Borel $\\sigma$-field (on whose component generated\nby $\\C$ we consider the topology of uniform convergence on bounded \nintervals). $\\Xi$ plays the role of joint path-current configuration \nspace. As a measurable space, it is endowed with a family of \n$\\sigma$-algebras \n$\\{\\A_{st}\\}_{t>s}$ such that $\\A_{st} = \\sigma(X_u, C_{uv}(\\varphi) \n: u,v \\in [t,s], \\varphi \\in \\D )$. Similarly, we can define the \nforward filtration $\\{\\A_{t}^+\\}_t = \\{\\A_{+\\infty,t}\\}_t$ and the \nbackward filtration $\\{\\A^-_t\\}_t = \\{\\A_{t,-\\infty}\\}_t$. The above \ndefinitions make sense also in the case that the parameter $t$ is \nrestricted to a bounded interval $I\\subset \\mathbb{R}$; in this case we denote \nwith $\\Xi_I$ the corresponding space. Whenever the limits make sense \nwe define $C^{+}_t(\\varphi) = \\lim_{s\\to\\infty} C_{ts}(\\varphi)$ and \n$C^-_t(\\varphi) = \\lim_{s\\to -\\infty}C_{st}(\\varphi)$.\n\n\n\n\n\\begin{definition}\n\\label{def:forward-current}\nA \\emph{forward current} (on $I\\subset \\mathbb{R}$) is a measure $\\eta$ on\n$\\Xi$ (on $\\Xi_I$) such that the process $X$ is an \n$(\\eta,\\{\\mathop{\\mathrm{{\\cal F}}}\\nolimits_t^+\\}_t)$-semimartingale and\n\\begin{equation}\n \\label{eq:curr-constr-0}\nC_{st}^X(\\varphi) = \\int_s^t \\varphi(u,X_u) dX_u, \n\\qquad \\eta-\\text{a.s.} \n\\end{equation}\nfor any $(s,t)\\in \\Delta$ (or $\\Delta_I$) and any adapted $\\varphi \\in \n\\D$ where on the right hand side we have the standard It\\^o integral on \nthe semimartingale $X$. When $X$ is Brownian motion, we call $\\eta$ \n\\emph{(forward) Brownian current}. \n\\end{definition}\nIn order not to multiply terminology, unless confusion may arise we \nwill use the term \\emph{current} also for the elements of $\\D'$, of \n$\\Xi$, and for the laws on $\\Xi$ without making explicit distinction. \nFor unspecified bounded intervals $I$ we use the notation $C_I^X$ for\nthe associated current with integrator $X$.\n\nNext we want to construct a (non-trivial) measure on $\\Xi$ for the\nBrownian current. Thus we start from Wiener measure $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$ (similarly \nwe could have worked with the Brownian bridge $\\mathop{\\mathrm{{\\cal W}}}\\nolimits_I^{x,y}$) and prove \nthat there exists a map $F: \\mathcal{X} \\ni \\omega \\mapsto F(\\omega)\\in\\C$ such\n that\n\\begin{equation}\n \\label{eq:curr-constr-0-bis}\nF(\\cdot)_{st}(\\varphi) = \\int_s^t \\varphi(u,X_u) dX_u, \n\\qquad \\mathop{\\mathrm{{\\cal W}}}\\nolimits-\\text{a.s.} \n\\end{equation}\nfor any adapted $\\varphi \\in \\D$, with the standard It\\^o integral \nat the right hand side. Then a measure $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp$ on $\\Xi$ can be \ndefined as the law of the couple $(X,F)$ under the measure $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$ and \nit will be a forward current. The existence of a regular version of \nmap $F$ is an interesting problem in itself which can be addressed \nby using the techniques developed in~\\cite{FGGT, FGR} for what \nconcerns the regular dependence on $\\varphi$. Unfortunately, the \ntopology $\\D$ which is implied by such approaches is unsuitable for \nour applications. Here we prefer to use the theory of rough paths \nwhich will provide the necessary regularity for the $F$ map in a \nmore convenient topology.\n\nIn Section ~\\ref{sec:rough-paths} we developed the basic tools of\nrough-path theory that we need in order to lift Wiener measure to \ncurrents. We do this next.\n\nFor any $\\alpha > 1$, let $\\D_\\alpha$ be the completion of the space of \nsmooth test vector fields with respect to the norm\n$$\n\\|\\varphi\\|_{\\D_\\alpha} = \\sup_{k \\in \\Z} (1+|k|)^\\alpha \n\\|\\varphi\\|_{\\rho,2,k,k+1}.\n$$ \nIn the following we will fix $\\alpha>1$ but otherwise arbitrarily small and \nwrite $\\D_\\alpha = \\D$.\n\\begin{lemma}\n\\label{lemma:rough}\nFor any $\\gamma \\ge 1\/3$, $\\rho > 1-\\gamma$, $t>s$ and $x,y\\in \\mathbb{R}^d$, \nthere exists a family of random variables $F \\in \\C$ such that\n(\\ref{eq:curr-constr-0}) holds with respect to $\\mathop{\\mathrm{{\\cal W}}}\\nolimits_{[s,t]}^{x,y}$, \nand which satisfy the pathwise bound\n\\begin{equation}\n \\label{eq:F-bound}\n |F_{uv}(\\varphi)| \\le C_{\\gamma,\\rho,|t-s|} |u-v|^\\gamma (1+N_{[s,t]}(X))^3 \n \\|\\varphi\\|_{\\rho,2,s,t}\n\\end{equation}\nfor any $\\varphi \\in \\D$, $(u,v) \\in \\Delta_{[s,t]}$, where\n$N_{[s,t]}(X) = (\\|X\\|_{\\gamma,[s,t]} + \n\\|\\mathbb{X}^2\\|_{2\\gamma,[s,t]})$ and where\n$C_{\\gamma,\\rho,|t-s|}$ depends only on $\\gamma,\\rho,|t-s|$.\n\\end{lemma}\n\n\\begin{proof}\nDefine\n\\begin{equation*}\nF_{uv}(\\varphi) = \n\\lim_{\\delta\\tau_\\alpha \\to 0} \n\\sum_\\alpha \\left(\\varphi(\\tau_\\alpha,X_{\\tau_\\alpha}) \n+ \\nabla \\varphi(\\tau_\\alpha,X_{\\tau_\\alpha}) \n\\mathbb{X}^2_{\\tau_\\alpha,\\tau_{\\alpha+1}}\\right)\n\\end{equation*}\nwhere $\\delta\\tau_\\alpha$ is the mesh of the partition and \n$\\mathbb{X}^2_{st}\\in C(\\mathbb{R}^2;\\mathbb{R}^d \\times \\mathbb{R}^d)$ is the \ntwice iterated It\\^o integral with respect to $X$. By the results \non rough path theory in Section \\ref{sec:rough-paths} (compare \nTheorem \\ref{th:rough-main}), this limit exists whenever \n$N_{[s,t]}(X)<\\infty$ and $\\|\\varphi\\|_{2,\\rho,s,t}<\\infty$. Otherwise \nset $F_{uv}(\\varphi) = 0$. Then $F$ is a well defined random variable \nobeying (\\ref{eq:F-bound}). Moreover, $F_{su}(\\varphi)+F_{ut}\n(\\varphi)= F_{st}(\\varphi)$ and the locality property for $F$ holds\nby definition.\n\nBy straightforward estimates we can also prove that \n$\\mathbb{E}_{\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x,y}_I} [N_I(X)^3] < \\infty$ for any $x,y,I$. Using\nthis last result, the equivalence between $F$ and the It\\^o integral \nfor the adapted vector field $\\varphi$ can be proved by the same \napproach as the one used in the case of Wiener measure.\n\\end{proof}\n\n\nA direct consequence of Lemma~\\ref{lemma:rough} is that whenever \n$N_{[s,t]}(X) < \\infty$ the map $\\varphi \\mapsto F(\\omega)_{st}\n(\\varphi)$ can be considered as an element of $\\D'$. Moreover if \nwe let\n$$\n\\mathcal{N}_{\\alpha,p}(X) = \\sum_{k \\in \\Z} (1+|k|)^{-\\alpha} \nN_{[k,k+1]}(X)^p,\n$$\nthen whenever $\\mathcal{N}_{\\alpha,3}(X) < \\infty$, the boundary \ncurrents $C^+_t$ and $C^-_t$ are well defined for any $t$ as elements \nof $\\D'$.\n\n\\begin{lemma}\n\\label{lemma:W-lift}\nFor every bounded $I \\subset \\mathbb{R} $ there exists a unique Brownian\ncurrent $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{\\sharp,x,y}_I$ on $\\Xi_I$. A similar statement holds for\nthe measures $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp$ with first marginal $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$. Moreover, since\nunder $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$ we have $\\mathcal{N}_{\\alpha,3}(X) < \\infty$ a.s., the\nboundary currents are well defined under $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp$.\n\\end{lemma}\n\\begin{proof}\nThe existence of the lifted measure for Brownian bridge $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x,y}_I$ \nis essentially contained in Lemma~\\ref{lemma:rough}. Its uniqueness \nis a direct consequence of the property~(\\ref{eq:curr-constr-0}). \nThe proof in the case of $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$ is similar and we are left to prove that\nunder $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp$ the boundary currents are well defined. The integrals \nin every interval $[i,i+1]$ are well defined, moreover\n$$\n\\left|\\int_{i}^{i+1} \\varphi(u,X_u) dX_u\\right|\\le (1+|i|)^{-\\alpha} \n\\|\\varphi\\|_{\\mathcal D} [1+N_{[i,i+1]}(X)]^3 \n$$\nso that the series\n$\n\\int_{t}^{+\\infty}\\varphi(u,X_u) dX_u = \n\\sum_{i>t-1}\\int_{i\\vee t}^{i+1} \\varphi(u,X_u) dX_u\n$\nis absolutely convergent if $\\mathcal{N}_{\\alpha,3}(X) < \\infty$.\nUnder $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$ we have\n$\n\\mathbb{E}_{\\mathop{\\mathrm{{\\cal W}}}\\nolimits} [(1+N_{[i,i+1]}(X))^3|X_0=x] \\le C\n$ \nuniformly in $i \\in \\Z$ and $x\\in \\mathbb{R}^d$ thus\n$$\n\\mathbb{E}_{\\mathop{\\mathrm{{\\cal W}}}\\nolimits} [\\mathcal{N}_{\\alpha,3}(X)|X_0=x] = \n\\sum_{i \\in \\Z}(1+|i|)^{-\\alpha} \n\\mathbb{E}_{\\mathop{\\mathrm{{\\cal W}}}\\nolimits} [N_{[i,i+1]}(X)^3|X_0=x] \n\\le \nC \\sum_{i \\in \\Z} (1+|i|)^{-\\alpha} < \\infty\n$$\nas soon as $\\alpha > 1$. This implies that $\\mathcal{N}_{\\alpha,3}(X)$\nis $\\mathop{\\mathrm{{\\cal W}}}\\nolimits$-a.s. finite.\n\\end{proof}\n\n\n\\begin{remark}\n\\rm{\nNote that the lifting of a measure from $\\mathcal{X}$ to $\\Xi$ is in general \nnot unique. For instance, we could decide to add some other term to \nthe definition of the current,\n$$\n\\tilde C_{st}^X(\\varphi) = C_{st}^X(\\varphi) + \\int_s^t \\mathrm{div}\\,\n\\varphi(t, X_t) dt\n$$\nand obtain a different lifted measure (which is no longer a forward \ncurrent). It would be interesting to explore whether different lifts \nmay have a different physical meaning in the models. For instance, \nnon-relativistic particles with spin can be (partially) described by \na current $C^\\sigma$ defined as\n$$\nC^X_{st}(\\varphi)[\\sigma] = C_{st}^X(\\varphi) + \\int_s^t \\sigma_t \n\\cdot \\mathrm{curl}\\, \\varphi(t, X_t) dt\n$$\nwhere $(\\sigma_t)_t$ is a vector-valued Poisson process describing\nthe spin of the particle~\\cite{Jona,Jona2,HL1}. \n}\n\\end{remark}\n\n\n\\subsection{It\\^o current}\n\\label{sec:ito-current}\nBeside defining Brownian currents, it will be useful for our purposes \nbelow to define currents for Brownian paths subjected to a potential \n(or penalty function) $V$. The reason is that we need a sufficiently\nstrong confining mechanism of paths in order to investigate the effect\nof a pair interaction on them (given by double stochastic integrals). \nThe translation invariant ($V \\equiv 0$) regime is presently little\nunderstood. \n\nWhile we properly introduce potentials only in the next section, we \nrequire here that the integral kernel\n\\begin{equation} \n\\label{fkf}\n\\hat\\pi_{t-s}(x,y) = \\int e^{- \\int_s^t V(X_u)du} \\, \nd\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x,y}_{[s,t]}(X), \\quad \\forall s0$ the semigroup $S_t$ is a bounded operator \nfrom $L^p$ to $L^q$ for every $ 1 \\leq p \\leq q \\leq \\infty$, and by \nthe Feynman-Kac formula and the Hille-Yoshida and Stone Theorems it\ncan be written as $S_t=e^{-tH}$, with $H$ coinciding with the\nSchr\\\"odinger operator $H_0+V$ on $C_0^\\infty(\\mathbb{R}^d)$. In addition, \n$e^{-tH}f$ is a continuous function for every $f \\in L^p, p \\in \n[1,\\infty]$, $\\forall t > 0$. \n\nLet $\\Psi$ be a ground state of the Schr\\\"odinger operator $H$, i.e., \na normalized eigenfunction in $L^2(\\mathbb{R}^d,dx)$ lying at $\\inf\\mathop{\\mathrm{Spec}}\\nolimits H$. \nUnder the conditions given in Section 4.1. this ground state is unique \nand has a strictly positive version. Using this we define the\nprobability measure $\\nu$ on $(\\mathcal{X},\\mathop{\\mathrm{{\\cal F}}}\\nolimits,\\mathop{\\mathrm{{\\cal W}}}\\nolimits)$ by\n\\begin{equation} \n\\label{pphi1}\n\\nu(A) = Z_T^{-1} \\int dx \\, \\Psi(x) \\int dy \\, \\Psi(y) \\int 1_{A}(X) \ne^{-\\int_{-T}^T(X_t)dt} \\, d\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x,y}_T(X)\n\\end{equation}\nwith normalizing constant $Z_T$, $A \\in \\mathop{\\mathrm{{\\cal F}}}\\nolimits_{T}$. This measure can be\nextended to a measure $\\nu$ on the full $\\mathop{\\mathrm{{\\cal F}}}\\nolimits$ by making use of the \nfacts $e^{-tH}\\Psi = \\Psi$ and $\\norm[2]{\\Psi} = 1$. The Feynman-Kac \nformula and the Markov property of Brownian motion imply that \n$\\{\\nu_T\\}$ given on $\\mathop{\\mathrm{{\\cal F}}}\\nolimits_{T}, \\, T>0$, define a consistent family of \nprobability measures. In particular, $\\nu$ satisfies the DLR equations \nand thus is a Gibbs measure relative to Brownian motion for potential \n$V$; for further details see \\cite{BL}. \n\nMoreover, $\\nu$ is the law of a reversible diffusion process with \nstationary distribution $d\\omega = \\Psi^2 \\, dx$ and stochastic \ngenerator $H_{\\omega}$ acting in $L^2(\\mathbb{R}^d,d\\omega)$ as\n$$ \nH_{\\omega} f = \\frac{1}{\\Psi} H (\\Psi f) = H_0 f - \n\\left(\\nabla \\log \\Psi, {\\nabla f}\\right)_{\\mathbb{R}^d}.\n$$\nThis process is called \\emph{It\\^o diffusion} (or \n\\emph{$P(\\phi)_{1}$-process} in quantum field theory). Its transition \nprobabilities are given by \n\\begin{equation} \\label{transition prob}\n\\nu ( f(X_{t+s}) | X_{s} = x) = \\int \\pi_{t}(x,y) f(y) \\, d\\omega(y),\n\\end{equation}\nwhere\n\\begin{equation} \n\\pi_{t}(x,y) = \\frac{\\hat\\pi_{t}(x,y)}{\\Psi(x)\\Psi(y)}\n\\label{pito}\n\\end{equation}\nis the transition density of $\\nu$ with respect to its stationary \ndistribution. The It\\^o process is Markovian, reversible, and has a \nversion with continuous paths. Moreover, it is a Brownian\nsemi-martingale with respect to either the forward or the backward \nfiltration, in particular it is the stationary solution of the forward \nstochastic differential equation \n\\begin{equation}\n \\label{eq:diffusion-eq}\ndX_t = \\nabla \\log \\Psi(X_t) dt + dB_t, \n\\end{equation}\nwhere $B_t$ is Brownian motion with respect to the forward filtration.\n\nFor the It\\^o bridge, i.e., the regular conditional probability \n$\\nu^{x,y}_T$ of $\\nu$ given $X_{-T} = x$, $X_{T} = y$, we will use\nthe following representation. Take (\\ref{pphi1}) describing the density \nof the measure $\\nu_T$ with respect to Brownian motion. Then\n\\begin{equation}\n \\label{eq:bridge-2}\n\\nu_T^{x,y}(A) = \\frac{ \\Pi_{2T}(x,y)}{Z_T\\, \\Psi(x) \\Psi(y) \\pi_{2T}(x,y)}\n\\mathbb{E}_{\\mathop{\\mathrm{\\widehat{\\cal W}}}\\nolimits^{x,y}_{[-T,T]}} \\left[ 1_A(X) e^{- V_{[-T,T]}(X) }\\right], \n\\end{equation}\nwith $\\Pi_t$ the Brownian transition kernel (\\ref{(2.2)}). This\nformula can be checked by noting that \n\\begin{equation*}\n \\begin{split}\n&\n\\hat \\mathbb{E}_{\\nu_T} \\left[f(x) g(y) \\mathbb{E}_{\\nu_T^{x,y}}[Q] \\right] \\\\ & \n\\quad \n= Z_T^{-1} \\hat \\mathbb{E}_{\\mathcal{W}_{[-T,T]}} \\left[f(x)g(y)\\Psi(x)\\Psi(y) \n\\mathbb{E}_{\\mathop{\\mathrm{\\widehat{\\cal W}}}\\nolimits_{[-T,T]}^{x,y}}[Q(X) e^{- \\int_{-T}^TV(X_t)dt}] \\right] \\\\ & \n\\quad \n= Z_T^{-1} \\mathbb{E}_{\\mathcal{W}_{[-T,T]}} \\left[\\Psi(X_{-T})\\Psi(X_T) \nf(X_T) g(X_{-T}) Q(X) e^{-\\int_{-T}^T V(X_t)dt}\\right] \\\\ & \n\\quad \n= \\mathbb{E}_{\\nu_T} \\left[f(X_T) g(X_{-T})Q(X) \\right]\n\\end{split}\n\\end{equation*}\nwhere $\\hat \\mathbb{E}$ denotes the expectation in a new probability\nspace whose coordinate process is denoted $\\hat X$, and with $\\hat\nX_{-T} = x$, $\\hat X_T = y$. \n\nThe assumption that $V$ is Kato-class (see Section 4.1) implies that\nthe It\\^o bridge measure $\\nu^{x,y}_{I}$ is absolutely continuous with\nrespect to the Brownian bridge measure $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x,y}_{I}$. This allows us \nto use the lifting result proved in Lemma~\\ref{lemma:W-lift} to show \nthat also $\\nu^{x,y}_{I}$ allows a lift to the space of currents. \n\\begin{lemma}\n\\label{lemma:nu-lift}\nFor every bounded $I \\subset \\mathbb{R}$ there exists a forward current \n$\\nu^{\\sharp,x,y}_I$ on $\\Xi_I$ such that is first marginal is\n$\\nu^{x,y}_I$. A similar statement holds for the stationary measures \n$\\nu^\\sharp$ with first marginal $\\nu$. Moverover, since under $\\nu$ \nwe have $\\mathcal{N}_{\\alpha,3}(X) < \\infty$ a.s., the boundary \ncurrents are well defined under $\\nu^\\sharp$.\n\\end{lemma}\n\\begin{proof}\nWe have\n$$\n\\mathbb{E}_{\\nu^{x,y}_I} [N_I(X)^3] \\le \nC \\left(\\mathbb{E}_{\\widehat \\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x,y}_I} [N_I(X)^6]\\right)^{1\/2} \n\\left(\\mathbb{E}_{\\nu^{x,y}_I} [e^{-2\\int_I V(X_t)dt}]\\right)^{1\/2} \n\\le C.\n$$\nHence the map $F$ defined in Lemma~\\ref{lemma:rough} is well defined\nand coincides almost surely with the It\\^o integral. This allows to\nconstruct the lifted measures $\\nu^{\\sharp,x,y}_I$ and $\\nu^{\\sharp}$. \nMoreover, by using stationarity of $\\nu$ we have that\n$\n\\mathbb{E}_{\\nu} [N_{[i,i+1]}(X)^3] \\le C\n$\nuniformly in $i\\in \\Z$, thus the boundary currents are well defined \nunder $\\nu^\\sharp$.\n\\end{proof}\n\n\n\\section{Gibbs measures on Brownian currents}\n\\subsection{Conditions on the potentials}\n\\label{sec:potentials}\nWe use the same terminology of the usual DLR theory and introduce \n``potentials\" and ``energy functionals\" below. \n\nAn \\emph{external potential} is a Lebesgue measurable function $V: \n\\mathbb{R}^d \\to \\mathbb{R}$ that we will choose from the Kato class, i.e., an element \nof the space $\\mathcal{K}(\\mathbb{R}^d)$ defined by the condition\n\\begin{equation}\n\\lim_{r \\to 0} \\sup_{x \\in \\mathbb{R}^d} \\int_{B_r(x)} \n|g(y-x)V(y)| \\, dy = 0,\n\\end{equation}\nwith $B_r(x)$ the ball of radius $r$ centered at $x$, and \n\\begin{equation}\n g(x) = \\left\\{ %\n\\begin{array}{ll} \n |x| & \\mbox{if} \\;\\; d = 1 \\\\ \n -\\ln|x| & \\mbox{if} \\;\\; d = 2 \\label{k2}\\\\\n |x|^{2-d} & \\mbox{if} \\;\\; d \\geq 3.\n \\end{array} \\right. \n\\end{equation}\nThis space is large enough to contain many choices of interest,\nwhile allowing the Feynman-Kac formula for the Schr\\\"odinger\nsemigroup $e^{-tH}$, $t \\geq 0$, to hold. This is generated by \nthe Schr\\\"odinger operator $H = H_0 + V$ defined on \n$L^2(\\mathbb{R}^d,dx)$ as a form sum ($V$ regarded as a multiplication \noperator). For Kato-class potentials $H$ is essentially \nself-adjoint on the form core $C^\\infty_0(\\mathbb{R}^d)$. In addition, \nwe will require of $V$ to be such that \n\\begin{enumerate}\n\\item[(1)]\n$H$ has a unique strictly positive eigenfunction (ground state) \n$\\Psi$ at $E = \\inf\\mathop{\\mathrm{Spec}}\\nolimits H$, with the property that $\\Psi \\in\nL^1 \\cup L^\\infty$;\n\\item[(2)] \n$e^{-tH}$ is intrinsically ultracontractive.\n\\end{enumerate}\nRecall the meaning of the latter property. Write $d\\omega = \\Psi^2 \ndx$ on $\\mathbb{R}^d$ as before, and define the isometry (\\emph{ground \nstate transform}) $j: L^2(\\mathbb{R}^d,d\\omega) \\rightarrow L^2(\\mathbb{R}^d,dx)$, \n$f \\mapsto \\Psi f$. Then $D(H_\\omega) = j^{-1}D(H)$ and $H_\\omega \nf=(j^{-1}Hj)f = (1\/\\Psi) H(\\Psi f) = -(1\/2)\\Delta f-({\\nabla\\ln\\Psi},\n\\nabla f)_{\\mathbb{R}^d}$, for every $f \\in D(H_\\omega)$. The associated \nsemigroup $e^{-tH_\\omega}$ exists for all $f\\in L^2(\\mathbb{R}^d,d\\omega)$ \nand $t \\geq 0$. $e^{-tH}$ is \\emph{intrinsically ultracontractive} \nwhen $e^{-tH_\\omega}$ is \\emph{ultracontractive}, i.e., it maps \n$L^2(\\mathbb{R}^d, d\\omega)$ into $L^\\infty(\\mathbb{R}^d,d\\omega)$ continuously. \nEquivalently, this means that $\\|{e^{-tH_\\omega}}\\|_{2,\\infty} < \n\\infty$, $\\forall t> 0$, and it is a monotonically decreasing \nfunction in $t$. Moreover, the integral kernel (\\ref{pito}) of \n$e^{-tH_\\omega}$ satisfies $0 \\leq \\pi_t(x,y) \\leq \n\\|e^{-(t\/2)H_\\omega}\\|_{2,\\infty}^2$ almost surely.\n\nThese conditions are in particular satisfied for $V$ bounded \nfrom below, continuous, and sufficiently confining, i.e., for \nwhich there exist constants $C_1, C_2 > 0$, $C_3, C_4\\in \\mathbb{R}$, \nand $a,b$ with $2 < a < b < 2a - 2$ such that the positive \npart of the potential, $V^+ = \\sup \\{0,V\\}$ satisfies\n\\begin{equation} \\label{iuccond}\nC_1|x|^a + C_3 \\leq V^{+}(x) \\leq C_2|x|^b + C_4.\n\\end{equation}\n\n\n\\bigskip\n\nA \\emph{pair interaction potential} is a Lebesgue measurable \nfunction $W: \\mathbb{R}^d \\times \\mathbb{R} \\to \\mathbb{R}$, even in both of its \nvariables, which we require to \n\\begin{enumerate}\n\\item[(1)]\nhave positive Fourier transform;\n\\item[(2)]\nsatisfy the regularity condition that there exists \n$M_{I,\\beta} \\in \\mathbb{R}$ such that\n\\begin{equation}\n\\sup_{x\\in \\mathbb{R}^d ,t\\in I} \\|W(x,t)\\|_{\\D_{\\beta}} \\leq\nM_{I,\\beta},\n\\label{a22}\n\\end{equation}\nfor $\\beta > \\max\\{\\alpha, 3\\}$ and every bounded $I \\subset \n\\mathbb{R}$.\n\\end{enumerate}\nThe requirement $\\beta>\\alpha$ is needed for having a well \ndefined interaction energy (actually in our applications for \nany $\\beta>1$ there is a suitable $\\alpha$ so that this holds), \nwhile $\\beta>3$ is a decay condition sufficient for ensuring \nthe convergence of the cluster expansion in Chapter 5 below. \n \nAn example satisfying these conditions is $W^\\rho$ seen in \nSection 1.\n\n\\bigskip\nFinally we write down the energies appearing in the definition of \nthe densities of Gibbs measures we are going to study. With given \npair potential $W$, for all $a,b \\in \\mathcal{D}'$ consider the \nfunction $\\psi_{k,\\varpi}(x,t) = e^{ik \\cdot x +i\\varpi t}\\in \n\\mathcal{D}$ and define the (possibly unbounded) quadratic form \n$$\n\\langle a,b \\rangle_W := \\int \\widehat W(k,\\varpi) \na(\\psi_{k,\\varpi})b(\\overline{\\psi_{k,\\varpi}}) dk d\\varpi. \n$$\nBy using the quadratic form, \nfor all bounded $I \\subset \\mathbb{R}$ and every $\\mathbb{X} = (X, C^X), \n\\, \\mathbb{Y} = (Y, C^Y) \\in \\Xi$ define the \\emph{internal energy \nfunctional}\n\\begin{equation}\nH_I(\\mathbb{X}) = V_I(X) + \\frac{\\lambda }{2}\\langle \nC^X_I, C^X_I \\rangle_W,\n\\label{hin}\n\\end{equation}\nand \n\\emph{interaction functional}\n\\begin{equation}\nH_I(\\mathbb{X}|\\mathbb{Y}) = V_I(X) + \n\\frac{\\lambda}{2}\\langle C^X_I, C^X_I \\rangle_W + \n\\lambda \\langle C^X_I, C^Y_{I^c} \\rangle_W,\n\\label{hint}\n\\end{equation}\nwith parameter $\\lambda \\in \\mathbb{R}$, where we wrote $V_I(X) = \n\\int_I V(X_t) dt$.\n\n\n\n\n\n\n\\subsection{Gibbs specifications}\n\n\nOn $\\Xi$ with its associated $\\sigma$-Borel field $\\A$ we take now \n$\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp$ as reference measure and define a Gibbs specification.\n\\begin{definition}[Gibbs specification]\nTake the regular version $\\mathop{\\mathrm{{\\cal W}}}\\nolimits_I^\\sharp(d\\mathbb{X}|\\mathbb{Y})$ of \nthe measure $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp$ conditional on $\\mathbb{Y}$ in $\\A_{I^c}$, \nand $H_I(\\mathbb{X})$ given by (\\ref{hin}). We call the family of\nprobability kernels $\\{\\mu^\\sharp_I\\}_I$ on $\\Xi$ indexed by the \nbounded intervals $I \\subset \\mathbb{R}$, \n\\begin{equation}\n\\label{eq:curr-finite-vol}\n\\mu^\\sharp_I(d\\mathbb{X}) = \\frac{e^{-H_I(\\mathbb{X})}}\n{Z_I} \\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp_I(d\\mathbb{X}) \n\\end{equation}\na \\emph{Gibbs specification on Brownian currents with free \nboundary condition}. Take $H_I(\\mathbb{X}|\\mathbb{Y})$ given \nby (\\ref{hint}). We call the family $\\{\\rho^\\sharp_I\\}_I$ on\n$\\Xi$,\n\\begin{equation}\n\\label{eq:spec}\n\\rho^\\sharp_I(d\\mathbb{X}|\\mathbb{Y}) = \n\\frac{e^{-H_I(\\mathbb{X}|\\mathbb{Y})}}{Z_I(\\mathbb{Y})} \n\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp_I(d\\mathbb{X} | \\mathbb{Y}) \n\\end{equation}\na \\emph{Gibbs specification on Brownian currents with boundary\ncondition $\\mathbb{Y}$}. \n\\end{definition}\n\n\\begin{definition}[Gibbs measure]\nA probability measure $\\mu$ on $(\\Xi,\\A,\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp)$ is a \n\\emph{Gibbs measure} for the potentials $V$ and $W$ if it is \nconsistent with the specification $\\{\\rho^\\sharp_I\\}_I$, i.e.,\nthere exists a version of its conditional probabilities with\nrespect to the family $\\{\\A_{I^c}\\}_I$ which agrees with \n$\\{\\rho^\\sharp_I\\}_I$ for all bounded $I \\subset \\mathbb{R}$. \n\\end{definition}\nIn the following chapter our main concern will be to prove the\nexistence of such Gibbs measures. \n\nOn Gibbs specifications here is a first result. \n\\begin{lemma}\nThe family $\\{ \\rho_I^\\sharp \\}_I$ is consistent, i.e., for \nevery pair of bounded intervals $I \\subset J \\subset \\mathbb{R}$ \nwe have $\\int\\int F(\\mathbb{X})\\rho_I^\\sharp(d\\mathbb{X}|\\mathbb{Y})\n\\rho_J^\\sharp(d\\mathbb{Y}|\\mathbb{Z}) = \\int F(\\mathbb{X})\n\\rho_J^\\sharp(d\\mathbb{X}|\\mathbb{Z})$, for any bounded measurable $F:\n\\Xi \\to \\mathbb{R}$.\n\\end{lemma}\n\\begin{proof}\nThe family $\\{ \\mathop{\\mathrm{{\\cal W}}}\\nolimits_I^\\sharp \\}_I$ is consistent by its definition. \nHence\n\\begin{equation*}\n \\begin{split}\n\\int \\int & F(\\mathbb{X})\n\\rho_I^\\sharp(d\\mathbb{X}|\\mathbb{Y})\n\\rho_J^\\sharp(d\\mathbb{Y}|\\mathbb{Z}) \n\\\\ & =\n \\int \\int F(\\mathbb{X}) \\frac{e^{-V_I(X)-(\\lambda\/2)\n \\langle C^X_I, C^X_I \\rangle_W - \n \\lambda\\langle C^X_I,C^Y_{I^c}\\rangle_W}}{Z_I(\\mathbb{Y})}\n\\\\ & \\quad \\quad \\times\n \\frac{e^{-V_J(Y)-(\\lambda\/2)\\langle C^Y_J, C^Y_J \\rangle_W-\n \\lambda\\langle C^Y_J,C^Z_{J^c} \\rangle_W}}{Z_J(\\mathbb{Z})}\n\\mathop{\\mathrm{{\\cal W}}}\\nolimits_I^\\sharp(d\\mathbb{X}|\\mathbb{Y})\\mathop{\\mathrm{{\\cal W}}}\\nolimits_J^\\sharp(d\\mathbb{Y}|\\mathbb{Z}) \n\\\\ &= \n \\int \\int F(\\mathbb{X}) \\frac{e^{-V_I(X)-(\\lambda\/2)\n \\langle C^X_I, C^X_I \\rangle_W-\n \\lambda\\langle C^X_I,C^Y_{I^c}\\rangle_W}}{Z_I(\\mathbb{Y})}\n\\\\ & \\quad \\quad \\times\n{e^{-V_I(Y)-(\\lambda\/2)\\langle C^Y_I, C^Y_I \\rangle_W\n-\\lambda \\langle C^Y_I,C^Z_{J^c} \\rangle_W -\n\\lambda\\langle C^Y_I,C^Y_{K} \\rangle_W }}{}\n\\\\ & \\quad \\quad \\times\n \\frac{e^{-V_K(Y)-(\\lambda\/2)\\langle C^Y_K, C^Y_K \\rangle_W-\n \\lambda\\langle C^Y_K,C^Z_{J^c} \\rangle_W}}{Z_J(\\mathbb{Z})}\n\\mathop{\\mathrm{{\\cal W}}}\\nolimits_I^\\sharp(d\\mathbb{X}|\\mathbb{Y})\\mathop{\\mathrm{{\\cal W}}}\\nolimits_J^\\sharp(d\\mathbb{Y}|\\mathbb{Z}) \n \\end{split}\n\\end{equation*}\nwhere we split off $C^Y_J = C^Y_I + C^Y_K$ with $K = J\\backslash \nI$. This gives for the right hand side\n\\begin{equation*}\n \\begin{split}\n\\quad &\n \\int \\int F(\\mathbb{X}) \\frac{e^{-V_I(X)-(\\lambda\/2)\n \\langle C^X_I,C^X_I \\rangle_W-\n \\lambda\\langle C^X_I,C^Y_{I^c}\\rangle_W}}{Z_I(\\mathbb{Y}_2)}\n\\\\ & \\quad \\quad \\times\n\\left( \\int {e^{-V_I(Y_1)-(\\lambda\/2)\\langle C^{Y_1}_I,C^{Y_1}_I\n\\rangle_W-\\lambda\\langle C^{Y_1}_I,C^Z_{J^c} \\rangle_W -\n\\lambda\\langle C^{Y_1}_I,C^{Y_2}_{K} \\rangle_W }}{} \n \\mathop{\\mathrm{{\\cal W}}}\\nolimits_I^\\sharp(d\\mathbb{Y}_1|\\mathbb{Y}_2)\\right)\n\\\\ & \\quad \\quad \\times\n \\frac\n{e^{-V_K(Y_2)-(\\lambda\/2)\\langle C^{Y_2}_K, C^{Y_2}_K \\rangle_W-\n\\lambda\\langle C^{Y_2}_K,C^Z_{J^c} \\rangle_W}}{Z_J(\\mathbb{Z})}\n\\mathop{\\mathrm{{\\cal W}}}\\nolimits_I^\\sharp(d\\mathbb{X}|\\mathbb{Y}_2) \\mathop{\\mathrm{{\\cal W}}}\\nolimits_K^\\sharp(d\\mathbb{Y}_2|\\mathbb{Z}), \n \\end{split}\n\\end{equation*}\nwhere we used the fact that \n$\\int \\mathop{\\mathrm{{\\cal W}}}\\nolimits_I^\\sharp(d\\mathbb{Y}_1|\\mathbb{Y}_2)\\mathop{\\mathrm{{\\cal W}}}\\nolimits_K^\\sharp\n(d\\mathbb{Y}_2|\\mathbb{Z}) = \\mathop{\\mathrm{{\\cal W}}}\\nolimits_J^\\sharp(d\\mathbb{Y}_1|\\mathbb{Z})$. Note that \nthe expression between the brackets equals $Z_I(\\mathbb{Y}_2)$, \nthus we further obtain\n\\begin{equation*}\n \\begin{split}\n\\quad & \n \\int \\int F(\\mathbb{X}) e^{-V_I(X)-(\\lambda\/2)\n \\langle C^X_I, C^X_I \\rangle_W-\n \\lambda\\langle C^X_I,C^Y_{I^c} \\rangle_W}\n\\\\ & \\quad \\quad \\times\n \\frac\n{e^{-V_K(Y_2)-(\\lambda\/2)\\langle C^{Y_2}_K, C^{Y_2}_K \\rangle_W -\n\\lambda\\langle C^{Y_2}_K,C^Z_{J^c} \\rangle_W}}{Z_J(\\mathbb{Z})}\n\\mathop{\\mathrm{{\\cal W}}}\\nolimits_I^\\sharp(d\\mathbb{X}|\\mathbb{Y}_2) \\mathop{\\mathrm{{\\cal W}}}\\nolimits_K^\\sharp(d\\mathbb{Y}_2|\\mathbb{Z}) \n\\\\ \\quad &= \n \\int F(\\mathbb{Y})\n\\frac{e^{-V_J(Y)-(\\lambda\/2)\\langle C^{Y}_J, C^{Y}_J \\rangle_W -\n\\lambda\\langle C^{Y}_J,C^Z_{J^c} \\rangle_W}}{Z_J(\\mathbb{Z})}\n\\mathop{\\mathrm{{\\cal W}}}\\nolimits_J^\\sharp(d\\mathbb{Y}|\\mathbb{Z}) = \\int F(\\mathbb{Y}) \n\\rho_J(d\\mathbb{Y}|\\mathbb{Z}).\n \\end{split}\n\\end{equation*}\n\\end{proof}\n\n\nThe forward current $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp$ has the key property that\n$\nC^X_{st}(\\varphi) = \\int_s^t \\varphi(u,X_u) dX_u\n$, \n$\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp$-a.s. for all $(s,t)\\in \\Delta$ and all adapted \n$\\varphi\\in \\D$. This will enable us to show that the finite \nvolume measures $\\mu_I$ coincide with the marginals of the \nmeasures $\\mu^\\sharp_I$ on the first component of the product \n$\\Xi$. The specification $\\{\\rho^\\sharp_I\\}$ can then be \nconsidered as a suitable rigorous replacement for the DLR \ndescription of the infinite-volume limit. A Gibbs measure on \n$\\mathcal{X}$ will then be a measure for which there exists a unique \nlift to the space $\\Xi$ of currents satisfying the relation~\n(\\ref{eq:curr-constr-0}) ensuring the identification of the \ncurrent with the stochastic integral and which satisfy the \nDLR conditions with respect to the specification \n$\\{\\rho^\\sharp_I\\}$.\n\nTo show that the specification is well defined we rewrite the \nvarious terms using the fact that, under the measure \n$\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp_I(d\\mathbb{X}|\\mathbb{Y})$ we have pathwise \nequality between the current $C^X$ and the stochastic integral \nwith respect to $X$ for adapted integrands belonging to $\\D$. \nThen\n\\begin{equation*}\n \\begin{split}\n\\langle C_I^X, C_I^X \\rangle_W & = \\int \\widehat W(k,\\varpi)\nC_I^X(\\psi_{k,\\varpi}) C_I^X(\\overline{\\psi_{k,\\varpi}})\ndk d\\varpi\n \\\\ & = \\int \\widehat W(k,\\varpi) |C_I^X(\\psi_{k,\\varpi})|^2\n dk d\\varpi \n \\\\ & = \\int \\widehat W(k,\\varpi)\n\\left|\\int_I \\psi_{k,\\varpi}(t,X_t) dX_t \\right|^2 \ndk d\\varpi = 2 W_I(X)\n \\end{split}\n\\end{equation*}\nand\n\\begin{equation*}\n \\begin{split}\n\\langle C_I^X, C_{I^c}^Y \\rangle_W & = \n\\int \\widehat W(k,\\varpi) C_I^X(\\psi_{k,\\varpi}) C_{I^c}^Y\n(\\overline{\\psi_{k,\\varpi}}) dk d\\varpi \n\\\\\n & =\nC_I^X\\left( \\int \\widehat W(k,\\varpi)\\psi_{k,\\varpi} \nC_{I^c}^Y(\\overline{\\psi_{k,\\varpi}}) dk d\\varpi \\right)\n\\\\\n & =\nC_I^X\\left( w^{C_{I^c}^Y}\\right) = \\int_{-T}^{T}\nC_{I^c}^Y(W(\\cdot - X_s,\\cdot - s)) dX_s,\n \\end{split}\n\\end{equation*}\nwith\n$$\nw^C(x,t) = \\int \\widehat W(k,\\varpi)\\psi_{k,\\varpi}(t,x)\nC(\\overline{\\psi_{k,\\varpi}}) dk d\\varpi = \nC(W(x-\\cdot,t-\\cdot)).\n$$\nBy using these equalities it is seen that the \nspecification~(\\ref{eq:spec}) takes the form\n\\begin{equation}\n\\label{eq:spec2}\n\\rho^\\sharp_I(d\\mathbb{X}|\\mathbb{Y}) =\n\\frac{e^{-V_I(X)-\\lambda W_I(X)-\n\\lambda \\int_I w^{C_{I^c}^Y}(u,X_u)dX_u}}{Z_I(\\mathbb{Y})} \n\\mathop{\\mathrm{{\\cal W}}}\\nolimits^\\sharp_I(d\\mathbb{X} | \\mathbb{Y}) \n\\end{equation}\nand it is well defined as soon as the exponential weight is integrable\nand the integral is different from zero. The conditions on $V$ and $W$ \nmake sure this is true. Indeed, for Kato-class potentials exponential \nintegrability is a consequence of Khasminskii's Lemma \\cite{Sim82}. \nMoreover, since the Fourier transform of $W$ is positive by assumption \nand $\\lambda >0$, the internal energy term is negative and thus \nexponentially integrable without any further restriction. For the \ninteraction with the boundary current we have\n$$\n|w^{C_{I^c}^Y}(x,t)| = |C_{I^c}^Y(W(x,t))| \\le M_I \\|C_{I^c}^Y\\|_{\\D'} \n$$\nwith $M_I = \\sup_{x\\in \\mathbb{R}^d,t\\in I} \\|W(x,t)\\|_{\\D}$, which by\ncondition (2) on $W$ is finite. Hence the stochastic integral in the \nexponent has a bounded and adapted integrand and thus by standard \ntechniques it follows that it is exponentially integrable for any\nvalue of $\\lambda$.\n\n\n\\medskip\nBy making use of the It\\^o current defined in\nSection~\\ref{sec:ito-current}, the specification $\\{\\rho^\\sharp_I\\}_I$ \ncan be finally written as\n\\begin{equation}\n\\label{eq:specc2}\n\\rho^\\sharp_I(d\\mathbb{X}|\\mathbb{Y}) =\n\\frac{e^{-\\lambda W_I(X)-\n\\lambda \\int_{-T}^T w^{C_{I^c}^Y}(u,X_u)dX_u}}{\\mathcal Z_I(\\mathbb{Y})} \n\\nu^\\sharp_I(d\\mathbb{X} | \\mathbb{Y}) . \n\\end{equation}\nNote that this is a forward current on $\\Xi$ but by the above results it \ncan be obtained as the unique lift of its marginal on $\\mathcal{X}$ satisfying the\nidentification~(\\ref{eq:curr-constr-0}) between currents and stochastic\nintegrals. \n\n\\begin{remark}\n\\rm{\nThe specification~(\\ref{eq:spec2}) seems to depend only on the path\nand the currents appearing in the definition of the vector-field \n$w^C$ that describes the interaction with the boundary paths. The point \nof introducing measures and specifications on currents resides in the \nfact that we are not able to describe (\\ref{eq:spec2}) in terms of \npaths alone. The framework of stochastic currents is not the only\npossibility to solve this difficulty. A different way to proceed is \nconsidering directly rough paths and defining the measures and \nspecifications on the space of (step-2) rough paths, i.e., formally of\ncouples $(X,\\mathbb{X}^2)$, where $\\mathbb{X}^2$ is the twice iterated\nintegral associated with the paths $X$. This would solve the problem\nof stochastic integrals, which can then be defined as rough integrals, \nand with suitable growth conditions on the rough paths we would be \nalso allowed to define the interaction terms with boundary paths (over\nunbounded time intervals) and specifications similarly to that on the \ncurrents. Our approach is motivated essentially by the consideration \nthat currents are more basic objects than rough paths. We prefer to\nsee rough path theory as a tool for obtaining stochastic currents in\nuseful topologies. Indeed, in principle the construction of good \nversions of stochastic integrals can be carried out without recourse \nto rough paths~\\cite{FGGT,FGR}. \n}\n\\end{remark}\n\n\n\n\\section{Existence of Gibbs measures for Brownian currents}\n\\label{sect:cluster}\n\n\\subsection{Cluster representation}\n\\label{sec:cluster-rep}\n\nIn the following we will construct a Gibbs measure that is consistent\nwith the specification $\\{\\rho^\\sharp_I\\}_I$. This will be achieved \nby breaking up paths according to a sequence of bounded subintervals \nof the real line, and constructing Gibbs measures for bounded intervals. \nTaking limits over these Gibbs measures will result in a Gibbs measure \non $\\mathcal{X}$ whose lifted measure to $\\Xi$ is consistent with the given\nspecification. As mentioned before, a reasonably confining $V$ is \nneeded to make sure that the paths are not allowed to escape to \ninfinity with large probability. \n\nThe following notion of convergence will be used below to discuss\nGibbs measures. Let generally $E$ \nbe a metric space, and $C(\\mathbb{R},E)$ the space of continuous paths\n$\\{X_t; t \\in \\mathbb{R}\\}$ with values in $E$. For any bounded interval $I \n\\subset \\mathbb{R}$ let $\\mathop{\\mathrm{{\\cal E}}}\\nolimits_I \\subset \\mathop{\\mathrm{{\\cal E}}}\\nolimits$ be a sub-$\\sigma$-field of the Borel \n$\\sigma$-field $\\mathop{\\mathrm{{\\cal E}}}\\nolimits$ of $E$ generated by the evaluations $\\{X_t: t \\in I\\}$. \nA sequence of probability measures $(m_n)_{\\{n \\in \\mathbb{N}\\}}$ on \n$C(\\mathbb{R},E)$ is said to converge locally weakly to the probability measure \n$m$ if for any such $I$ the restrictions $m_n|_{\\mathop{\\mathrm{{\\cal E}}}\\nolimits_I}$ converge weakly \nto the measure $m|_{\\mathop{\\mathrm{{\\cal E}}}\\nolimits_I}$. \n\nThe main result of this paper is the following\n\n\\begin{theorem}\nSuppose $V$ and $W$ satisfy the assumptions stated in Section \n\\ref{sec:potentials}. Take any unbounded increasing sequence \n$(T_n)_{n\\geq 0}$ of positive real numbers, and suppose $0 < |\\lambda| \n\\leq \\lambda^*$ with $\\lambda^*$ small enough. Then the local weak \nlimit $\\lim_{n \\rightarrow \\infty} \\mu_{T_n} = \\mu$ exists on $\\mathcal{X}$ \nand does not depend on the choice of sequence $T_n$. Its unique lift \n$\\mu^\\sharp$ on $\\Xi$ is a Gibbs probability measure consistent with\nthe specification $\\{\\rho^\\sharp_I\\}_I$.\n\\label{exist}\n\\end{theorem}\n\\begin{proof}\nWe develop a cluster expansion, i.e., choose the coupling parameter \n$\\lambda$ sufficiently small for being able to control the measure \nfor the interaction switched on ($\\lambda \\neq 0$) in terms of a \nconvergent perturbation series around the free case ($\\lambda = 0$). \nThe theorem follows then through Propositions \\ref{p0}, \\ref{p2},\n\\ref{pclustest} and \\ref{prop:infinite-measure} below. \n\\end{proof}\n\n\\medskip\nTake a division of $[-T,T]$ into disjoint intervals $\\tau_k = (t_k, \nt_{k+1})$, $k = 0,..., N-1$, with $t_0 = -T$ and $t_N = T$, each of \nlength $b$, i.e. fix $b = 2T\/N$; for convenience we choose $N$ to be \nan even number so that the origin is endpoint to some intervals. We \nbreak up a path $X$ into pieces $X_{\\tau_k}$ by restricting it to \n$\\tau_k$. The total energy contribution of the pair interaction then\nbecomes\n\\begin{equation}\nW_T(X) = \\frac{1}{2} \\sum_{i,j = 0}^{N-1} \n\\langle C^X_{\\tau_i},C^X_{\\tau_j}\\rangle_W = \n\\sum_{0 \\leq i < j \\leq N-1} W_{\\tau_i,\\tau_j}\n\\label{sum}\n\\end{equation}\nwhere with the notation $\\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ij} = \n\\langle C^X_{\\tau_i},C^X_{\\tau_j}\\rangle_W$ we have\n\\begin{equation}\nW_{\\tau_i,\\tau_j} \n= \\left\\{ \n\\begin{array}{ll}\n\\vspace{0.2cm} \n\\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ij} + \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ji} & \\mbox{if $|i-j| \\geq 2$} \\nonumber \\\\\n\\vspace{0.2cm} \n\\frac{1}{2} (\\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ii} + \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{jj}) + \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ij} + \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ji} \n& \\mbox{if $|i-j| = 1$, and \n$i \\neq 0$, $j \\neq N - 1$} \\nonumber \\\\\n\\vspace{0.2cm} \n\\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ij} + \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ji} + \\frac{1}{2} \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{00} \n& \\mbox{if $i = 0$ and $j = 1$} \\nonumber \\\\ \n\\vspace{0.2cm} \n\\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ij} + \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ji} + \\frac{1}{2} \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{N-1 \\; N-1} \n& \\mbox{if $i = N-1$ and $j = N-2$.} \\nonumber \\\\\n\\end{array} \\right.\n\\end{equation}\nTo keep the notation simple we do not make explicit the $X$ \ndependence in these objects.\n\nBy using (\\ref{sum}) we obtain\n\\begin{equation}\ne^{-\\lambda W_T} = \\prod_{0 \\leq i < j \\leq N-1} \n(e^{-\\lambda W_{\\tau_i,\\tau_j}} + 1 - 1) = 1 + \\sum_{{\\cal R} \\neq \n\\emptyset} \\prod_{(\\tau_i,\\tau_j) \\in {\\cal R}} \n(e^{-\\lambda W_{\\tau_i,\\tau_j}} - 1).\n\\label{sum1}\n\\end{equation}\nHere the summation is performed over all nonempty sets of different \npairs of intervals, i.e. ${\\cal R} = \\{(\\tau_i,\\tau_j): \n(\\tau_i, \\tau_j) \\neq (\\tau_{i'},\\tau_{j'}) \\; \\mbox{whenever} \\; (i,j) \n\\neq (i',j') \\}$. \n\nA break-up of the paths involves a corresponding factorization of the \nreference measure into It\\^o bridges for each subinterval. Put $X_{t_k} = \nx_k$ for the positions at the time-points of the division, $\\forall \nk = 0,...,N$, with $-T = t_0 < t_1 < ... < t_N = T$. We write for a \nshorthand\n\\begin{equation}\n\\nu_T^{\\mathbf{x}}(\\cdot) \\equiv \\nu_T(\\, \\cdot \\,| X_{t_0} = x_0, \\ldots , \nX_{t_N} = x_N) = \\prod_{k=0}^{N-1} d\\nu_{\\tau_k}^{x_k,x_{k+1}}(\\cdot).\n\\label{mar}\n\\end{equation}\nLet $p_{t_0,...,t_N} (x_0,...,x_N)$ be the density with respect to \n$\\prod_{k=0}^N d\\omega(x_k)$, $d\\omega = \\Psi^2 dx$, of the joint \ndistribution of positions of the path $X$ recorded at the time-points \nof the division. \nBy Markovianness it follows that\n\\begin{eqnarray}\np_{t_0,...,t_N} (x_0,...,x_N) \n& = &\n\\prod_{k=0}^{N-1} \\pi_b(x_{k+1},x_k) = \\prod_{k=0}^{N-1} \n(\\pi_b(x_{k+1},x_k) - 1 + 1) \n\\nonumber \\\\ \n& = &\n1 + \\sum_{\\cal S \\neq \\emptyset} \\prod_{k: \\tau_k \\in {\\cal S}} \n(\\pi_b(x_{k+1},x_k) - 1),\n\\nonumber\n\\end{eqnarray}\nwhere $\\pi_t$ is the transition kernel for the It\\^o diffusion given by\n(\\ref{pito}). The summation runs over all nonempty sets ${\\cal S} = \n\\{\\tau_k = (t_k,t_{k+1})\\}$ of different pairs of consecutive time-points. \n\nIn order to have a systematic control over these sums we introduce:\n\n\\begin{itemize}\n\\item[(1)] \n \\emph{Contours.} \\hspace{0.2cm} \nTwo distinct pairs of intervals $(\\tau_i,\\tau_j)$ and \n$(\\tau_{i'},\\tau_{j'})$ will be called directly connected and denoted \n$(\\tau_i,\\tau_j) \\sim (\\tau_{i'},\\tau_{j'})$ if one interval of the pair \n$(\\tau_i,\\tau_j)$ coincides with one interval of the pair $(\\tau_{i'},\n\\tau_{j'})$. A set of connected pairs of intervals is a collection \n$\\{(\\tau_{i_1},\\tau_{j_1}),..., (\\tau_{i_n},\\tau_{j_n})\\}$ in which \neach pair of intervals is connected to another through a sequence of \ndirectly connected pairs, i.e., for any $(\\tau_i,\\tau_j) \\neq \n(\\tau_{i'},\\tau_{j'})$ there exists $\\{(\\tau_{k_1},\\tau_{l_1}),..., \n(\\tau_{k_m},\\tau_{l_m})\\}$ such that $(\\tau_i,\\tau_j) \\sim \n(\\tau_{k_1},\\tau_{l_1}) \\sim ... \\sim (\\tau_{k_m},\\tau_{l_m}) \n\\sim (\\tau_{i'},\\tau_{j'})$. A maximal set of connected pairs of \nintervals is called a contour, denoted by $\\gamma$. We denote by \n$\\bar\\gamma$ the set of all intervals that are elements of the pairs \nof intervals belonging to contour $\\gamma$, and by $\\ga^*$ the set of \ntime-points of intervals appearing in $\\bar\\ga$. Two contours $\\ga_1, \n\\ga_2$ are disjoint if they have no intervals in common, i.e. \n$\\bar \\ga_1 \\cap \\bar \\ga_2 = \\emptyset$. Clearly, $\\cal R$ can \nbe decomposed into sets of pairwise disjoint contours: \n${\\cal R} = \\cup_{r \\geq 1} {\\cal R}_r$, where ${\\cal R}_r = \n\\{\\ga_1,...,\\ga_r\\}$ with $\\bar\\ga_i \\cap \\bar \\ga_j = \\emptyset$, \n$i \\neq j$; $i,j = 1,...,r$. \n\n\\item[(2)] \\emph{Chains.} \\hspace{0.2cm} \nA collection of consecutive intervals $\\{\\tau_{j},\\tau_{j+1}...,\n\\tau_{j+k} \\}$, $j \\geq 0$, $j+k \\leq N-1$ is called a chain. \nAs in the case of contours, $\\bar \\rh$ and $\\rh^*$ mean the set of \nintervals belonging to the chain $\\rh$ and the set of time-points in \n$\\rh$, respectively. Two chains $\\rh_1, \\rh_2$ are called disjoint if \nthey have no common time-points, i.e. $\\rh_1^* \\cap \\rh^*_2 = \\emptyset$. \nDenote by $\\partial^-\\rh$ resp. $\\partial^+\\rh$ the leftmost resp. \nrightmost time-points belonging to $\\rh$. \n\n\\item[(3)]\n \\emph{Clusters.} \\hspace{0.2cm} \nTake a (non-ordered) set of disjoint contours and disjoint chains, \n$\\Ga = \\{\\gamma_1,...,\\gamma_r;\\varrho_1,...,\\varrho_s\\}$, with some \n$r \\geq 1$ and $s \\geq 0$. Note that such contours and chains may have \ncommon time-points. The notation $\\Ga^* = (\\cup_i \\ga^*_i) \\cup \n(\\cup_j \\rh_j^*)$ means the set of all time-points appearing as\nbeginnings or ends of intervals belonging to some contour or chain \nin $\\Gamma$. Also, we put $\\bar \\Ga = (\\cup_i \\bar\\ga_i) \\cup (\\cup_j \n\\bar\\rh_j)$ for the set of intervals appearing in $\\Ga$ through \nentering some contours or chains. $\\Gamma$ is called a cluster if \n$\\Ga^*$ is a connected collection of sets (in the usual sense), and \nfor every $\\varrho \\in \\Gamma$ we have that $\\partial^-\\rh, \n\\partial^+\\rh \\in \\cup_{j=1}^r\\ga^*_j$. This means that in a cluster \nchains have no loose ends. We denote by $\\mathop{\\mathrm{{\\cal K}}}\\nolimits_N$ the set of all clusters \nfor a given $N$. \n\n\\end{itemize}\n\nWith these notations the sum in (\\ref{sum1}) is then further expanded as\n\\begin{equation}\n\\sum_{{\\cal R} \\neq \\emptyset} \\prod_{(\\tau_i,\\tau_j) \\in \\cal R} \n(e^{-\\lambda W_{\\tau_i,\\tau_j}} - 1) = \\sum_{r \\geq 1} \n\\sum_{\\{\\gamma_1,...,\\gamma_r\\}} \\prod_{k = 1}^r \\prod_{(\\tau_i,\\tau_j) \n\\in \\gamma_k} (e^{-\\lambda W_{\\tau_i,\\tau_j}} - 1)\n\\label{sum2}\n\\end{equation}\nwhere now summation goes over collections $\\{\\gamma_1,...,\\gamma_r\\}$ \nof contours such that $\\bar\\ga_k \\cap \\bar\\ga_{k'} = \\emptyset$ \nunless $k = k'$. In a similar way (\\ref{mar}) appears in the form\n\\begin{equation}\n\\sum_{\\cal S \\neq \\emptyset} \\prod_{k: \\tau_k \\in {\\cal S}} \n\\left( \\pi_b(x_{k+1}, x_k) - 1 \\right) = \n\\sum_{s \\geq 1} \\sum_{\\{\\varrho_1,...,\\varrho_s\\}} \\prod_{j=1}^s \n\\prod_{k: \\tau_k \\in \\varrho_j} \\left(\\pi_b(x_{k+1},x_k) - 1 \\right).\n\\label{sum3}\n\\end{equation}\nHere $\\{\\varrho_1,...,\\varrho_s\\}$ is a collection of disjoint chains,\nand this formula justifies how we defined them. \n\nFor every cluster $\\Ga = \\{\\ga_1,...,\\ga_r;\\rh_1,...,\\rh_s\\} \\in \\mathop{\\mathrm{{\\cal K}}}\\nolimits_N$ \ndefine the function\n\\begin{equation}\n\\kappa_\\Ga = \\prod_{l=1}^r \\prod_{(\\tau_i,\\tau_j) \\in \\gamma_l} \n(e^{-\\lambda W_{\\tau_i,\\tau_j}} - 1) \\prod_{m=1}^s \n\\prod_{k: \\tau_k \\in \\rh_m}\\left(\\pi_b(x_{k+1},x_k) - 1\\right). \n\\label{kappa}\n\\end{equation}\nAlso, introduce the auxiliary probability measure \n\\begin{equation}\nd\\chi_N(X) = \\prod_{k=0}^{N-1} d\\nu_{\\tau_k}^{x_k,x_{k+1}}(X_{\\tau_k}) \n\\prod_{k=0}^N d\\omega(x_k),\n\\label{auxi}\n\\end{equation}\nand look at \n\\begin{equation}\nK_{\\Ga} = \\mathbb{E}_{\\chi} [\\kappa_\\Ga],\n\\label{aux}\n\\end{equation}\nwhere $\\chi$ is the unique extension over the real line of the family \nof consistent probabilities $\\{\\chi_N\\}_{N\\ge 1}$. Note that \n$\\int(\\pi_b(x_{k+1},x_k)-1) d\\omega(x_{k+1}) = \\int(\\pi_b(x_{k+1},x_k) \n-1) d\\omega(x_k) = 0$. This is the reason why from a cluster we rule \nout chains having loose ends; for any such chain $\\mathbb{E}_{\\chi_N} \n[\\kappa_\\Ga] = 0$.\n\nDefine \n\\begin{equation*}\n\\phi^T(\\Ga_1,...,\\Ga_n) = \\begin{cases}\n1 & \\text{if $n=1$}\\\\\n\\sum_{G \\in \\G^n} \\prod_{\\{i,j\\}\\in G} \n(- 1_{\\Ga^*_i \\cap \\Ga^*_j \\neq \\emptyset}) & \\text{if $n>1$},\n\\end{cases}\n\\end{equation*}\nwith $\\G^n$ the set of connected graphs on the vertex set\n$\\{1,\\dots,n\\}$. Note that $\\phi^T(\\Ga_1,...,\\Ga_n)=0$ if the graph \non the vertex set $\\{1,\\dots,n\\}$ with edges $\\{i,j\\}$ drawn \nwhenever $\\Gamma^*_i \\cup \\Gamma^*_j \\neq \\emptyset$, is connected.\n\nBy putting (\\ref{sum2}), (\\ref{mar}), (\\ref{sum3}), (\\ref{kappa}) and \n(\\ref{aux}) together we obtain the cluster representation of the\npartition function $Z_T$.\n\\begin{proposition}\n\\label{p0}\nFor every $T = Nb\/2 > 0$ we have \n\\begin{equation}\n\\label{clexp}\nZ_T = 1 + \\sum_{n \\geq 1} \\sum_{\\{\\Ga_1,...,\\Ga_n\\} \\in \\mathop{\\mathrm{{\\cal K}}}\\nolimits_N\n\\atop \\Ga^*_i \\cap \\Ga^*_j = \\emptyset, i \\neq j} \\prod_{l=1}^n \nK_{\\Ga_l}.\n\\end{equation}\nIf the activities $K_{\\Ga}$ satisfy the bound\n\\begin{equation}\n\\sum_{\\Ga \\in \\mathop{\\mathrm{{\\cal K}}}\\nolimits_N \\atop \\Ga^* \\ni 0, |\\bar\\Ga| = n} |K_{\\Ga}| \n\\;\\leq\\; c \\; \\eta^n\n\\label{conv}\n\\end{equation}\nfor $\\eta > 0$ small enough, then the series above and at the right \nhand side of\n\\begin{equation}\n\\log Z_T = \\sum_{n \\geq 1} \\sum_{\\{\\Ga_1,...,\\Ga_n\\} \\in \\mathop{\\mathrm{{\\cal K}}}\\nolimits_N \n\\atop 0 \\in \\Ga_1^*} \\phi^T(\\Ga_1,...,\\Ga_n) \\prod_{l=1}^n K_{\\Ga_l}\n\\end{equation}\nare absolutely convergent, uniformly in $N$, and the latter one \ngives the logarithm of the partition function for the interval \n$[-T,T]$.\n\\end{proposition}\n\n\\noindent\nThe expression of the logarithm and the absolute convergence of the \nsums are a general result of cluster expansion techniques, for details \nof proof see~\\cite{MM}.\n\n\n\\subsection{Convergence of cluster expansion}\n\n\\begin{proposition}\n\\label{p2}\nSuppose that there exist a function $D: \\mathbb{N}\\times\\mathbb{N} \n\\rightarrow (0,\\infty)$ and numbers $\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits, C>0$ with\n$\n\\sup_{i\\in\\mathbb{N}} \\sum_{j\\in\\mathbb{N}} D(i,j) \\le C\n$\nsuch that for every $N > 0$ and every cluster \n$\\Ga = \\{\\ga_1,...,\\ga_r;\\rh_1,...,\\rh_s\\} \\in \\mathop{\\mathrm{{\\cal K}}}\\nolimits_N$, the bound \n\\begin{equation}\n|K_{\\Gamma}| \\; \\leq \\; \\left(\\prod_{l=1}^r \\prod_{(\\tau_i,\\tau_j) \n\\in \\gamma_l} \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits D(i,j)\\right) \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits^{\\sum_{m=1}^s |\\bar\\rh_m|}\n\\label{clustest-abs}\n\\end{equation}\nholds. Then there is a constant $c > 0$ and a function \n$0 < \\eta(\\varepsilon) < 1$ with $\\eta \\to 0$ as $\\varepsilon \\to 0$ such that\n\\begin{equation}\n\\sum_{\\Ga \\in \\mathop{\\mathrm{{\\cal K}}}\\nolimits_N \\atop \\Ga^* \\ni 0, |\\bar\\Ga| = n} |K_{\\Ga}| \n\\;\\leq\\; c \\; \\eta^n.\n\\label{convv}\n\\end{equation}\n\\end{proposition}\nThe function $D(i,j)$ will be specified in Proposition \n\\ref{pclustest} below. \n\n\\medskip\n\\begin{proof}\nWe put for a shorthand \n$\\D(\\ga) = \\prod_{(\\tau_i,\\tau_j) \\in \\ga} \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits D(i,j)$.\nConsider the function of complex variable $z$\n\\begin{equation}\nH(z;\\varepsilon) = \\sum_{\\Ga\\in\\mathop{\\mathrm{{\\cal K}}}\\nolimits_N: \\Ga^* \\ni 0} K_{\\Ga} z^{|\\bar \\Ga|} = \n\\sum_{\\Ga\\in\\mathop{\\mathrm{{\\cal K}}}\\nolimits_N: \\Ga^* \\ni 0 \\atop \\Ga \\supset\\; \\mbox{\\tiny{one contour}}} \nK_{\\Ga} z^{|\\bar\\Ga|} + \\sum_{\\Ga\\in\\mathop{\\mathrm{{\\cal K}}}\\nolimits_N: \\Ga^* \\ni 0 \\atop \\Ga \\supset\\;\n\\mbox{\\tiny{more than one contour}}}K_{\\Ga} z^{|\\bar\\Ga|}\n\\label{H}\n\\end{equation}\nWe show that for sufficiently small $\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits > 0$ this is an analytic \nfunction of $z$ in a circle of radius $R(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)$ which diverges as \n$\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits \\to 0$. Moreover, we show that within this circle $H(z;\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)$ \nis uniformly bounded in $\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits$. This will then imply (\\ref{convv}) by \nchoosing $\\eta(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits) = 1\/R(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)$. \n\nWe start by estimating the second sum; the first is simpler as it\ninvolves clusters having a single contour. Our strategy is first to\nbound it by a sum taken over graphs whose vertices are the contours of\nthe same cluster. The sums over graphs will then be bounded by sums\ntaken over trees.\n\n\\medskip\n\\noindent\n\\emph{\\textbf{Bounds by sums over graphs}} \\quad\nFor each $r \\geq 2$ consider in $\\mathop{\\mathrm{{\\cal K}}}\\nolimits_N$ those clusters $\\Ga$ that have $r$ \ncontours. For given $\\Ga \\in \\mathop{\\mathrm{{\\cal K}}}\\nolimits_N$ let $\\V_r = \\{\\ga_1,...,\\ga_r\\}$ be the \ncollection of these contours. We construct connected graphs $G$ by drawing \nedges between the elements of $\\V_r$ considered as vertex set. \nConnected graphs are those for which either $\\ga_i^* \\cap \\ga_j^* \\neq \\emptyset$ \nor there exists $\\rh_l \\in \\Ga$ such that $\\rh^*_l \\cap \\ga_i^* \\neq \\emptyset \n\\neq \\rh_l^* \\cap \\ga_j^*$. Let $\\G_r$ denote the set of all possible such graphs. \n\nConsider the collection of {\\it reduced} chains\n$\\{\\hat\\rh_1,...,\\hat\\rh_s\\}$ with the properties:\n\\begin{enumerate}\n\\item[(1)]\nfor every pair $\\{\\ga_i,\\ga_j\\} \\in G \\in \\G_r$, $\\ga_i^* \\cap \\ga_j^* = \\emptyset$, \nthere is at least one chain $\\hat\\rh_l$ of this collection connecting $\\ga_i$ and \n$\\ga_j$ (i.e. $\\overline{\\hat \\rh_l} \\cap \\bar \\ga_i = \\emptyset = \\overline{\\hat \n\\rh_l} \\cap \\bar \\ga_j$ and $\\hat \\rh_l^* \\cap \\ga_i^* \\neq \\emptyset \\neq \\hat \n\\rh_l^* \\cap \\ga_j^*$), and for any pair $\\{\\ga_i,\\ga_j\\} \\not \\in G$ such a chain \ndoes not occur;\n\\item[(2)]\n$\\cup_i \\bar\\ga_i \\cap \\cup_j \\overline{\\hat\\rh}_j = \\emptyset$ and \n$\\cup_j \\{\\partial^-{{\\hat\\rh}_j},\\partial^+ {{\\hat\\rh}_j}\\} \\subset \n\\cup_i \\ga_i^*$; \n\\item[(3)]\n$0 \\in (\\cup_i \\ga_i^*) \\cup (\\cup_j {\\hat\\rh}_j^*)$;\n\\item[(4)]\neach chain $\\hat\\rh_k$ connects a pair $\\{\\ga_i,\\ga_j\\} \\in G \\in\n\\G_r$ \nor fills a gap within a contour $\\ga_i$.\n\\end{enumerate}\nWe call a collection of reduced chains compatible with graph $G$ if it\nsatisfies the conditions above and denote it \n$\\{\\hat\\rh_1,...,\\hat\\rh_s\\}_G$. Note that each $\\hat\\rh$ can join\nonly one pair of contours. A collection of reduced chains is then\nconstructed through the following steps:\n\\begin{enumerate}\n\\item[(1)]\nfirst remove all chains $\\rh_l \\in \\Ga$ for which $\\bar\\rh_l \\subset \\cup_{i=1}^r \n\\bar\\ga_i$;\n\\item[(2)]\nfor all remaining chains $\\rh_k \\in \\Ga$, $\\bar \\rh_k \\not\\subset \\cup_{i=1}^r \n\\bar\\ga_i$ remove all intervals from the set $\\bar\\rh_k \\cap (\\cup_{i=1}^r \\bar\\ga_i)$;\n\\item[(3)]\nof the remaining intervals form all possible collections of non-empty chains denoted \nby $\\{\\hat\\rh_1,...,\\hat\\rh_s\\}$.\n\\end{enumerate} \nThen by Proposition \\ref{clustest} we write\n\\begin{eqnarray}\n\\label{onecont}\n\\lefteqn{\n\\big|\\sum_{\\Ga: \\Ga^* \\ni 0 \\atop \\Ga \\supset\\;\n\\mbox{\\tiny{more than one contour}}} K_\\Ga z^{|\\bar\\Ga|}\\big| \n\\; \\leq \\label{est} } \\\\ &&\n\\sum_{r\\geq2} \\sum_{\\{\\ga_1,...,\\ga_r\\}} \\sum_{G \\in \\G_r} \n\\sum_{s \\geq 0} \\prod_{i=1}^r ((|z|(1+\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|))^{|\\bar\\ga_i|} \n\\D(\\ga_i) \\sum_{\\{\\hat\\rh_1,...,\\hat\\rh_s\\}_G \\atop \n\\mbox{\\tiny {$0 \\in \\cup_i \\gamma_i^* \\cup {\\hat\\rh}_i^*$}}} \n\\prod_{j=1}^s(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{|\\overline {\\hat\\rh_j}|}. \\nonumber\n\\end{eqnarray}\nNote that for fixed $\\{\\ga_1,...,\\ga_r\\}$ the collection\n$\\{\\hat\\rh_1,...,\\hat\\rh_m\\}$ can be obtained from many possible\ncollections of chains $\\{\\rh_1,...,\\rh_s\\}$. This gives the factor \n$(1 + \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{|\\bar\\ga_i|}$ appearing at the right hand side of \n(\\ref{est}). From now on we assume that $\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|\\le 1$ so that we \ncan estimate this factor by $2^{|\\bar\\ga_i|}$.\n\n\nNow consider the last sum above involving the reduced chains.\nIn this sum, either $0$ belongs to a contour or some chains. In\nthe second case there is a factor of $(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(0,\\{\\gamma\\})}$\nappearing in the sum, so we can estimate the sum by\n\\begin{equation*}\n \\begin{split}\n& (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(0,{\\gamma})\/2} \\sum_{\\{\\hat\\rh_1,...,\\hat\\rh_s\\}_G \n\\atop \\mbox{\\tiny {$0 \\in \\cup_i \\gamma_i^* \\cup {\\hat\\rh}_i^*$}}}\n\\prod_{j=1}^s (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{|\\overline {\\hat\\rh_j}|\/2}\n \\le \n \\sum_k (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(0,\\gamma_k)\/2}\n \\sum_{\\{\\hat\\rh_1,...,\\hat\\rh_s\\}_G \n\\atop \\mbox{\\tiny {$0 \\in \\cup_i \\gamma_i^* \\cup {\\hat\\rh}_i^*$}}}\n\\prod_{j=1}^s (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{|\\overline {\\hat\\rh_j}|\/2}.\n \\end{split}\n\\end{equation*}\nFor each contour $\\gamma_i$ the sum over all reduced chains belonging\nto this contour cannot be larger than $2^{|\\overline{\\gamma_i}|}$\nsince the number of such chains is bounded by $\\overline{\\gamma_i}$\n(when every chain separates each two successive intervals in the\ncontour). Moreover for each couple of contours $(\\gamma_i,\\gamma_j)$\nthe contribution to the sum of the chains connecting them is given by\n$$\n2^{|\\overline{\\gamma_i}|+|\\overline{\\gamma_j}|} \n(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\gamma_i,\\gamma_j)\/2}\n$$\nsince there is at least one chain longer than\n$\\mathop{\\mathrm{dist}}\\nolimits(\\gamma_i,\\gamma_j)$ and the rest of the chains contribute\ninto the combinatorial prefactor. This gives\n\\begin{eqnarray}\n\\label{morecont-2}\n\\lefteqn{\n\\big|\\sum_{\\Ga: \\Ga^* \\ni 0 \\atop \\Ga \\supset\\;\n\\mbox{\\tiny{more than one contour}}}\nK_\\Ga z^{|\\bar\\Ga|}\\big| \\; \\leq \\label{estt} } \\\\ &&\n\\sum_{r\\geq2} \\sum_{\\{\\ga_1,...,\\ga_r\\}}\n \\sum_k \\prod_{i=1}^r (8|z|)^{|\\bar\\ga_i|} \\D(\\ga_i) \n(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(0,\\gamma_k)\/2} \\sum_{G \\in \\G_r}\n \\prod_{\\{\\gamma_i,\\gamma_j\\}\\in G} \n(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\gamma_i,\\gamma_j)\/2}.\n \\nonumber\n\\end{eqnarray}\n\n\\medskip\n\\noindent\n\\emph{\\textbf{Bounds by sums over trees}} \\quad\nWe use the tree-graph bound (cf. Lemma 8, Ch. 2, Sect. 4 \nof \\cite{MM}) to get\n\\begin{equation}\n\\sum_{G\\in \\G_r} \\prod_{\\{\\ga_i,\\ga_j\\} \\in G} (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^\n{\\mathop{\\mathrm{dist}}\\nolimits (\\gamma_i,\\gamma_j)\/2} \n\\leq \n\\prod_{i,j =1}^r (1 + (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\gamma_i,\\gamma_j)\/2}) \n\\sum_{T \\in \\T_r} \\prod_{\\{\\ga_i,\\ga_j\\} \\in T} \n(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\gamma_i,\\gamma_j)\/2}\n\\label{mm8}\n\\end{equation}\nwhere $\\T_r$ is the set of trees on the vertex set\n$\\{\\gamma_1,\\dots,\\gamma_r\\}$. Moreover we have\n$$\n\\prod_{i,j =1}^r (1 + (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\gamma_i,\\gamma_j)\/2}) \n\\le\n2^{2 r} \\le 2^{2\\sum_i |\\overline\\gamma_i| },\n$$\nthus (\\ref{onecont}) is further estimated by\n\\begin{eqnarray}\n\\label{morecont-3}\n\\lefteqn{\n\\big|\\sum_{\\Ga: \\Ga^* \\ni 0 \\atop \\Ga \\supset\\;\n\\mbox{\\tiny{more than one contour}}}\nK_\\Ga z^{|\\bar\\Ga|}\\big| \\; \\leq \\label{esttt} } \\\\ &&\n\\sum_{r\\geq2} \\sum_{\\{\\ga_1,...,\\ga_r\\}}\n \\sum_k \\prod_{i=1}^r (32|z|)^{|\\bar\\ga_i|} \\D(\\ga_i) \n(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(0,\\gamma_k)\/2} \\sum_{T \\in \\T_r}\n \\prod_{\\{\\gamma_i,\\gamma_j\\}\\in G} (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits\n(\\gamma_i,\\gamma_j)\/2}.\n\\nonumber\n\\end{eqnarray}\n\n\nTake the trees over vertex set $\\{1,...,r\\}$ obtained through \n$\\ga_k \\mapsto k$, $\\forall k = 1,...,r$; denote them $\\tilde T$ \nand the set of all such trees by $\\tilde \\T_r$. Then we re-sum in\n(\\ref{morecont-3}):\n\\begin{equation}\n \\begin{split}\n& \\mbox{r.h.s.}\\; (\\ref{morecont-3}) \\; \\leq\\\\\n& \\quad \\sum_{r=2}^\\infty \\frac{1}{r!} \n\\sum_{\\tilde T \\in \\tilde \\T_r} \n\\sum_{i^*=1}^r \n\\sum_{(\\ga_1,...,\\ga_r)} \\prod_{i=1}^r (32|z|)^{|\\bar\\ga_i|} \n(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(0,\\ga_{i^*})\/2} \\D(\\ga_i) \n\\prod_{\\{i,j\\} \\in \\tilde T}\n(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\ga_i,\\ga_j)\/2}. \n \\end{split}\n\\label{hae}\n\\end{equation}\nThe change of bracket indicates that the third sum here is performed \nover all ordered collections of disjoint contours. Fix an enumeration\nof $\\tilde T$ and pick its first element $i_0$. We estimate first \n\\begin{equation}\n\\sum_{(\\ga_1,...,\\ga_r)} \\prod_{i=1}^r (32|z|)^{|\\bar\\ga_i|} \\D(\\ga_i) \n\\prod_{\\{i,j\\} \\in \\tilde T} (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\ga_i,\\ga_j)\/2}.\n\\label{hva}\n\\end{equation}\nLet $j_0 \\neq i_0$ be an end vertex of tree $\\tilde T$ being joint \nonly with vertex $k_0$. Then\n\\begin{eqnarray}\n\\lefteqn{\n\\hspace{-1.5cm}\n\\sum_{\\ga_{j_0}} (32|z|)^{|\\bar\\ga_{j_0}|} \\D(\\ga_{j_0}) (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^\n{\\mathop{\\mathrm{dist}}\\nolimits(\\ga_{k_0},\\ga_{j_0})\/2} \\label{i0}} \\\\ &&\n\\;\\leq\\; \\sum_{\\tau'' \\in \\bar\\ga_{k_0}} \\sum_{\\ga_{j_0}} \n\\sum_{\\tau' \\in \\bar \\ga_{j_0}} (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\tau',\\tau'')\/2} \n(32|z|)^{|\\bar\\ga_{j_0}|} \\D(\\ga_{j_0}) \n\\nonumber \\\\ &&\n\\;\\leq\\; \\sum_{\\tau'' \\in \\bar\\ga_{k_0}} \\sum_{\\tau'} \n(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\tau',\\tau'')\/2} \\sum_{\\ga_{j_0}: \\tau' \\in \\bar \n\\ga_{j_0}} (32|z|)^{|\\bar\\ga_{j_0}|} \\D(\\ga_{j_0}). \\nonumber \n\\end{eqnarray}\nHere we used that $(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\ga,\\ga')\/2} \n\\leq \n\\sum_{\\tau \\in \\ga, \\tau' \\in \\ga'} (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\tau,\\tau')\/2}$.\nBy using Lemma \\ref{lsumd} below and the bound\n\\begin{equation}\n\\sum_{k=2}^\\infty a^{k-1} k^m \\leq \\frac{2^{m}m! e a}{1 - e a}\n\\label{geo}\n\\end{equation}\nobtained via complex integration, we further estimate (\\ref{i0}) by\n\\begin{equation}\n\\label{j0}\n\\sum_{\\tau'' \\in \\bar\\ga_{k_0}} \\sum_{\\tau'} \n(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\tau',\\tau'')\/2} \n\\sum_{k=2}^\\infty (32|z|)^k (C\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)^{k-1} \n\\leq \n\\frac{64 C |\\bar\\ga_{k_0}| \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|}{(1 - (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{1\/2})\n(1 - 32C \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)}. \n\\end{equation}\nFrom now on we choose $z$ such that $32C \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z| < 1$ holds.\n\nNext we go on by taking the next vertex of $\\tilde\\T$ in line, say\n$j_1 \\neq i_0$ connecting with $k_1$. We iterate the procedure for the\nnew tree obtained by deleting from $\\tilde T$ the vertex $j_0$ and\nedge $(j_0, k_0)$. If $j_1 \\neq k_0$, we get again an estimate of the\ntype (\\ref{j0}). If $j_1 = k_0$, we estimate\n\\begin{equation*}\n\\sum_{\\tau'' \\in \\ga_{k_1}} \\sum_{\\tau'} (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\tau',\\tau'')\/2} \n\\sum_{\\ga_{j_1}: \\tau' \\in \\ga}|\\bar\\ga_{j_1}|(32|z|)^{|\\bar\\ga_{j_1}|}\n\\D(\\ga)\n\\leq \n\\const{q2} |\\bar\\ga_{k_1}| \\sum_{k=2}^\\infty (32|z|)^k \nk (C \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)^{k-1},\n\\end{equation*}\nwith some $\\defconst{q2} > 0$. Continuing this procedure inductively \nwe get after summation over $\\ga_{j_m}$, $j_m \\neq i_0$, connected \nto $\\ga_{k_m}$, the net contribution\n\\begin{eqnarray*}\n{\\const{q2}|\\bar\\ga_{k_m}|} \\sum_{k=1}^\\infty (32|z|)^k k^{l_{j_m} -1} \n(C\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)^{k-1} \n&\\leq&\n{\\const{q2}|\\ga_{k_m}|} \\sum_{k=2}^\\infty (\\const{q3}|z|)^k \nk^{l_{j_m}-1} (C\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)^{k-1} \\\\\n&\\leq&\n\\const{q9} \\; |\\ga_{k_m}||z| \\sum_{k=2}^\\infty ({\\const{q5}} \n\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|)^{k-1} k^{l_{j_1}-1}\n\\label{ind}\n\\end{eqnarray*}\nwhere $l_{j_m}$ is the degree of vertex $j_m$, i.e. the number \nof edges of $\\tilde T$ incident to $j_m$, and $\\defconst{q3},\n\\defconst{q4} > 0$, $\\defconst{q5} = C\\const{q3}$, $\\defconst{q9} \n= \\const{q3}\\const{q4}$ is the long sequence of constants. By using \n(\\ref{geo}) again, we estimate (\\ref{hva}) further for fixed \n$\\ga_{i_0}$ and $\\tilde T$ to get\n\\begin{equation*}\n\\sum_{\\ga_k: \\; k \\neq i_0 \\atop k = 1,...,r} \\prod_{\\{i,j\\} \\in \\tilde T} \n(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(\\tau_i,\\tau_j)\/2} \\prod_{i \\neq i_0} \\D(\\ga_i) (32|z|)^\n{|\\bar\\ga_i|} \\leq\n|\\bar\\ga_{i_0}| ^{l_{i_0}} \\prod_{k \\neq i_0} 2^{l_{j_k}} (l_{j_k}-1)! \\; \n(k_3 \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|^2)^{r-1} \n\\end{equation*}\nwhere we used that $\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z| \\leq 1\/(\\const{q5} e)$. Thus we need \nfurthermore (see (\\ref{hae}))\n\\begin{eqnarray*}\n\\lefteqn{\n\\sum_{\\ga_{i_0}} |\\bar\\ga_{i_0}|^{l_{i_0}} \\D(\\ga_{i_0})\n(16|z|(1+\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|))^{|\\ga_{i_0}|} (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(0,\\ga_{i_0})\/2} } \\\\ &&\n\\hspace{0.3cm} \\leq \n\\sum_{\\tau} (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(0,\\tau)\/2} \\sum_{\\ga_{i_0}: \\tau \\in \\ga_{i_0}} \n|\\bar\\ga_{i_0}|^{l_{i_0}} \\D(\\ga_{i_0}) (32|z|)^{|\\ga_{i_0}|}.\n\\end{eqnarray*}\nBy a repetition of the arguments above we get \n\\begin{eqnarray*}\n\\sum_{\\ga_{i_0}: \\tau \\in \\ga_{i_0}} |\\bar\\ga_{i_0}|^{l_{i_0}}\n\\D(\\ga_{i_0}) (32|z|)^{|\\ga_{i_0}|}\n& \\leq &\n\\sum_{k=2}^\\infty k^{l_{i_0}} (C\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)^{k-1} (32|z|)^k \\\\\n& \\leq &\n\\const{q6} 2^{l_{i_0}} l_{i_0}! \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|^2,\n\\end{eqnarray*}\nwith $\\defconst{q6} > 0$. Summation over $\\tau$ gives\n$\n\\sum_\\tau (\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits|z|)^{\\mathop{\\mathrm{dist}}\\nolimits(0,\\tau)\/2} \\leq \\; \\const{q7}, \n$\nwith some $\\defconst{q7} > 0$, hence we finally obtain for fixed \n$\\tilde T$ and $i_0$\n\\begin{eqnarray*}\n2^{l_{i_0}} l_{i_0}! \\prod_{j_k \\neq i_0} 2^{l_k} (l_{j_k}-1)! \n\\const{q7} (\\const{q6} \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|^2)^{r-1}\n&\\leq&\n\\const{q7} (2^2 \\const{q6} \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|^2)^{r-1} \\prod_{k=1}^r l_{j_k}! \n\\end{eqnarray*}\nwhere we used the fact $\\sum_{k=1}^r l_{j_k} = 2(r-1)$ for trees. \nAn upper bound on the number of trees with vertices $\\{1,...,r\\}$ \nand degrees $\\{l_1,...,l_r\\}$ is \\cite{MM}\n\\begin{equation}\n\\frac{2^{r-2} (r-2)!}{\\prod_{j=1}^r l_j!}. \n\\label{inc}\n\\end{equation}\nMoreover, the number of collections $\\{l_1,...,l_r\\}$ such that $l_i > \n0$ and $\\sum_i l_i = 2(r-1)$ is bounded from above by $2^{2(r-1)}$.\nHence, by summing over $i_0$ and combining this estimate with\n(\\ref{inc}), we get\n\\begin{equation}\n\\sum_{\\Ga: \\Ga^* \\ni 0 \\atop \\Ga \\supset\\;\\mbox{\\tiny{more than one\n contour}}} K_\\Ga |z|^{|\\bar\\Ga|} \\leq c \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|^2\n\\end{equation}\nwith some constant $c > 0$. This completes the estimate of the second\nterm in (\\ref{H}). The first term there can be handled in a similar\nway with substantial simplifications due to the fact that only one \ncontour occurs in the clusters. \n\nIt is seen then that by choosing $z$ such that $\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits |z|^2 \\leq\n\\mbox{const}$, the sum $\\sum_\\Ga K_{\\Ga} z^{|\\bar \\Ga|}$ converges\nand is bounded. Hence $H(z)$ is an analytic function within a circle \nof radius $R(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)$ with $R(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits) \\to \\infty$ as $\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits \\to 0$, and is \nbounded by a constant independent of $\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits$. Thus \n\\begin{equation}\n\\sum_{\\Ga: \\Ga^* \\ni 0 \\atop |\\bar\\Ga| = n} |K_{\\Ga}| \\;\\leq\\; \n\\mbox{const} \\; R(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)^{-n} \\; := \\; \\mbox{const} \\; \\eta(\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits) ^n,\n\\end{equation}\nwith suitable constants.\n\\end{proof}\n\nFinally we show the lemma referred to in the proof above. \n\\begin{lemma}\n\\label{lsumd}\nThere is a constant $C > 0$ such that for any interval $\\tau$ \nand integer $k \\geq 2$ \n\\begin{equation}\n\\sum_{\\ga: \\bar\\ga \\ni \\tau \\atop |\\bar\\ga|=k} \\D(\\ga) \\; \\leq\\; \n(C\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)^{k-1}.\n\\label{sumd}\n\\end{equation}\n\\end{lemma}\n\\begin{proof}\n\\begin{equation}\n\\sum_{\\ga: \\tau \\in \\bar\\ga \\atop |\\bar\\ga| = k} \\D(\\ga) = \n\\sum_{\\{\\tau_1,...,\\tau_k\\}} \\sum_{G\\in \\hat\\G}\n\\prod_{\\{\\tau_i,\\tau_j\\} \\in G} \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits D(i,j)\n\\end{equation}\nHere $\\hat\\G$ denotes the set of connected graphs with vertices \n$\\tau_1,...,\\tau_k$. Note that for fixed $i_0$ we have \n$\\sum_{\\tau_j: j \\neq i_0} \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits D(i_0,j) \\le C \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits$ with some \n$C > 0$. Thus by using (\\ref{mm8}) we find\n\\begin{equation}\n\\sum_{G\\in\\hat\\G} \\prod_{\\{\\tau_i,\\tau_j\\} \\in G} \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits D(i,j)\n\\leq \ne^{k C \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits} \\sum_{T\\in \\hat\\T} \\prod_{\\{\\tau_i,\\tau_j\\} \\in T} \n\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits D(i,j) \n\\label{102}\n\\end{equation}\nwhere $\\hat\\T$ are trees for the same vertex set as for $\\hat\\G$. \nNext order the collection $\\{\\tau_1,...,\\tau_k\\}$ further to get \n\\begin{equation}\n\\sum_{\\{\\tau_1,...,\\tau_k\\}} \\sum_{T \\in \\hat T} \\prod_{\\{i,j\\} \n\\in \\tilde \\T} \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits D(i,j) =\n\\frac{1}{k!} \\sum_{(\\tau_1,...,\\tau_k)}\\sum_{\\tilde T \\in\\tilde\\T} \n\\prod_{\\{i,j\\} \\in \\tilde T} \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits D(i,j),\n\\label{101}\n\\end{equation}\nwith the same $\\tilde \\T$ as previously. We then obtain inductively\n\\begin{equation}\n\\sum_{(\\tau_1,...,\\tau_k)} \\prod_{(\\tau_i,\\tau_j) \\in \\T} \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits D(i,j)\n\\leq (C \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)^{k-1}.\n\\end{equation}\nSince the number of trees having $k$ vertices is $k^{k-2}$ \\cite{MM}, \nby using Stirling's formula, (\\ref{102}) and (\\ref{101}) we complete \nthe proof of the lemma.\n\\end{proof}\n\n\n\\subsection{Cluster estimates}\nHaving the abstract cluster expansion at hand, we turn now to\nestablishing the bounds~(\\ref{clustest-abs}) on the cluster\nactivities. \n\\begin{proposition}\nThere exists $\\delta > 1$ and a function\n$\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits(\\lambda)<\\infty$ with $\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits(\\lambda) \\to 0$ as $\\lambda \\to 0$ and\n$b = b(\\lambda)\\ge 1$ such that for every $N > 0$ and every cluster\n$\\Ga = \\{\\ga_1,...,\\ga_r;\\rh_1,...,\\rh_s\\} \\in \\mathop{\\mathrm{{\\cal K}}}\\nolimits_N$, the bound \n\\begin{equation}\n|K_{\\Gamma}| \\; \\leq \\; \\prod_{l=1}^r \\prod_{(\\tau_i,\\tau_j) \\in \n\\gamma_l} \\frac{\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits}{(1+b|i-j-1|)^\\delta }\\prod_{m=1}^s \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits^{|\\bar\\rh_m|}\n\\label{clustest}\n\\end{equation}\nholds.\n\\label{pclustest}\n\\end{proposition}\nThe first product (over the contours) above is our $D(i,j)$ in\nProposition \\ref{p2} above, and it is readily seen that it satisfies\nthe condition given there.\n\n\\medskip\n\\begin{proof}\nBy H\\\"older inequality\n\\begin{eqnarray}\n|K_{\\Ga}| \n& \\leq & \n\\prod_{l=1}^r \\prod_{(\\tau_i,\\tau_j) \\in \\gamma_l} \\left(\\int \\mathbb{E}_{\\nu_T^{\\mathbf{x}}}\n\\left[|e^{-\\lambda W_{\\tau_i,\\tau_j}} - 1|^{n_{ij}}\\right]\nd\\omega^{\\otimes N}(\\mathbf{x}) \\right)^{1\/n_{ij}} \n\\times \\\\\n&& \\hspace{0.3cm} \\times \n\\prod_{m=1}^s \\prod_{k: \\tau_k \\in \\rh_m}\\left(\\int|\\pi_b(x_{k+1},x_k) - 1|^\\beta \nd\\omega(x_k)d\\omega(x_{k+1})\\right)^{1\/\\beta} \\nonumber\n\\end{eqnarray}\nwith suitable exponents. We choose $\\beta = 4$, $n_{ij} = A |i-j+1|^{\\Delta}$, \nwith $\\Delta > 1$ to be specified below. Taken with correct \nmultiplicities, we pick $A$ such that\n$$\n\\frac{2}{\\beta} + \\sum_{j \\in \\mathbb{N} \\atop j \\geq i}\\frac{2}{n_{ij}} =\n\\frac{1}{2} + \\frac{2}{A}\\sum_{k=1}^\\infty \\frac{1}{k^\\Delta} \\leq 1.\n$$\nThe first part of estimate (\\ref{clustest}) follows by Lemma \\ref{holderw}, \nthe second by Lemma \\ref{holderpfi} below. By choosing $b= - \\log |\\lambda|\/\n(\\Lambda+C)$ with suitable $C>0$, we have\n$\n\\lambda e^{Cb} = e^{-\\Lambda b} = |\\lambda|^{\\Lambda\/(\\Lambda+C)} =: \\mathop{\\mathrm{{\\varepsilon}}}\\nolimits,\n$\nthus the estimate (\\ref{clustest}) is finally obtained.\n\\end{proof}\n\n\\begin{lemma}\n\\label{lemma:trans}\nFor large enough $b > 0$ there is a constant $C > 0$ such that\n\\begin{equation}\n|\\pi_b(x,y) - 1| \\,\\leq\\, Ce^{-\\Lambda b},\n\\end{equation}\nuniformly in $x,y \\in \\mathbb{R}^d$, where $\\Lambda > 0$ is the spectral gap of the \nSchr\\\"odinger operator $H = H_0 + V$. \n\\label{holderpfi}\n\\end{lemma}\n\\begin{proof}\nBy assumption the potential is chosen so that $H$ is intrinsically \nultracontractive, i.e., for each $b > 0$, $C_b = \\norm[L^{\\infty}(\\mathbb{R}^{2d})]{\\pi_b} \n< \\infty$. \n By the semigroup property of $\\pi_b$ and the fact that $\\int \\pi_b(x,y) d\\omega(y) = \n 1$ for each $x$, for $b > 2$ we have\n \\begin{equation}\n \\label{iucestimate}\n \\begin{split}\n |\\pi_b(x,y) - 1| \n & = \n \\Big| \\int d\\xi \\int d\\eta \\, \\pi_1(x,\\xi)\\Psi^2(\\xi) (\\pi_{b-2}(\\xi,\\eta) - 1) \n \\Psi^2(\\eta) \\pi_1(\\eta,y) \\Big| \\\\\n & \\leq \n C_1^2 \\int d\\xi \\int d\\eta \\, \\Psi(\\xi) | \\hat\\pi_{b-2}(\\xi,\\eta) - \n \\Psi(\\xi)\\Psi(\\eta)| \\Psi(\\eta) \\\\\n & \\leq \n C_1^2 \\left( \\int d\\xi \\int d\\eta (\\hat\\pi_{b-2}(\\xi,\\eta) - \\Psi(\\xi) \\Psi(\\eta))^2 \n \\right)^{1\/2} \\\\\n & = \n C_{1}^2 e^{-(b-2)(E_1-E)} \\left(\\sum_{k\\geq 2} e^{-2(b-2)(E_k- E_1)} \\right)^{1\/2} \n\\leq \n Ce^{-\\Lambda b},\n \\end{split}\n \\end{equation}\nwhere $E = \\inf \\mathop{\\mathrm{Spec}}\\nolimits H$.\nThe last but first step comes about as follows. $\\norm[L^{\\infty}(\\mathbb{R}^{2d})]{\\pi_b} < \\infty$ \nimplies that $\\hat\\pi_b \\in L^2(\\mathbb{R}^{2d},dx)$ for each $b > 0$. Thus $e^{-bH}$ is a \nHilbert-Schmidt operator for each $b>0$, in particular $H$ has a purely discrete spectrum \nwith eigenvalues $E < E_{1} \\leq E_{2} \\leq \\ldots$. With $P_\\Psi$, the projection onto \nthe subspace of $L^2(\\mathbb{R}^d,dx)$ spanned by $\\Psi$, the last equality gives the Hilbert-Schmidt \nnorm of $e^{-(b-2)H} - P_\\Psi$.\n\\end{proof}\n\n\\medskip\n\\begin{lemma}\nAssume that $\\beta > 3$. Then there exists $\\delta > 1$ and\nconstants $\\defconst{ggg}, \\defconst{ggg2} < \\infty$ such that, \nfor $\\lambda$ small enough\n\\begin{equation}\n\n\\begin{split}\n\\left(\\int \\mathbb{E}_{\\nu_T^{\\mathbf{x}}}\\left[ |e^{\\lambda\nW_{\\tau_i,\\tau_j}}-1|^{n_{ij}}\\right] d\\omega^{\\otimes N}\n(\\mathbf{x}) \\right)^{1\/n_{ij}} \n& \\le\n\\const{ggg} \\lambda e^{C b} (1+b|i-j-1|)^{\\delta}\n\\end{split}\n\\end{equation}\nfor all $i,j$ and $b \\ge 1$.\n\\label{holderw}\n\\end{lemma}\n\\begin{proof}\nNote that the conditional expectation $\\mathbb{E}_{\\nu_T^{\\mathbf{x}}}\n[|e^{\\lambda W_{\\tau_i,\\tau_j}}-1|^{n_{ij}}]$ depends only on \n$x_i,x_{i+1},x_{j},x_{j+1}$. Thus we can write with a slight \nabuse of notation,\n$$\n\\int \\mathbb{E}_{\\nu_T^{\\mathbf{x}}} \n\\left[|e^{\\lambda W_{\\tau_i,\\tau_j}}-1|^{n_{ij}}\\right] \nd\\omega^{\\otimes N}(\\mathbf{x}) =\n\\int \\mathbb{E}_{\\nu_T^{\\mathbf{x}}} \n\\left[|e^{\\lambda W_{\\tau_i,\\tau_j}}-1|^{n_{ij}}\\right] \nd\\omega^{\\otimes 4}(x_i, x_{i+1}, x_j, x_{j+1})\n$$\nwhen $|i-j|>1$, otherwise one integral must be ignored in this \nexpression. Then by (\\ref{eq:bridge-2}) the problem reduces to \nestimates on the multiple Brownian bridge $\\widehat\\mathcal{W}_T^{\\mathbf{x}}$:\n\\begin{equation*}\n \\begin{split}\n& A = \\int \\mathbb{E}_{\\nu_T^{\\mathbf{x}}} \n \\left[|e^{\\lambda W_{\\tau_i,\\tau_j}}-1|^{n_{ij}}\\right] \n d\\omega^{\\otimes 4}(x_i,x_{i+1},x_j,x_{j+1}) \\\\ \n& \\le\n\\int \\frac{\\mathbb{E}_{\\widehat\\mathcal{W}_T^{\\mathbf{x}}} \n \\left[|e^{\\lambda W_{\\tau_i,\\tau_j}}-1|^{n_{ij}} \n e^{-V_{\\tau_i}-V_{\\tau_j}}\\right]}\n {Z_b \\Psi(x_i) \\Psi(x_{i+1}) \\pi_{b}(x_i,x_{i+1})}\n \\frac{\\Pi_{b}(x_i,x_{i+1}) \\Pi_{b}(x_j,x_{j+1})}\n {Z_b \\Psi(x_j) \\Psi(x_{j+1}) \\pi_{b}(x_j,x_{j+1})} \n d\\omega^{\\otimes 4}(x_i,x_{i+1},x_j,x_{j+1}) \\\\ \n& \\le \nC Z_b^{-2} \\int \\mathbb{E}_{\\widehat\\mathcal{W}_T^{\\mathbf{x}}} \n \\left[|e^{\\lambda W_{\\tau_i,\\tau_j}}-1|^{n_{ij}} \n e^{-V_{\\tau_i}-V_{\\tau_j}}\\right]\n \\Pi_{b}(x_i,x_{i+1}) \\Pi_{b}(x_j,x_{j+1}) \\times \\\\ \n& \\qquad \\times \n\\Psi(x_i) \\Psi(x_{i+1}) \\Psi(x_j) \\Psi(x_{j+1})\ndx_i dx_{i+1} dx_{j} dx_{j+1},\n \\end{split}\n\\end{equation*}\nwhere we used Lemma \\ref{holderpfi} and chose $b$ large enough \nso that\n$\n\\sup_{x,y} |\\pi_{b}(x,y)-1|\\le C e^{-\\Lambda b} \\le 1\/2\n$.\nFor $i+1 0$. Furthermore, write\n$\n W_{\\tau_i,\\tau_j} = \\int_{\\tau_i} dX_t \\int_{\\tau_j} dX_s g_{ts}\n$\n(see (\\ref{eq:J-rewriting}) below) and estimate the double integral below\nby using the Burkholder-Davis-Gundy inequality,\n\\begin{equation*}\n \\begin{split}\n\\widehat\\mathbb{E}_{x_i,x_j} \\left[|\n W_{\\tau_i,\\tau_j}|^{2n}\\right] & \\le c_{n} \\widehat\\mathbb{E}_{x_i,x_j} \\left[\\left|\n \\int_{\\tau_i} dt\\left | \\int_{\\tau_j} dX_s g_{ts}\\right|^2\n \\right|^{n}\\right] \n\\\\ & \\le c_{n} |\\tau_i|^{n-1} \\int_{\\tau_i} dt \\widehat\\mathbb{E}_{x_i,x_j} \\left[\n \\left | \\int_{\\tau_j} dX_s g_{ts}\\right|^{2n} \\right] \n\\\\ & \\le c_{n}^2 |\\tau_i|^{n-1} |\\tau_j|^{n-1} \\int_{\\tau_i} dt \\int_{\\tau_j} ds \\widehat\\mathbb{E}_{x_i,x_j} \\left[\n |g_{ts}|^{2n} \\right] \n\\\\ & \\le c_{n}^2 b^{2n} C^n (1+b|i-j-1|)^{-2 n \\beta}.\n \\end{split}\n\\end{equation*}\nNow we estimate also the exponential of the energy. By $e^{|x|} \\le 2\n\\cosh x$ we get\n \\begin{equation*}\n \\begin{split}\n \\widehat\\mathbb{E}_{x_i,x_j} \\left[ e^{4 n |\\lambda\n W_{\\tau_i,\\tau_j}|} \\right] \\le \\widehat\\mathbb{E}_{x_i,x_j} \\left[ e^{4 n \\lambda\n W_{\\tau_i,\\tau_j}} \\right] + \\widehat\\mathbb{E}_{x_i,x_j}\n \\left[ e^{- 4 n \\lambda\n W_{\\tau_i,\\tau_j}} \\right]. \n \\end{split}\n\\end{equation*}\nEach of the expectations in the right hand side can be similarly \nestimated by using Lemma~\\ref{lemma:final-exp-bound} below:\n\\begin{equation*}\n \\begin{split}\n\\widehat\\mathbb{E}_{x_i,x_j} \\left[ e^{4 n |\\lambda\n W_{\\tau_i,\\tau_j}|} \\right] \\le (1-n^2 a^2)^{-\\const{b}} \\le C\n \\end{split}\n\\end{equation*}\nwith \n$\na = \\const{a}\n \\lambda b (1+ b |i-j-1|)^{-\\beta} \n$ and $na \\le 1\/2$.\nHence, by making use these inequalities (uniform in $x_i,x_j$), \nand the fact that $\\Psi \\in L_1(\\mathbb{R}^d,dx)$, we arrive at\n\\begin{equation}\n \\label{eq:ewbound}\n\\begin{split}\n& \n\\left(\\int \\widehat\\mathbb{E}_{x_i,x_j} \n\\left[ |e^{\\lambda W_{\\tau_i,\\tau_j}}-1|^{n_{ij}}\n e^{-V_{\\tau_i}-V_{\\tau_j}}\\right] \n\\Psi(x_i) dx_i \\Psi(x_j) dx_j \\right)^{1\/n_{ij}} \\\\\n&\n\\hspace{1cm} \\le \nC c_{n_{ij}}^{1\/n_{ij}}\\lambda be^{C b}(1+b|i-j-1|)^{-\\beta}. \n\\end{split}\n\\end{equation}\nBy using the estimate $c_k \\le (2k)^{2k}$ for the universal \nconstant in the Burkholder-Davis-Gundy inequality, we furthermore\nobtain\n$$\n c_{n_{ij}}^{1\/n_{ij}} (1+b|i-j-1|)^{-\\beta} \\le (2n_{ij})^2\n (1+b|i-j-1|)^{-\\beta} .\n$$\nRecall that $n_{ij} = A |i-j|^\\Delta$ with $\\Delta >1$ and choose \n$\\Delta <\\beta$ so that $an_{ij} \\to 0 $ as $|i-j|\\to\\infty$ and \ncondition $an_{ij}\\le 1\/2$ is satisfied uniformly in $i,j$ for \n$\\lambda$ small enough. Thus there is a constant $\\defconst{fff} > \n0$ such that\n\\begin{equation}\n\\begin{split}\n\\text{r.h.s. (\\ref{eq:ewbound})}\n\\, \\le \\,\n\\const{fff} \\lambda b^{2-2\\Delta} e^{C b}(1+b|i-j-1|)^{2\\Delta-\\beta}\n\\, \\le \\, \n\\const{fff} \\lambda e^{C b}(1+b|i-j-1|)^{-\\delta}, \n\\end{split}\n\\end{equation}\nfor all $|i-j| > 1$ and $b \\ge 1$, where $\\delta = \\beta-2\\Delta>1$\n(this is possible by choosing $\\Delta>1$ but small enough). For the \ncases $|i-j|=1$ we can follow a similar strategy to estimate\n$$\n\\left(\\mathbb{E}_{\\nu_T^{\\mathbf{x}}} \\left[|e^{\\lambda W_{\\tau_i,\\tau_j}}-1|^{2n_{ij}} \ne^{-V_{\\tau_i}-V_{\\tau_j}}\\right]\\right)^{1\/2n_{ij}} \n\\le \n\\defconst{tt2} e^{\\defconst{tt1} b}, \n$$ \nwhere the constants do not depend on either $b$ or $n_{ij}$.\n\\end{proof}\n\n\\subsection{Energy estimates}\nHere we estimate $\\widehat \\mathbb{E}_{x,y} [e^{\\lambda W_{\\tau_i,\\tau_j}}\n]$.\nBelow we will prove two lemmas that give the basic estimates by\nmaking use of the following result to control exponential\nintegrability of stochastic integrals. The first lemma will be \noften used for controlling exponential moments of stochastic integrals.\n\\begin{lemma}\n\\label{lemma:exp-bound}\nIf $X$ is Brownian motion and $f$ is an $\\mathop{\\mathrm{{\\cal F}}}\\nolimits$-adapted process, we \nhave the bound\n\\begin{equation}\n \\label{eq:exp-bound}\n\\mathbb{E} [e^{\\int_0^b f_s dX_s}] \\le \n\\left\\{\\mathbb{E} [e^{2 \\int_0^b |f_s|^2 ds} ] \\right\\}^{1\/2}.\n\\end{equation}\n\\end{lemma}\n\\begin{proof}\nThe proof is a combination of Cauchy-Schwartz and Girsanov Theorems:\n \\begin{equation*}\n \\begin{split}\n \\mathbb{E} [e^{\\int_0^b f_s dX_s}] &= \n \\mathbb{E} [e^{\\int_0^b f_s dX_s - \\int_0^b |f_s|^2 ds +\n \\int_0^b |f_s|^2 ds} ]\n\\\\ & \\le \\left\\{\\mathbb{E} [e^{2 \\int_0^b f_s dX_s - \\frac{1}{2} \\int_0^b\n |2 f_s|^2 ds}]\\right\\}^{1\/2}\\left\\{\\mathbb{E} [e^{\n 2 \\int_0^b |f_s|^2 ds} ] \\right\\}^{1\/2}\n\\\\ & \\le \\left\\{\\mathbb{E} [e^{\n 2 \\int_0^b |f_s|^2 ds} ] \\right\\}^{1\/2}.\n \\end{split}\n \\end{equation*}\n\\end{proof}\n\n\\medskip\n\\noindent\n\\emph{\\textbf Estimates for separated intervals} \\quad\nWe turn to estimating exponentials of energy contributions in \n(\\ref{sum}) by starting with pairs of intervals that are not \nadjacent. Let thus $i>j+1$; in this case the exponent has the \nform $W_{\\tau_i,\\tau_j} = \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ij} + \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ji}$ with\n\\begin{equation}\n \\label{eq:J-rewriting}\n\\mathop{\\mathrm{{\\cal J}}}\\nolimits_{i,j} = \\langle C^X_{\\tau_i},C^X_{\\tau_j}\\rangle_W = \n\\int_{t_i}^{t_i+b} dX_t \\int_{t_j}^{t_j+b} d X_s\\, W(X_t-X_s,t-s). \n\\end{equation}\nNote that under the\nmeasure $\\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x_i}_{\\tau_i} \\otimes \\mathop{\\mathrm{{\\cal W}}}\\nolimits^{x_j}_{\\tau_j}$\nthe two currents $\\mathbb{X}_{\\tau_i}$ and $\\mathbb{X}_{\\tau_j}$ are independent \nand the interaction energy can be written as a double stochastic\nIt\\^o integral and estimated by using tools borrowed from stochastic \nanalysis.\n\\begin{lemma}\n\\label{lemma:final-exp-bound}\nLet $i>j+1$. There exist positive constants $\\defconst{a},\\defconst{b}$ \nsuch that whenever\n\\begin{equation}\n \\label{eq:cond-bound-1}\na := \\lambda \\const{a} b (1+ |t_i-t_j-b|)^{-\\beta} < 1, \n\\end{equation}\nwe have\n\\begin{equation*}\n \\begin{split}\n\\widehat \\mathbb{E}_{x,y} [e^{\\lambda W_{\\tau_i,\\tau_j}}] &\n\\le \\left[1-a^2\\right]^{-\\const{b}}.\n \\end{split}\n\\end{equation*} \n\\end{lemma}\n\\begin{proof}\nWe have\n\\begin{equation*}\n \\begin{split}\n\\left(\\widehat \\mathbb{E}_{x,y} [e^{\\lambda W_{\\tau_i,\\tau_j}}]\\right)^4 &\n\\le\n\\left(\\widehat \\mathbb{E}_{x,y} [e^{2\\lambda \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ij}}]\\right)^2 \n\\left(\\widehat \\mathbb{E}_{x,y} [e^{2\\lambda \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ji}}]\\right)^2. \n \\end{split}\n\\end{equation*}\nBy using (\\ref{eq:exp-bound}) we obtain\n\\begin{equation*}\n \\begin{split}\n\\left(\\widehat \\mathbb{E}_{x,y} [e^{\\lambda \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ij}}]\\right)^2 &\\le\n\\widehat \\mathbb{E}_{x,y} \\left[\\exp\\left(2 \\lambda^2 \\int_{t_i}^{t_i+b} dt\n \\left|\\int_{t_j}^{t_j+b} d X_s W(X_t-X_s,t-s)\n \\right|^2\\right)\\right]\n\\\\ &\\le \\int_{t_i}^{t_i+b} \\frac{dt}{b}\n\\widehat \\mathbb{E}_{x,y} \\left [\\exp\\left(2 b \\lambda^2 \n \\left|\\int_{t_j}^{t_j+b} d X_s W(X_t-X_s,t-s) \\right|^2\\right)\\right ]\n\\\\ &\\le \\int_{t_i}^{t_i+b} \\frac{dt}{b}\n\\widehat \\mathbb{E}_{x,y} \\mathbb{E}_G \\left[\\exp\\left(2 \\lambda\n G \\sqrt{b}\\int_{t_j}^{t_j+b} d X_s W(X_t-X_s,t-s)\\right)\\right],\n \\end{split}\n\\end{equation*}\nwith $G$ a normal random variable defined on a new probability space\nand $\\mathbb{E}_G$ the related Gaussian expectation. Then, using again \n(\\ref{eq:exp-bound}) yields\n\\begin{equation*}\n \\begin{split}\n\\left(\\widehat \\mathbb{E}_{x,y} [e^{\\lambda W_{\\tau_i,\\tau_j}}]\\right)^2 &\n\\le \\int_{t_i}^{t_i+b} \\frac{dt}{b} \n\\left(\\widehat \\mathbb{E}_{x,y} \\mathbb{E}_G \\left[\\exp\\left(8 \\lambda^2 G^2 \nb \\int_{t_j}^{t_j+b} ds |W(X_t-X_s,t-s)|^2 \\right)\\right]\\right)^{1\/2}.\n \\end{split}\n\\end{equation*}\nThe assumptions on $W$ give\n\\begin{equation*}\n \\begin{split}\n \\int_{t_j}^{t_j+b} ds |W(Y_t-Z_s,t-s)|^2 & \\le \n \\int_{t_j}^{t_j+b} ds \\frac{C}{(1+|t-s|)^{2\\beta}}\n \\le {\\const{c}} b (1+|t_i-t_j-b|)^{-2\\beta}\n \\end{split}\n\\end{equation*}\nfor some constant $\\defconst{c}$,\nwhere we used that $t \\in [t_i,t_i+b]$. Hence\n\\begin{equation*}\n \\begin{split}\n\\left(\\widehat \\mathbb{E}_{x,y} [e^{2\\lambda \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ij}}]\\right)^2 &\n\\le \\left(\n\\mathbb{E}_G \\left[\\exp\\left[32 \\lambda^2 \\const{c} b^2 \n(1+|t_i-t_j-b|)^{-2\\beta} G^2 \\right]\\right]\\right)^{1\/2}.\n \\end{split}\n\\end{equation*}\nThe Gaussian integration can be performed explicitly, yielding\n\\begin{equation*}\n \\begin{split}\n\\widehat \\mathbb{E}_{x,y} [e^{2\\lambda \\mathop{\\mathrm{{\\cal J}}}\\nolimits_{ij}}] &\n\\le \n \\left(1-32 \\lambda^2 \\const{c} b^2 (1+|t_i-t_j-b|)^{-2\\beta}\n \\right)^{-d\/8},\n \\end{split}\n\\end{equation*}\n as soon as\n$\n1-32 \\lambda^2 \\const{c} b^2(1+ |t_i-t_j-b|)^{-2\\beta} > 0.\n$\nThus the claim follows.\n\\end{proof}\n\n\\medskip\n\\noindent\n\\emph{\\textbf Estimates for adjacent intervals} \\quad\nThe interaction energy estimates between adjacent intervals are \ngiven by\n\n\\begin{lemma}\nFor all $i = 0,...,N-2$ we have\n$\n\\widehat \\mathbb{E}_{x,y} [e^{\\lambda W_{\\tau_i,\\tau_{i+1}}}] \\le\n\\defconst{diag} e^{\\defconst{diag2} b } < \\infty\n$\nfor sufficiently small $\\lambda$.\n\\end{lemma}\n\\begin{proof}\nBy using that $W(x,t)$ is bounded and arguments similar to those \nof the previous lemma, the required exponential integrability of\n$W_{\\tau_i,\\tau_{i+1}}$ easily follows, at least for sufficiently\nsmall $\\lambda$.\n\\end{proof}\n\n\\subsection{Properties of the cluster expansion}\nWe finally show how the convergence of the cluster expansion of \n$Z_T$ and $\\log Z_T$ seen in Proposition \\ref{p0} imply existence \nof a limit Gibbs measure $\\mu$. \n\nFor any subset $A \\subset \\mathbb{R}$ let\n$$\nZ^A_T = 1 + \\sum_{n\\geq 1} \n\\sum_{{\\Ga_1,...,\\Ga_r \\in \\C_N \\atop \\Ga^*_i \\cap \\Ga^*_j \n= \\emptyset, \\; i \\neq j} \\atop A \\cap (\\cup_i\\Ga^*_i) \n= \\emptyset} \\prod_{i=1}^r K_{\\Ga_i}\n$$\nand write $Z_T^\\Gamma := Z_T^{\\overline \\Gamma \\cup \\Gamma^*}$.\nBy the cluster expansion we have\n$$\n\\log Z^\\Gamma_T = 1 + \\sum_{n\\geq 1} \n\\sum_{{\\Ga_1,...,\\Ga_r \\in \\C_N } \\atop \\Ga^* \\cap (\\cup_i\\Ga^*_i) \n= \\emptyset} \\phi^T(\\Ga_1,...,\\Ga_n)\\prod_{i=1}^r K_{\\Ga_i}.\n$$\nMoreover we can define the correlation functions for the clusters by\n\\begin{equation}\n \\label{eq:cluster-corr}\nf^\\Gamma_T = \\frac{Z_T^\\Gamma}{Z_T} = \\exp\\left( - \\sum_{n\\geq 1} \n\\sum_{{\\Ga_1,...,\\Ga_r \\in \\C_N } \\atop \\Ga^* \\cap (\\cup_i\\Ga^*_i) \n\\neq \\emptyset} \\phi^T(\\Ga_1,...,\\Ga_n)\\prod_{i=1}^r K_{\\Ga_i} \\right) \n\\end{equation}\nand let $f^\\Gamma = \\lim_{T\\to\\infty} f^\\Gamma_T $ as the limit\nexists by the cluster estimates above and general arguments of \ncluster expansion~\\cite{MM}. Moreover we have the uniform estimate\n\\begin{equation}\n \\label{eq:cluster-corr-estimate}\n|f^\\Gamma_T| \\le 2^{|\\overline \\Gamma|} \n\\end{equation}\nfor $\\lambda$ small enough (the constant $2$ can actually be\nreplaced with any number larger than $1$, provided $\\lambda$ is\nchosen correspondingly small). Then existence of the infinite time \nlimit for the measures $\\{\\mu_T\\}_T$ follows easily and we have\n\\begin{proposition}\n \\label{prop:infinite-measure}\nThe local limit $\\mu = \\lim_{T\\to\\infty} \\mu_T$ exists and satisfies \nthe equality\n\\begin{equation}\n \\label{eq:limit-eq-cluster}\n\\mathbb{E}_{\\mu} [F_S] =\n\\mathbb{E}_{\\chi} [F_S] f^{S} + \\sum_{n\\geq 1} \n\\sum_{{\\Ga_1,...,\\Ga_r \\in \\C \\atop \\Ga^*_i \\cap \\Ga^*_j \n= \\emptyset, \\; i \\neq j} \\atop i: S \\cap \\Ga^*_i \n\\neq \\emptyset}\n \\mathbb{E}_{\\chi}[F_S \\prod_{i=1}^r \\kappa^{\\Ga_i}] \nf^{\\cup \\overline \\Ga}\n\\end{equation}\nfor any bounded, $\\mathcal{F}_S$-measurable function $F_S$, where \n$S$ is a finite union of intervals of the partition considered in \nthe cluster expansion. Moreover, the measure $\\mu$ is invariant \nwith respect to time shift.\n\\end{proposition}\n\\begin{proof}\nWe have\n\\begin{equation}\n\\mathbb{E}_{\\mu}[F_S] = \\lim_{T \\rightarrow \\infty} \n\\frac{\\int e^{-\\lambda W_T(X)} F(X) d\\nu_T(X)}\n{\\int e^{-\\lambda W_T(X)} d\\nu_T(X)} = \n\\lim_{T \\rightarrow \\infty} \\frac{Z_T(F)}{Z_T}.\n\\end{equation}\nBy the cluster expansion we are led to \n\\begin{eqnarray}\nZ_T(F) = \n\\mathbb{E}_{\\chi} [F_S] Z_T^{S} + \\sum_{n\\geq 1} \n\\sum_{{\\Ga_1,...,\\Ga_r \\in \\C_N \\atop \\Ga^*_i \\cap \\Ga^*_j \n= \\emptyset, \\; i \\neq j} \\atop i: S \\cap \\Ga^*_i \n\\neq \\emptyset}\n \\mathbb{E}_{\\chi}[F_S \\prod_{i=1}^r \\kappa^{\\Ga_i}] \nZ_T^{\\cup \\overline \\Ga}.\n\\label{ce}\n\\end{eqnarray}\nIf $F_S$ is bounded, standard arguments show that the series \non the right hand side is absolutely convergent uniformly in \n$N$ and thus~(\\ref{eq:limit-eq-cluster}) follows. Given the \nuniqueness of the limiting measure, its invariance with respect \nto time shifts is a direct consequence of the invariance of the \npotentials and of the It\\^o-measure~(for more details \nsee~\\cite{LM}).\n\\end{proof}\n\n\\begin{corollary}\n\\label{cor:equiv-mu-nu}\nLet $F\\in\\mathcal{F}_{[0,b]}$ be a positive random variable. \nThen\n$\n\\mathbb{E}_{\\mu}[F] \\leq C (\\mathbb{E}_{\\nu}[F^2])^{1\/2} \n$. \n\\end{corollary}\n\\begin{proof}\nBy using Proposition ~\\ref{prop:infinite-measure} for $S=[0,b]$ \n(which for fixed $N$ is the interval $\\tau_{N\/2}$ of the \npartition) we have\n$$\n\\mathbb{E}_{\\mu} [F_S] = \\mathbb{E}_{\\chi} [F_S] \nf^S + \\sum_{\\Ga_0: \\bar\\Ga_0 \\cap [0,b] \\neq \\emptyset} \n\\mathbb{E}_{\\chi} [F_S \\kappa^{\\Ga_0}] f^{\\Ga_0}, \n$$\nwhere in the second term the sum is over the only cluster which\ncan overlap with $S$. We have $|f^S| \\le 2$, $|f^{\\Ga_0}| \\le \n2^{|\\Ga_0^*|}$. On the other hand, by using Lemma~\\ref{lemma:trans} \nand choosing $b$ large enough to ensure that \n$\\sup_{x,y}|\\pi_b (x,y)-1|\\le 1\/2$, we have\n\\begin{equation}\n\\mathbb{E}_{\\chi} [F_S] = \\int F_S(X) \\frac{d\\nu^{x_0,x_1}(X)}\n{\\pi_b(x_0,x_1)} \\pi_b (x_0,x_1) d\\omega(x_0) d\\omega(x_1) \n\\; \\leq \\; \nC \\; \\mathbb{E}_{\\nu} [F_S]. \n\\nonumber\n\\end{equation}\nFurthermore,\n$\n\\mathbb{E}_{\\chi^N} [\\kappa_a^{\\Ga_0}] \\leq \n\\left(\\mathbb{E}_{\\chi^N} [\\xi_a]\\right)^{1\/2} \\; \n\\left(\\mathbb{E}_{\\chi^N} [(\\kappa^{\\Ga_0})^2]\\right)^{1\/2}\n$\nand by the same arguments as in Proposition~\\ref{p2} above we \nobtain the bound\n$$\n\\sum_{\\Ga_0: [0,b]\\in \\bar \\Ga_0 } [ \\mathbb{E}_{\\chi} \n(\\kappa^{\\Ga_0})^2]^{1\/2} \\; 2^{|\\bar\\Ga_0|} \\; \\leq \\; \n\\mbox{const}\n$$\nwith some constant. Hence we get that\n$\n\\mathbb{E}_{\\mu}[F_S] \\le C\n\\mathbb{E}_{\\nu}[F_S] + C\n(\\mathbb{E}_{\\nu}[F_S^2])^{1\/2} \n$,\nwhich implies the claim.\n\\end{proof}\n\n\\begin{theorem}\n\\label{th:mu-lift}\nThere exists a unique forward current $\\mu^\\sharp$ on $\\Xi$ \nsuch that its $\\Xi$-marginal is $\\mu$. Moreover, under $\\mu$ \nwe have $\\mathcal{N}_\\alpha(X) < \\infty$ almost surely and \nthe boundary currents are well defined under $\\mu^\\sharp$.\n\\end{theorem}\n\\begin{proof}\nCorollary~\\ref{cor:equiv-mu-nu} implies that the measure $\\mu$ \nis absolutely continuous with respect to $\\nu$, thus the almost \nsure events of $\\nu$ carry over to $\\mu$ and we can consider \nthe lifted measure $\\mu^\\sharp$. This further implies that\n$$\n\\mathbb{E}_\\mu[N_{[k,k+1]}(X)^3] \n\\le \n\\left(\\mathbb{E}_\\nu[(N_{[k,k+1]}(X))^6]\\right)^{1\/2} \n\\le \nC, \n$$\nindependently of $k\\in\\Z$ and then $\\mathbb{E}_\\mu[\\mathcal{N}_\n\\alpha(X)] < \\infty$ for any $\\alpha > 1$. This last \ncondition guarantees the existence of the boundary currents \nunder the measure $\\mu^\\sharp$.\n\\end{proof}\n\n\\section{Properties of the Gibbs measure}\n\n\\subsection{Dependence on boundary conditions and DLR uniqueness}\n\nUniqueness in DLR sense means that for any increasing sequence of real \nnumbers $\\{T_n\\}_n$, $T_n \\uparrow \\infty$, and any corresponding \nsequence of boundary conditions $\\{Y_n\\}_n \\subset \\Xi$ we have \n$\\mathbb{E}_{\\rho^\\sharp_{T_n}(\\cdot |\\mathbb{Y}_n)}[F_B] \\to \\mathbb{E}_\\mu[F_B]$, \nfor every bounded $B \\subset \\mathbb{R}$, and each bounded and local (i.e., \nmeasurable with respect to $\\mathop{\\mathrm{{\\cal F}}}\\nolimits_B$) function $F_B$ on $\\Xi$. However,\nsuch a strong statement cannot be made in this context and we have to\nrestrict the class of allowed boundary conditions to be able to\ncontrol the limit. Fix $\\alpha > 1$ and let\n\\begin{equation}\n\\Xi_a = \\{Y \\in \\Xi : \\mathcal{N}_\\alpha(Y) \\le a \\},\n\\quad \\Xi_* = \\cup_{a > 0} \\Xi_a\n\\end{equation} \nbe the set of \\emph{allowed boundary conditions} carrying full \n$\\mu^\\sharp$ measure. Then we have\n\\begin{theorem}\n\\label{th:uniq}\nFor any $a>0$ the measure $\\mu$ is unique in DLR sense for any \nsequence of boundary conditions $(\\mathbb{Y}_n)_n$ in $\\Xi_a$, i.e.\n\\begin{equation}\n\\lim_{n\\to\\infty} \\mathbb{E}_{\\rho^\\sharp_{T_n}(\\cdot|\\mathbb{Y}_n)}[F_B] =\n\\mathbb{E}_\\mu[F_B].\n\\label{dlruniq}\n\\end{equation}\n\\end{theorem}\n\\begin{proof}\nWe consider the class of bounded local functions $F_S$ on $\\Xi$ indexed \nby bounded intervals $S \\subset \\mathbb{R}$ (that is, $F_S$ is measurable with \nrespect to $\\mathop{\\mathrm{{\\cal F}}}\\nolimits_S$). It suffices to prove that for any increasing sequence \n$\\{T_n\\}$, $T_n \\to \\infty$, and any corresponding sequence of boundary \nconditions $(\\mathbb{Y}_n)_n \\subset \\Xi_a$ (\\ref{dlruniq}) holds for arbitrary \n$F_S$ of the above class. To show this we express the conditional \nexpectations appearing above in terms of the cluster representation. We \nsuppose without loss that $S$ consists of a finite union of intervals of \nthe partition of $[-T,T]$. \n\nFrom now on we follow the steps of the construction of the cluster\nrepresentation in Section \\ref{sec:cluster-rep}. Take the same partition \nof the interval $[-T,T]$ into disjoint segments as before. The interaction \nenergy can then be written as\n\\begin{equation}\nW_T(X|Y) = \n\\sum_{0 \\leq i < j \\leq N} W_{\\tau_i,\\tau_j}(X_{\\tau_i},X_{\\tau_j}) + \n\\sum_{0 \\leq i \\leq N-1} W_{\\tau_i,T}^Y (X_{\\tau_i}),\n\\end{equation}\nwith the same notations as before, and with \n\\begin{equation}\nW_{\\tau_i,T}^Y (X_{\\tau_i}) = 2 \\int_{\\tau_i}w^{C^{Y +}_{T}}(X_t,t) dX_t + \n2\\int_{\\tau_i} w^{C^{Y-}_{-T}}(X_t,t) dX_t. \n\\end{equation}\nBy (\\ref{a22}) the estimate \n\\begin{equation}\n\\label{a22y}\n\\sup_{x\\in \\mathbb{R}^d, t \\in \\tau} |w^{C^{Y \\pm}_{\\pm T}}(x,t)| \n\\leq \n\\frac{M_\\tau \\|C^{Y \\pm}_{\\pm T}\\|_{\\mathcal{D}'}}\n{\\left(\\mathop{\\mathrm{dist}}\\nolimits(\\tau,[-T,T]^c) + 1 \\right)^{\\beta-\\alpha}}\n\\end{equation}\neasily follows. (Here $[-T,T]^c = \\mathbb{R} \\backslash [-T,T]$ and \n$\\mathop{\\mathrm{dist}}\\nolimits(\\tau_k,[-T,T]^c) = \\min \\{kb, (N-k-1)b\\}$.) \n\nFix the positions $X_{t_0} = Y^-_{-T} = x_0$, $X_{t_1} = x_1$, ..., \n$X_{t_{N-1}} = x_{N-1}$, $X_{t_N} = Y^+_T = x_N$. Similarly to \n(\\ref{auxi}) introduce the auxiliary measure \n\\begin{equation}\nd\\chi_N^Y = \n\\prod_{k=0}^{N-1} \\frac{e^{-\\lambda W^Y_{\\tau_k,T}(X_{\\tau_k})}}\n{\\Z_{\\tau_k}^T(Y|x_k,x_{k+1})} d\\nu^{x_k,x_{k+1}}_b (X_{\\tau_k}) \n\\prod_{k=1}^{N-1} d\\omega(x_k) \n\\end{equation}\nwhere \n\\begin{equation}\n\\Z^T_{\\tau_k} (Y|x_k, x_{k+1}) = \\mathbb{E}_{\\nu^{x_k,x_{k+1}}_b} \n\\left[ e^{-\\lambda W_{\\tau_k,T}^Y (X_{\\tau_k})} \\right].\n\\end{equation}\nAlso, for every cluster $\\Ga$ consider the function $\\kappa_{\\Ga}^Y$ \ndefined similarly to (\\ref{kappa}). If $\\pm T \\not \\in \\Ga^*$, then \n$\\kappa^Y_\\Ga$ does not depend on $Y$. If $-T \\in \\Ga^*$ and\/or $T \\in \n\\Ga^*$, then $\\kappa^Y_\\Ga$ depends on $Y^-_{-T} = x_0$ and\/or $Y^+_T = \nx_N$, respectively. Next we define the weights\n\\begin{equation}\nK^Y_{\\Ga} = \\mathbb{E}_{\\chi^Y_N} [\\kappa^Y_\\Ga]\n\\end{equation}\nin the same manner as in (\\ref{aux}). The partition function $Z_T(Y)$ \ncan be expressed similarly to (\\ref{clexp}) with these altered objects. \nNote that\n\\begin{enumerate}\n\\item[(1)]\nfor sufficiently small $|\\lambda| \\neq 0$ the cluster estimate \n(\\ref{convv}) obtained in Proposition \\ref{p2} stays essentially \nvalid, i.e., \n\\begin{equation}\n\\label{eq:bound-KY}\n\\sum_{\\Ga: \\Ga^* \\ni 0 \\atop |\\bar \\Ga| = n} |K^Y_\\Ga| \\;\\leq\\; \nc \\eta'(\\lambda) ^n,\n\\end{equation}\nwith $\\eta'(\\lambda)$ going to zero as $\\lambda \\to 0$;\n\\item[(2)]\nfor any fixed $\\Ga$ we have\n$\n\\lim_{T\\to \\infty} \\kappa^Y_\\Ga = \\kappa_\\Ga\n$ and \n\\begin{equation}\n\\label{eq:convergence-KY}\n\\lim_{T\\to \\infty} K^Y_\\Ga = K_\\Ga,\n\\end{equation}\nboth uniformly convergent in $\\mathbb{Y} \\in \\Xi_a$. \n\\end{enumerate}\nThe proof of these statements goes by the same arguments used \nin the previous section and it will be omitted. The \nbound~(\\ref{eq:bound-KY}) can be proven as in Lemma~\\ref{p2}.\nIndeed, by using Lemma~\\ref{lemma:exp-bound} and the bound~(\\ref{a22y}) \non the influence of the boundary current we have a handle to control \nthe exponential moments of $W_{\\tau_i,T}^Y(X_{\\tau_i})$ in terms of \nthe norm of the boundary current and repeat the proof of \nLemma~\\ref{holderw} to obtain the necessary estimates on cluster \nactivities (with constants depending on $a$). A good control of the \nexponential moments is the key to obtain (\\ref{eq:convergence-KY}).\n\n\nThen in (\\ref{dlruniq}) we have\n\\begin{equation}\n\\mathbb{E}_{\\mu_T} [F_S|Y] = \\mathbb{E}_{\\chi^Y_T} \\left[F_S \\right] \n\\; f^S_T(Y) + \\sum_{n\\geq 1} \\sum_{\\{\\Ga_1,...,\\Ga_m\\}: \n\\Ga^*_i \\cap \\Ga_j^* \\neq \\emptyset \\atop \\Ga_i^* \\cap S \\neq 0, \n\\Ga_i \\subset [-T,T], i = 1,...,m} \n\\mathbb{E}_{\\chi^Y_T} \\left [F_S \\prod_{i=1}^m \\kappa^Y_T \\right] \\; \nf_T^{\\cup \\bar\\Ga}(Y)\n\\label{eee}\n\\end{equation}\nwith the same notations as in (\\ref{ce}) and $f^A_T(Y)= Z_T^A(Y)\/Z_T(Y)$.\n\nTake now a collection of intervals $\\{\\tau_i\\} = \\U$; the partition \nfunction $ Z_T^{\\U}(Y) := Z_{T}^{\\cup_{\\tau_i \\in \\U} \\tau^*_i} (Y)$ \ncan then be written like in~(\\ref{clexp}) except for changing $K_\\Ga$ \nfor $K^Y_\\Ga$.\n\\begin{lemma}\nFor sufficiently small $|\\lambda| \\neq 0$ we have the following \nproperties of $f^{\\U}_T(Y) := Z_T^\\U (Y)\/Z_T(Y)$. On the one hand,\n\\begin{equation}\n|f^{\\U}_T(Y)| \\;\\leq\\; 2^{|\\U|},\n\\label{zup}\n\\end{equation}\nwith $|\\U|$ denoting the number of intervals contained in $\\U$. \nOn the other hand,\n\\begin{equation}\n\\lim_{T \\to \\infty} f^{\\U}_T(Y) = f^{\\U}\n\\label{zap}\n\\end{equation}\nuniformly in $\\mathbb{Y} \\in \\Xi_a$. Moreover, $f^{\\U}$ also satisfies \n(\\ref{zup}) above. \n\\end{lemma}\n\\begin{proof}\nBoth statements are direct consequences of the \nbounds (\\ref{eq:bound-KY}) and of (\\ref{eq:convergence-KY})\ntogether with the cluster representation of the correlation \nfunctions~(\\ref{eq:cluster-corr}). By putting\n\\begin{equation}\n\\hat K_\\Ga^Y = \\left\\{ \n\\begin{array}{ll}\nK_\\Ga^Y & \\mbox{if $\\Ga$ lies inside $[-T,T]$} \\nonumber \\\\\n0 & \\mbox{otherwise} \n\\end{array} \\right.\n\\end{equation}\nand using dominated convergence we obtain that $f_{\\U}^Y \\to \nf_{\\U}$ uniformly as $T \\to \\infty$.\n\\end{proof}\n\nWe now return to the expression (\\ref{eee}). By ergodicity \nof the reference measure \n\\begin{equation}\n\\lim_{T\\to \\infty} \\mathbb{E}_{\\chi_N^Y} [F_S] = \n\\mathbb{E}_{\\chi} [F_S], \n\\end{equation}\nand hence the first term of (\\ref{eee}) converges to \n$\\mathbb{E}_{\\chi}[F_S] f_S$. By the same argument as above \nwe also obtain\n\\begin{equation}\n\\lim_{N \\to \\infty} \\mathbb{E}_{\\chi_N^Y} \n\\left[F_S \\prod_{i=1}^m \\kappa_{\\Ga_i}^Y \\right] \n= \n\\mathbb{E}_{\\chi}\\left[F_S \\prod_{i=1}^m \\kappa_{\\Ga_i}\\right] \n\\label{ele}\n\\end{equation}\nuniformly in $Y$, and \n\\begin{equation*}\n|\\mathbb{E}_{\\chi_N^Y} \n[F_S \\prod_{i=1}^m \\kappa_{\\Ga_i}^Y ]| \\;\\leq\\; (\\sup|F_S|) \n\\prod_{i=1}^m |K^Y_{\\Ga_i}|,\n\\label{maa}\n\\end{equation*}\nwhich implies\n\\begin{eqnarray*}\n\\sum_{\\{\\Ga_1,...,\\Ga_m\\} \\atop \\Ga^*_j \\cap S^* = \n\\emptyset, j = 1,...,m} \n\\left ( \\prod_{i=1}^m |K^Y_{\\Ga_i}|\\right) \n2^{|\\cup_i \\bar\\Ga_i|} 2^{|S^*|}\n& \\leq &\n2^{|S^*|} \\sum_{m=1}^{|S^*|} {|S^*| \\choose m} \n\\left( \\sum_{\\Ga: \\Ga \\ni 0} 2^{|\\bar\\Ga|} |K^Y_\\Ga| \\right)^m \\\\\n& \\leq &\n2^{|S^*|} \\sum_{m=1}^{|S^*|} {|S^*| \\choose m} \n\\left( \\sum_{n=2}^\\infty 2^n c \\eta(\\lambda)^n \\right)^m < \\infty.\n\\end{eqnarray*}\nHere $S^*$ denotes the time points occurring in $S$. By using now \nthis estimate together with (\\ref{ele}) and applying Lebesgue's \ndominated convergence theorem once again, we arrive at\n\\begin{equation}\n\\mathbb{E}_{\\chi} [F_S] f_S + \\sum_{m=1}^\\infty \n\\sum_{\\{\\Ga_1,...,\\Ga_m\\} \\atop \\Ga_i^* \\cap S^* \\neq\n\\emptyset, i = 1,...,m} \\mathbb{E}_{\\chi} \n\\left[F_S \\prod_{i=1}^m \\kappa_{\\Ga_i} \\right] \nf_{S \\cup (\\cup_i \\bar\\Ga_i)} = \\mathbb{E}_\\mu [F_S].\n\\end{equation}\n\\end{proof}\n\n\\begin{corollary}\nThe measure $\\mu^\\sharp$ satisfies the DLR equations\n$$\n\\int_\\Xi \\mathbb{E}_{\\rho^\\sharp_T(\\cdot|\\mathbb{Y})}[F_S] \nd\\mu^\\sharp(\\mathbb{Y}) = \\mathbb{E}_{\\mu^\\sharp}[F_S]\n$$ \nfor any $S0} \\Xi_a$ has full $\\mu^\\sharp$-measure so that \nthe left hand side of the equality makes sense. Fix $a>0$. By \nTheorem~\\ref{th:uniq} we have\n$$\n\\int_\\Xi \\mathbb{E}_{\\rho^\\sharp_T(\\cdot|\\mathbb{Y})}[F_S] d\\mu^\\sharp(\\mathbb{Y}) \n= \n\\lim_{T_2\\to+\\infty}\\int_\\Xi \\mathbb{E}_{\\rho^\\sharp_T(\\cdot|\\mathbb{Y})}[F_S] \nd\\rho^\\sharp_{T_2}(\\mathbb{Y}|\\mathbb{Z})\n$$\nuniformly in $Z\\in \\Xi_a$. Then the DLR consistency of the \nspecification implies\n$$\n\\int_\\Xi \\mathbb{E}_{\\rho^\\sharp_T(\\cdot|\\mathbb{Y})}[F_S|\\mathbb{Y}] d\\mu^\\sharp(\\mathbb{Y}) = \n\\lim_{T_2\\to+\\infty} \\mathbb{E}_{\\rho^\\sharp_{T_2}(\\cdot|\\mathbb{Z})}[F_S] \n= E_{\\mu^\\sharp}[F_S],\n$$\nproving the statement.\n\\end{proof}\n\n\n\\subsection{Typical path configurations}\nIn this section we show that most of $\\mu$'s weight is\nconcentrated on paths that can be characterized by a growth \ncondition. The proof of this depends on a lemma which was\nalready shown in \\cite{LM}, however, we include it here for \nmaking the presentation more self-contained. \n\\begin{theorem}\nSuppose (\\ref{iuccond}) holds with $a = b = s$, and\n$\\mu$ is a probability measure obtained by Theorem \\ref{exist} \nfor $V$ and $W$. Then there is a number $C > 0$ and a functional \n$R(X)$, such that \n\\begin{equation}\n|X_t| \\leq \\left(C \\log (|t| + 1)\\right)^{1\/(s+1)} + R(X)\n\\end{equation}\n$\\mu$-almost surely.\n\\label{tpb}\n\\end{theorem}\n\\begin{proof}\nThe strategy of proving this theorem is to derive the typical \nbehaviour of $\\mu$ from the typical behaviour of the reference \nprocess. This follows through Lemma \\ref{ldphi} below. Then\ncombining this lemma with Corollary~\\ref{cor:equiv-mu-nu} gives\n\\begin{equation*}\n \\begin{split}\n\\mu\\left(\\max_{0 \\leq t \\leq 1} |X_t| \\geq a\\right)\n& \\;\\le \\; \nC \\left[\\mathbb{E}_\\nu(1_{\\max_{0 \\leq t \\leq 1} |X_t| \\geq a})\\right]^{1\/2} \n\\le c' e^{-\\theta' a^{s+1}},\n \\end{split}\n\\end{equation*}\nwith some constants $c',\\theta' > 0$. Thus under the stationary \nmeasure $\\mu$ \n\\begin{equation}\n\\mu\\left(\\max_{n \\leq t \\leq n+1} |X_t| \\; \\geq \\; \n(k\\log n)^{1\/(s+1)}\\right) \\; \\leq \\; \\mbox{const} \n\\frac{1}{n^{k \\theta'}}\n\\end{equation}\nholds. Choosing $k$ so that $k \\theta' > 1$, the \nBorel-Cantelli Lemma implies that $\\mu$-almost surely\n\\begin{equation}\n|X(t)| \\; \\leq \\; (k \\log t)^{1\/(s+1)}\n\\end{equation}\nfor $t \\geq T^*$, with $T^* = T^*(X)$ sufficiently large. \nWriting $R(X) = \\max_{|t| \\leq T^*} |X(t)|$ completes the proof. \n\\end{proof}\n\nFinally we prove the lemma used above. \n\\begin{lemma}\nLet $\\nu$ be the measure of the It\\^o-process for $V$ satisfying \n(\\ref{iuccond}) with exponent $s > 2$, and $a > 0$. Then \nthere exist $C > 0$ and $\\theta > 0$ such that\n\\begin{equation}\n\\nu\\left(\\max_{0 \\leq t \\leq 1} |X_t| \\geq a \\right) \n\\; \\leq C \\; e^{-\\theta a^{s+1}}.\n\\end{equation}\n\\label{ldphi}\n\\end{lemma}\n\\begin{proof}\nFor the underlying It\\^o-process we have the Dirichlet operator \non $L^2(\\mathbb{R}^d,d\\nu)$\n\\begin{equation}\nLf = -\\Delta f + 2 (\\nabla \\log \\Psi, \\nabla f)\n\\end{equation}\nand Dirichlet form \n\\begin{equation}\n{\\cal E}(f,f) = - \\int f \\Delta f d\\omega + \n2 \\int f (\\nabla \\log \\Psi, \\nabla f) d\\omega,\n\\end{equation}\nwith $d\\omega = \\Psi^2 dx$, as before. By using Varadhan's Lemma \n(see Lemma 1.12, \\cite{KV86}), for any $f \\in L^2(d\\omega)$ and \nevery $N > 0$ \n\\begin{equation}\n\\nu \\left(\\max_{0\\leq t\\leq 1} |f(X_t)| \\geq N\\right) \n\\; \\leq \\; \\frac{3}{N} \\sqrt{{\\cal E}(f,f) + (f,f)}\n\\label{kv}\n\\end{equation}\nholds. Choose $f = f_a := 1_{\\{x \\in \\mathbb{R}^d: |x| \\geq a\\}} * \\phi$ \nby picking a mollifier $\\phi$ (with $||\\phi||_\\infty < \\infty$) \nso that the above convolution is in the domain of $L$. This can be \nchosen so that the smoothing of the edges of the indicator function \ntakes place in a sphere $S(a)$ of radius $a$ centred at the origin, \ni.e., with a suitable $\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits > 0$ we take $f_a(x) = 1$ for $x \\in \\mathbb{R}^d \n\\backslash S(a+\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)$, $f_a(x) = 0$ for $x \\in S(a-\\mathop{\\mathrm{{\\varepsilon}}}\\nolimits)$, and $f_a$ \nis a sufficiently smooth function ${\\tilde f}_a$ otherwise. Denote \nthese three domains by $D_1$, $D_2$ and $D_3$, respectively. Setting \n$N=1$ in (\\ref{kv}) yields\n$$\n\\nu \\left(\\max_{0 \\leq t \\leq 1} |X_t| \\geq a \\right) \\leq \\; \n3 \\sqrt{||f_a||^2_{L^2(d\\omega)} + (f_a, L f_a)_{L^2(d\\omega)}}.\n$$\nMoreover, we have\n$$\n||f_a||^2_{L^2(d\\omega)} = \\int f_a^2 (x) d\\omega(x) = \n\\int_{D_1} d\\omega(x) + \\int_{D_3} {\\tilde f}_a^2(x) d\\omega(x).\n$$\nUnder the hypothesis on $V$ the standard estimate \n$\n\\Psi(x) \\le C e^{-\\theta |x|^{s+1}}\n$\nholds by Carmona's results \\cite{Ca78} for the ground state $\\Psi$, \nwith some $C,\\theta > 0$. This bound further leads to\n\\begin{equation}\n\\int_{D_1}d\\omega(x) \\; \\leq \\; c e^{-\\theta a^{s+1}},\n\\end{equation}\nwhere $c,\\theta > 0$ are independent of $a$. A similarly estimate is \nvalid for $D_3$. On the other hand, since ${\\tilde f}_a$ is smooth \nenough and $\\max_{D_3} \\{ |\\nabla {\\tilde f}_a|,|\\Delta {\\tilde f}_a|, \n\\Delta {\\tilde f}_a^2 \\} \\leq m < \\infty$, we get\n$$\n({\\tilde f}_a, L{\\tilde f}_a) \\leq c' e^{-\\theta a^{s+1}},\n$$\nwith suitable $c'> 0$. A similar estimate is obtained also for the \nremaining two domains. \n\\end{proof}\n\n\n\n\\subsection{Mixing properties}\nSince $\\mu$ is constructed in a way that offers no immediate \naccess to computations with this measure, it is important to \nderive further basic information on $\\mu$ by using the cluster\nexpansion. We give here one last result of this paper. \n\\begin{theorem}\nLet $F,G$ be two bounded functions, the first measurable with respect \nto $\\mathop{\\mathrm{{\\cal F}}}\\nolimits_I$, the second with respect to $\\mathop{\\mathrm{{\\cal F}}}\\nolimits_J$, where $I,J$ are distinct \nintervals of the partition considered in the cluster expansion above. \nThen the estimate on the covariance \n\\begin{eqnarray*}\n&& \\mathop{\\mathrm{cov}}\\nolimits_\\mu (F;G) = \\mathbb{E}_\\mu[F G] - \\mathbb{E}_\\mu[F] \\; \n\\mathbb{E}_\\mu[G] \\\\\n&& |\\mathop{\\mathrm{cov}}\\nolimits_\\mu \\; (F;G)| \\;\\leq \\; \\mbox{\\rm{const}} \\; \n\\frac{\\sup{|F|} \\sup{|G|}} {|t-s|^\\vartheta + 1}\n\\end{eqnarray*}\nholds, where $\\vartheta > 0$ and the constant prefactor is \nindependent of $F,G$.\n\\end{theorem}\n\n\\medskip\n\\begin{proof}\nFirst recall formula~(\\ref{eq:limit-eq-cluster}) which applied to $F$ \n(and similarly to $G$) gives\n\\begin{equation}\n\\mathbb{E}_\\mu [F] = \\mathbb{E}_\\nu [F] f^{I} + \n\\sum_{\\Ga_0: I \\cap \\Ga_0^* \\neq \\emptyset} K^{\\Ga_0}(F) f^{\\Ga_0^*},\n\\label{F}\n\\end{equation}\nwhere we let $K_{\\Ga_0}(F) = \\mathbb{E}_{\\chi} [F \\kappa_{\\Ga_0}] $. \nFurthermore, consider $f^A$ estimated as before like $|f^A| \\leq \n2^{|A|}$. For $A_1 \\cap A_2 = \\emptyset$ we have\n\\begin{equation}\n|f^{A_1 \\cup A_2} - f^{A_1} f^{A_2}| \\; \\leq \\; \\mbox{const} \\; \n\\frac{2^{|A_1| + |A_2|}}{\\mathop{\\mathrm{dist}}\\nolimits(A_1,A_2)^\\zeta},\n\\label{cost}\n\\end{equation}\nwith some $\\zeta > 0$. This estimate can easily be obtained by the \ngeneral results in \\cite{MM}. Now we write\n\\begin{equation}\n\\begin{split}\n\\mathbb{E}_\\mu [F G] \n& = \n\\mathbb{E}_\\nu [F] \\; \\mathbb{E}_\\nu [G] f^{I \\cup J} \\\\\n& \\qquad + \n\\sum_{\\Ga_1 \\atop \\Ga^*_1 \\cap I \\neq \\emptyset , \\Ga^*_1 \\cap J \n= \\emptyset} \\mathbb{E}_{\\chi} [F \\kappa_{\\Ga_1}] \\; \n\\mathbb{E}_\\chi [G] f^{J \\cup \\Ga^*_1}\n+ \\sum_{\\Ga_2 \\atop \\Ga^*_2 \\cap J \\neq \\emptyset , \\Ga^*_2 \\cap I \n= \\emptyset} \\mathbb{E}_{\\chi} [G \\kappa_{\\Ga_2}] \\; \n\\mathbb{E}_\\chi [F] f^{I \\cup \\Ga^*_2}\n\\\\\n& \\qquad + \n\\sum_{\\Ga_1,\\Ga_2 : \\Ga^*_1 \\cap \\Ga^*_2 = \\emptyset \\atop \\Ga^*_2 \n\\cap J \\neq \\emptyset , \\Ga^*_2 \\cap I \\neq \\emptyset} \n\\mathbb{E}_\\chi [F \\kappa_{\\Ga_1}] \\; \n\\mathbb{E}_{\\chi} [G \\kappa_{\\Ga_2}] f^{\\Ga^*_1 \\cup \\Ga^*_2}\n\\sum_{\\Ga \\atop \\Ga^* \\cap J \\neq \\emptyset , \\Ga^* \\cap I \\neq \n\\emptyset} \\mathbb{E}_\\chi [F G \\kappa_{\\Ga}] f^{\\Ga^*}.\n\\end{split}\n\\end{equation}\nFrom here and (\\ref{F}) we obtain\n\\begin{equation}\n\\begin{split}\n\\mathop{\\mathrm{cov}}\\nolimits_\\mu(F; G) \n& = \n\\mathbb{E}_\\chi [F] \\; \\mathbb{E}_\\chi [G](f^{I\\cup J}-f^{I}f^{J}) \n\\\\ & \\qquad + \n\\sum_{\\Ga_1 \\atop \\Ga^*_1 \\cap I \\neq \\emptyset , \\Ga^*_1 \\cap J \n= \\emptyset} \\mathbb{E}_{\\chi} [F \\kappa_{\\Ga_1}] \\; \n\\mathbb{E}_\\chi [G] (f^{J \\cup \\Ga^*_1}- f^{J} f^{\\Ga^*_1})\n \\\\ & \\qquad \n+ \\sum_{\\Ga_2 \\atop \\Ga^*_2 \\cap J \\neq \\emptyset , \\Ga^*_2 \\cap I \n= \\emptyset} \\mathbb{E}_{\\chi} [G \\kappa_{\\Ga_2}] \\; \n\\mathbb{E}_\\chi [F] (f^{I \\cup \\Ga^*_2}-f^{I}f^{\\Ga^*_2})\n\\\\\n& \\qquad + \n\\sum_{\\Ga_1,\\Ga_2 : \\Ga^*_1 \\cap \\Ga^*_2 = \\emptyset \\atop \\Ga^*_2 \n\\cap J \\neq \\emptyset , \\Ga^*_2 \\cap I \\neq \\emptyset} \n\\mathbb{E}_\\chi [F \\kappa_{\\Ga_1}] \\; \\mathbb{E}_{\\chi} \n[G \\kappa_{\\Ga_2}] (f^{\\Ga^*_1 \\cup \\Ga^*_2}-f^{\\Ga^*_1}f^{\\Ga^*_2})\n \\\\ & \\qquad\n+ \n\\sum_{\\Ga \\atop \\Ga^* \\cap J \\neq \\emptyset , \\Ga^* \\cap I \\neq \n\\emptyset} \\mathbb{E}_\\chi [F G \\kappa_{\\Ga}] f^{\\Ga^*}\n \\\\ &\\qquad\n- \n\\sum_{\\Ga \\atop \\Ga^* \\cap J \\neq \\emptyset , \\Ga^* \\cap I \\neq \n\\emptyset} \\mathbb{E}_\\chi [F ] \\mathbb{E}_\\chi [ G \\kappa_{\\Ga}] \nf^{\\Ga^*} f^I f^{\\Ga^*}\n \\\\ & \\qquad\n- \n\\sum_{\\Ga \\atop \\Ga^* \\cap J \\neq \\emptyset , \\Ga^* \\cap I \\neq \n\\emptyset} \\mathbb{E}_\\chi [F \\kappa_{\\Ga}] \\mathbb{E}_\\chi [ G ] \nf^{\\Ga^*} f^J f^{\\Ga^*}\n \\\\ & \\qquad \n- \\sum_{\\Ga_1,\\Ga_2 : \\Ga^*_1 \\cap \\Ga^*_2 \\neq \\emptyset \\atop \n\\Ga^*_2 \\cap J \\neq \\emptyset , \\Ga^*_2 \\cap I \\neq \\emptyset} \n\\mathbb{E}_\\chi [F \\kappa_{\\Ga_1}] \\; \\mathbb{E}_{\\chi} \n[G \\kappa_{\\Ga_2}] f^{\\Ga^*_1}f^{\\Ga^*_2}\n\\end{split}\n\\end{equation}\nFor estimating the first four terms at the right hand side above we \nuse (\\ref{cost}) along with the bound\n\\begin{equation}\n|\\mathbb{E}_{\\mathop{\\mathrm{{\\cal P}}}\\nolimits} [F \\kappa_{\\Ga}]| \\; \\leq \\; \\frac{\\sup |F|}\n{(\\mathop{\\mathrm{diam}}\\nolimits \\Ga^*)^{\\zeta'} + 1} E_{\\Ga}(\\delta',\\varepsilon')\n\\end{equation}\n$i = 1, 2$, where $E_{\\Ga}(\\delta',\\varepsilon')$ is the function \nappearing at the right hand side of estimate~(\\ref{clustest}) with \nslightly modified entries ($\\delta',\\varepsilon'$ instead of $\\delta,\\varepsilon$; \n$\\delta' > 1$) so that (\\ref{convv}) still holds. Here $\\zeta' = \n\\delta - \\delta' > 0$, and we used in addition that \n\\begin{equation}\n\\frac{1}{(\\mathop{\\mathrm{diam}}\\nolimits \\Ga^*)^{\\zeta'} + 1} \\; \n\\frac{1}{\\mathop{\\mathrm{dist}}\\nolimits(I,\\Ga^*)^\\zeta + 1} \\; \\leq \\; \n\\frac{1}{\\mathop{\\mathrm{dist}}\\nolimits(I,J)^\\vartheta + 1}\n\\end{equation}\nwhenever $J \\cap \\Ga^* \\neq \\emptyset$ (similarly for $I$), and\n\\begin{equation}\n\\frac{1}{(\\mathop{\\mathrm{diam}}\\nolimits \\Ga^*_1)^{\\zeta'} + 1} \\; \n\\frac{1}{(\\mathop{\\mathrm{diam}}\\nolimits \\Ga^*_2)^{\\zeta'} + 1} \\; \n\\frac{1}{\\mathop{\\mathrm{dist}}\\nolimits(\\Ga^*_1,\\Ga^*_2)^\\zeta} \\; \\leq \\; \n\\frac{1}{\\mathop{\\mathrm{dist}}\\nolimits(I,J)^\\vartheta + 1}\n\\end{equation}\nfor $I \\cap \\Ga^*_1 \\neq \\emptyset$, $J \\cap \\Ga^*_2 \\neq \\emptyset$, \nand $\\vartheta = \\min \\{\\zeta,\\zeta'\\} > 0$.\n\nNext, in the fifth term above we use that $\\mathop{\\mathrm{diam}}\\nolimits \\Ga^* \\geq \\mathop{\\mathrm{dist}}\\nolimits(I,J)$ \nwhenever $I \\cap \\Ga^* \\neq \\emptyset$, $J \\cap \\Ga^* \\neq \\emptyset$, \nand that\n\\begin{equation}\n|\\mathbb{E}_{\\chi} [F G \\kappa_\\Ga]| \\; \\leq \\; \\sup|F| \\sup|G| \n\\frac{E_\\Ga(\\delta',\\varepsilon')} {(\\mathop{\\mathrm{diam}}\\nolimits \\Ga^*)^{\\zeta'} + 1}.\n\\end{equation}\nFor the remaining three terms in the sum above we apply the same argument. \nThus for the full sum the corresponding bounds become\n\\begin{equation}\n\\mbox{const} \\; \\frac{\\sup|F| \\sup|G|}{\\mathop{\\mathrm{dist}}\\nolimits(I,J)^\\vartheta + 1} \nE_{\\Ga_1}(\\delta',\\varepsilon') \nE_{\\Ga_2}(\\delta',\\varepsilon') \\; 2^{|\\Ga_1^*| + |\\Ga^*_2|}, \n\\end{equation}\nwhenever $I \\cap \\Ga^*_1 \\neq \\emptyset$, $J \\cap \\Ga^*_2 \\neq \\emptyset$, \nrespectively \n\\begin{equation}\n\\mbox{const} \\; \\frac{\\sup|F| \\sup|G|}{\\mathop{\\mathrm{dist}}\\nolimits(I,J)^\\vartheta + 1} \nE_{\\Ga}(\\delta',\\varepsilon') \\; 2^{|\\Ga^*|} \n\\end{equation}\nin the other cases. Then using Proposition~\\ref{p2} we can prove \nboundedness of the sums over $\\Ga_1, \\Ga_2$ or $\\Ga$, concluding the \nproof.\n\\end{proof}\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\@startsection{subsection}{3{\\@startsection{subsection}{3}\n\t\\z@{.5\\linespacing\\@plus.7\\linespacing}{.5\\linespacing}\n\t{\\bfseries\\itshape}} \\makeatother \n\\renewcommand{\\thesubsection}{\\thesection\\Alph{subsection}}\n\n\n\\makeatletter \\renewenvironment{proof}[1][\\proofname]{\n\t\\par\\pushQED{\\qed}\\normalfont\n\t\\topsep6\\mathfrak{p}@\\@plus6\\mathfrak{p}@\\relax\n\t\\trivlist\\item[\\hskip\\labelsep\\bfseries#1\\@addpunct{.}]\n\t\\ignorespaces}{\n\t\\popQED\\endtrivlist\\@endpefalse} \\makeatother\n\n\\def\\:{\\colon}\n\\numberwithin{equation}{section}\n\n\\renewcommand{\\bar}{\\overline}\n\n\n\\begin{document}\n\n\\title[Factorial affine surfaces with $\\kappa=0$ and trivial units]{Classification of smooth factorial affine surfaces \\\\ of Kodaira dimension zero with trivial units}\n\\subjclass[2010]{Primary: 14R05; Secondary: 14J26, 57R65, 57M99}\n\\keywords{affine surface, $\\mathbb{C}^{*}$-fibration, log Minimal Model program, knot surgery, Kirby diagram}\n\\author{Tomasz Pe\u0142ka}\n\\address{Basque Center for Applied Mathematics, Alameda de Mazarredo 14, 48009 Bilbao, Spain}\n\\email{tpelka@bcamath.org}\n\\author{Pawe\u0142 Ra\u017any}\n\\address{Instytut Matematyki, Uniwersytet Jagiello\u0144ski, ul. \u0141ojasiewicza 6, 30-348 Krak\u00f3w, Poland}\n\\email{pawel.razny@uj.edu.pl}\n\\begin{abstract}\n\tWe give a corrected statement of \\cite[Theorem 2]{GM_k<2}, which classifies smooth affine surfaces of Kodaira dimension zero, whose coordinate ring is factorial and has trivial units. Denote the class of such surfaces by $\\mathcal{S}_{0}$. An infinite series of surfaces in $\\mathcal{S}_{0}$, not listed in \\cite{GM_k<2}, was recently obtained by Freudenburg, Kojima and Nagamine \\cite{FKN_k0} as affine modifications of the plane. We complete their list to a series containing arbitrarily high-dimensional families of pairwise non-isomorphic surfaces in $\\mathcal{S}_{0}$. Moreover, we classify them up to a diffeomorphism, showing that each occurs as an interior of a $4$-manifold whose boundary is an exceptional surgery on a $2$-bridge knot \\cite{BW_surgery}. In particular, we show that $\\mathcal{S}_{0}$ contains countably many pairwise non-homeomorphic surfaces.\n\\end{abstract}\n\n\\maketitle\n\nWe work with complex algebraic varieties.\n\\section{Introduction}\n\nThe structure of smooth affine surfaces of non-general type is essentially understood, see \\cite[Chapter III]{Miyan-OpenSurf}. It can be summarized in a following way, which parallels the classical Enriques-Kodaira classification of smooth projective surfaces. Let $S$ be a smooth affine surface and let $\\kappa\\in \\{-\\infty,0,1,2\\}$ be its Kodaira--Iitaka dimension. If $\\kappa=-\\infty$ then $S$ admits a $\\mathbb{C}^{1}$-fibration \\cite[III.1.3.2]{Miyan-OpenSurf}. If $\\kappa=1$ then $S$ has a (canonical) $\\mathbb{C}^{*}$-fibration \\cite[III.1.7.1]{Miyan-OpenSurf}. The surfaces with $\\kappa=0$ are peculiar, but understandable: they are obtained in a controlled way from some specific minimal models, see \\cite[II.6.4]{Miyan-OpenSurf}. \n\nLet $\\mathcal{S}_{\\kappa}$ be the class of smooth affine surfaces $S$ of Kodaira--Iitaka dimension $\\kappa$, such that $\\mathbb{C}[S]^{*}=\\mathbb{C}^{*}$ and $\\mathbb{C}[S]$ is factorial, i.e.\\ all line bundles on $S$ are trivial. It is known \\cite[III.2.2.1]{Miyan-OpenSurf} that $S\\in \\mathcal{S}_{-\\infty}$ if and only if $S\\cong \\mathbb{C}^{2}$. The above description of affine surfaces with $\\kappa<2$ suggests that it is the class $\\mathcal{S}_{0}$ where one could look for unusual examples.\n\nIn \\cite[Theorem 2]{GM_k<2}, Gurjar and Miyanishi claimed that, up to an isomorphism, there are at most two surfaces in $\\mathcal{S}_{0}$. This result, however, was recently proved wrong by Freudenburg, Kojima and Nagamine \\cite{FKN_k0}, who constructed an infinite series of pairwise non-isomorphic surfaces in $\\mathcal{S}_{0}$. \n\nThe aim of this article is to complete the description of $\\mathcal{S}_{0}$ by proving the following theorem.\n\n\\begin{thm}\\label{THM}\nLet $\\mathcal{S}_{0}$ be the class of smooth affine surfaces $S$ such that the ring $\\mathbb{C}[S]$ is factorial, $\\mathbb{C}[S]^{*}=\\mathbb{C}^{*}$ and the logarithmic Kodaira--Iitaka dimension of $S$ is zero. Then $S\\in \\mathcal{S}_{0}$ if and only if\n\\begin{equation*}\nS\\cong S_{p_1,p_2}:= \\operatorname{Spec} \\mathbb{C}[x_1,x_2][(x_2x_1^{-\\deg p_1}-p_1(x_1^{-1}))x_1^{-1},(x_1x_2^{-\\deg p_2}-p_2(x_2^{-1}))x_2^{-1}]\n\\end{equation*}\nfor some monic polynomials $p_1,p_2\\in \\mathbb{C}[t]$. Moreover, $S_{p_1,p_2}\\not \\cong S_{p_1',p_2'}$ for $\\{p_1,p_2\\}\\neq \\{p_1',p_2'\\}$.\n\\end{thm}\n\n\\begin{rem}\\label{rem:literature}\n\t\\begin{enumerate}\n\t\t\\item\\label{item:GM} The two surfaces constructed in \\cite[Theorem 2]{GM_k<2} are both isomorphic to $S_{1,1}$, see \\cite[Theorem 4.4]{FKN_k0}.\n\t\tThis surface was extensively studied in \\cite[\\S 8]{Kal-Ku1_AL-theory}.\n\t\t\\item\\label{item:FKN} For $n\\geq 1$, the surface $V_{n}=\\operatorname{Spec} \\mathbb{C}[x,y][(x-1)\/(x^{n}y-1)]$ constructed in \\cite{FKN_k0} is isomorphic to $S_{t^{n-1},1}$ via $\\mathbb{C}[x_1,x_2][(x_2-1)x_1^{-n},(x_1-1)x_2^{-1}]\\ni (x_1,x_2)\\mapsto (x,1-x^{n}y)\\in \\mathbb{C}[x,y][(x-1)\/(x^{n}y-1)]$.\n\t\\end{enumerate}\n\\end{rem}\n\n\\begin{figure}[htbp]\n\t\\begin{tabular}{ccc}\n\t\t\\begin{subfigure}[t]{0.35\\textwidth}\n\t\t\t\\includegraphics[scale=0.25]{figures\/S_kl}\n\t\t\t\\caption{$p_1,p_2\\neq 1$}\n\t\t\t\\label{fig:Skl}\n\t\t\\end{subfigure}\n\t\t&\n\t\t\\begin{subfigure}[t]{0.35\\textwidth}\n\t\t\t\\includegraphics[scale=0.25]{figures\/S_k0}\n\t\t\t\\caption{$p_1\\neq 1$, $p_2=1$}\n\t\t\t\\label{fig:Sk0}\n\t\t\\end{subfigure}\n\t\t&\n\t\t\\begin{subfigure}[t]{0.15\\textwidth}\n\t\t\t\\includegraphics[scale=0.25]{figures\/S_00}\n\t\t\t\\caption{$p_1=p_2=1$}\n\t\t\t\\label{fig:S00}\n\t\t\\end{subfigure}\n\t\\end{tabular}\n\t\\caption{Standard boundaries of the surfaces $S_{p_1,p_2}$ from Theorem \\ref{THM}.}\n\t\\label{fig:boundaries}\n\\end{figure}\n\n\\begin{constr}\\label{constr}\n\tIn the remaining part of the article we will use the following geometric construction of the surface $S_{p_{1},p_{2}}$, or rather of its log smooth completion. Let $(x_1,x_2)$ be coordinates on $\\P^{1}\\times \\P^{1}$. For $t\\in \\P^{1}$, $j\\in \\{1,2\\}$ put $\\ll_{j,t}=\\{x_{j}=t\\}$. Write $p_1(t)=t^{d-1}+a_{1}t^{d-2}+\\dots+a_{d-1}$. We perform a sequence of $d$ blowups over $(0,1)$, as follows. The exceptional curve of the first blowup is parametrized by $x_{2}'=(x_2-1)x_1^{-1}\\in \\P^{1}$, where $x_{2}'=\\infty$ lies on the proper transform of $\\ll_{1,0}$. We blow up the point $x_{2}'=a_{1}$, and so on: after $m$-th blowup, the exceptional curve is parameterized by $x_{2}^{(m)}=x_2 x_1^{-m}-x_1^{-m}-a_{1}x_1^{-m+1}-\\dots-a_{m-1}$, and we blow up the point with $x_{2}^{(m)}=a_{m}$. Next, we use $p_2$ to analogously define a sequence of $\\deg p_2+1$ blowups over the preimage of $(1,0)$. We denote the resulting morphism by $\\phi\\colon X\\to \\P^{1}\\times \\P^{1}$, put $L_{j,t}:= \\phi^{-1}_{*}\\ll_{j,t}$ and denote by $A_{1},A_{2}\\subseteq X$ the exceptional curves of the last blowup over $(0,1)$ and $(1,0)$, respectively. Eventually, we put $D=\\phi^{*}(\\ll_{1,\\infty}+\\ll_{2,\\infty}+\\ll_{1,0}+\\ll_{2,0})_{\\mathrm{red}}-A_{1}-A_{2}$ and $S_{p_1,p_2}=X\\setminus D$.\n\t\n\tFigure \\ref{fig:Skl} shows the graph of $D+A_{1}+A_{2}$, where we use the following notation: each line denotes a curve isomorphic to $\\P^{1}$ of specified self-intersection number; the solid ones are the components of $D$, the dashed ones are $A_{1}$ and $A_{2}$; and $[(2)_{m}]$ is short for a chain of $m$ $(-2)$-curves (see Section \\ref{sec:log_surfaces}). \n\t\n\tIf $p_1,p_2\\neq 1$ or $p_1=p_2=1$, then $(X,D)$ is a \\emph{standard} completion of $S_{p_1,p_2}$ in the sense of \\cite{FKZ-weighted-graphs}, see Section \\ref{sec:standard}. To obtain a standard completion in the remaining case, blow up at the preimage of the point $(\\infty,\\infty)$ and contract $L_{1,\\infty}+L_{2,0}$: this gives a graph as in Figure \\ref{fig:Sk0}. In Corollary \\ref{cor:st}\\ref{item:distinct} we use \\cite[Corollary 3.36]{FKZ-weighted-graphs} to prove that $S_{p_1,p_2}\\not\\cong S_{p_1',p_2'}$ for $\\{p_1,p_2\\}\\neq \\{p_1',p_2'\\}$ just by comparing these graphs.\n\\end{constr}\n\nIn Section \\ref{sec:diff} we describe the diffeomorphism types of surfaces in $\\mathcal{S}_{0}$. The result is summarized in Theorem \\ref{thm:diff}. We refer to \\cite[\\S 4,5]{GS_Kirby} for a language of Kirby diagrams (see also Section \\ref{sec:Kirby_overview}) and use the notation of \\cite{BW_surgery} for $2$-bridge knots.\n\n\\begin{thm}\\label{thm:diff}\n\tThe surface $S_{p_{1},p_{2}}$ from Theorem \\ref{THM} is diffeomorphic to the interior of a $2$-handlebody on a $0$-framed $2$-bridge knot $K_{[2d_1,2d_2]}$, where $d_{j}=\\deg p_{j}+1$, $j\\in \\{1,2\\}$. \n\tMoreover, $S_{p_{1},p_{2}}$ is not homeomorphic to $S_{p_{1}',p_{2}'}$ for $\\{\\deg p_{1},\\deg p_{2}\\}\\neq \\{\\deg p_{1}', \\deg p_{2}'\\}$. \n\\begin{figure}[ht]\n\t\\includegraphics[scale=0.35]{figures\/knot}\n\t\\caption{Kirby diagram of a $4$-manifold whose interior is $S_{p_{1},p_{2}}$ with $\\deg p_{j}=d_{j}-1$.}\n\t\\label{fig:knot}\n\\end{figure}\t\n\\end{thm}\nTheorem \\ref{thm:diff} follows by some elementary handle slides from Proposition \\ref{prop:handles}, where we translate Construction \\ref{constr} to the Kirby diagram in Figure \\ref{fig:handles}. \n\t \t \\begin{figure}[ht]\n\t \t \t\\includegraphics[scale=0.4]{figures\/handles}\n\t \t \t\\caption{Kirby diagram equivalent to the one in Figure \\ref{fig:knot}, cf.\\ \\cite[Fig.\\ 5.14]{GS_Kirby}.}\n\t \t \t\\label{fig:handles}\n\t \t \\end{figure}\t\n\t \t \nIn Corollary \\ref{cor:not_homeo} we prove that $S_{p_{1},p_{2}}$ for different degrees of $p_{j}$ are not homeomorphic, because they have different homotopy types at infinity, see \\cite[1.21]{Fujita-noncomplete_surfaces}. Like in the proof of Theorem \\ref{THM}, we deduce it from the graph of $D$, this time applying the calculus of graph $3$-manifolds \\cite{Neumann-plumbing_graphs} to $M=\\d \\mathrm{Tub}(D)$, see \\cite{Mumford-surface_singularities}. That argument becomes more explicit once it is related to the fact that, by Theorem \\ref{thm:diff}, $M$ is a $0$-surgery on $K_{[2d_1,2d_2]}$. We exploit it in Section \\ref{sec:pi_1}, giving a thorough geometric description of $M$. The fundamental group $\\pi_{1}(M)=\\pi_{1}^{\\infty}(S_{p_{1},p_{2}})$ is computed in Proposition \\ref{prop:pi_1}.\t\n\n\\begin{rem}\n\tWe do not know which of the $S_{p_{1},p_{2}}$, for fixed degrees of $p_{1},p_{2}$, are biholomorphic. \n\\end{rem}\n\nAn interesting homogeneity property of $S_{1,1}$ was shown in \\cite[\\S 8]{Kal-Ku1_AL-theory}: although $S_{1,1}$ has no algebraic automorphisms except $(x_1,x_2)\\mapsto(x_2,x_1)$, it admits a lot of nice holomorphic ones, coming from the flows of algebraic vector fields. Theorem \\ref{thm:AAut} shows that the same holds for all $S_{p_{1},p_{2}}$. Denote by $\\mathrm{AAut}_{\\mathrm{hol}}(S_{p_1,p_2})$ the subgroup of $\\operatorname{Aut}_{\\mathrm{hol}}(S_{p_1,p_2})$ generated by elements of flows of complete algebraic vector fields on $S_{p_{1},p_{2}}$. \n\n\\begin{thm}\\label{thm:AAut}\n\tWe keep notation from Construction \\ref{constr}.\n\t\\begin{enumerate}\n\t\t\\item\\label{item:Aut_thm} If $p_1\\neq p_2$ then $\\operatorname{Aut}(S_{p_1,p_2})$ is trivial. Otherwise, $\\operatorname{Aut}(S_{p,p})=\\mathbb{Z}_{2}$ is generated by $(x_1,x_2)\\mapsto (x_2,x_1)$.\n\t\t\\item\\label{item:AAut_thm} The group $\\mathrm{AAut}_{\\mathrm{hol}}(S_{p_{1},p_{2}})$ acts $m$-transitively, for any $m$, on its open orbit $S_{p_1,p_2}\\setminus Z$, where:\n\t\t\\begin{enumerate}\n\t\t\t\\item $Z=\\emptyset$ if $p_{1}=p_{2}=1$,\n\t\t\t\\item $Z=((L_{3-j,1}\\cap A_{j}) \\cup A_{3-j})\\cap S_{p_1,p_2} \\cong \\{\\mathrm{pt}\\}\\sqcup \\mathbb{C}^{1}$ if $p_{j}=1, p_{3-j}\\neq 1$ for some $j\\in \\{1,2\\}$,\n\t\t\t\\item $Z=(A_{1}\\cup A_{2})\\cap S_{p_1,p_2}\\cong \\mathbb{C}^{1}\\sqcup \\mathbb{C}^{1}$ if $p_{1},p_{2}\\neq 1$.\n\t\t\\end{enumerate}\n\t\\end{enumerate}\n\\end{thm}\n\nPart \\ref{item:Aut_thm} is obtained as Corollary \\ref{cor:st}\\ref{item:Aut} from the uniqueness of standard boundary. To prove \\ref{item:AAut_thm} it suffices, by \\cite[Theorem 1.6]{KKL_avf}, to describe all $\\mathbb{C}^{*}$-fibrations of $S_{p_{1},p_{2}}$: this is done in Proposition \\ref{prop:fibrations}. \n\\bigskip\n\n\nWe now explain what is missing in the proof of \\cite[Theorem 2]{GM_k<2}, and how we are going to correct it. The aforementioned result is deduced from the (partial, but sufficient) classification of almost minimal pairs of Kodaira dimension zero given in \\cite[\\S 8]{Fujita-noncomplete_surfaces}. For the general statement see \\cite{Kojima_k0}. \n\nMore precisely, let $S\\in \\mathcal{S}_{0}$ and let $(X_0,D_0)$ be a smooth completion of $S$. Then by the Miyanishi's theory of peeling (see \\cite[p.\\ 107]{Miyan-OpenSurf} or Section \\ref{sec:peeling} below), there is a sequence of birational morphisms\n\\begin{equation*}\n(X_0,D_0)\\toin{\\psi_{1}}(X_{1},D_{1})\\toin{\\psi_{2}}\\dots \\toin{\\psi_{n}} (X_{n},D_{n}),\n\\end{equation*}\nsuch that $(X_{n},D_{n})$ is almost minimal, hence known, and each $\\psi_{i}$ is an snc-minimalization of $D_{i-1}+A$, where $A\\subseteq X_{i-1}$ is a $(-1)$-curve not contained in $D_{i-1}$ such that $A\\cdot D_{i-1}=1$. The latter process is called a \\emph{half-point attachment} in \\cite[8.15]{Fujita-noncomplete_surfaces}. Therefore, to find $S$ it suffices to identify an $(X_{n},D_{n})$ on the list in \\cite[8.70]{Fujita-noncomplete_surfaces} (cf.\\ \\cite[Proposition 1.5]{Kojima_k0}), and perform suitable half-point attachments.\n\nThe proof in \\cite[Theorem 2]{GM_k<2} proceeds exactly along these lines. However, what is meant in loc.\\ cit.\\ by a half-point attachment is just one blowup, the contraction of $A$ (cf.\\ e.g.\\ \\cite[Definition 4.15]{FZ-deformations}). Nonetheless, if $A$ meets a $(-2)$-curve $C\\subseteq D_{i-1}$ which is not branching in $D_{i-1}$, then after the contraction of $A$, the image of $D_{i-1}$ is not snc-minimal, and $\\psi_{i}$ should contract the image of $C$, too. \n\nThis is why in \\cite{GM_k<2} only $S_{1,1}$ is reconstructed. Indeed, it is shown (essentially) that $n=2$, $X_{2}=\\P^{1}\\times \\P^{1}$ and $D_{2}$ is a union of two vertical and two horizontal lines. But $(X,D)\\to (X_{2},D_{2})$ is obtained by two blowups at smooth points of $D_2$, see Figure \\ref{fig:S00}. The smooth completions of $S_{p_1,p_2}$, which are obtained from $(X_{2},D_{2})$ by two \\emph{sequences of blowups} over $D_{2}$, see Figure \\ref{fig:Skl}, are missing.\n\nIn Section \\ref{sec:proof} we repeat the above proof, filling in the missing part. Our approach contains no essentially new ingredient. However, since the assumption $S\\in \\mathcal{S}_{0}$ is substantially stronger than just $\\kappa(S)=0$, we do not need to rely on the structure theorem \\cite[8.70]{Fujita-noncomplete_surfaces}, as it is done in \\cite{GM_k<2}. So for the convenience of the reader, we make our proof more self-contained. The only general result we refer to is the theorem of Kawamata \\cite[2.2]{Kawamata-classification_surfaces} which asserts that if $\\kappa(S)=0$ then the positive part of the Zariski decomposition of $K_{X_{n}}+D_{n}$ is numerically trivial, see \\cite[II.6.2.1]{Miyan-OpenSurf}. \n\\bigskip\n\n\\textbf{Acknowledgments.} We would like to thank the referee for valuable comments, including pointing out a flaw in the earlier version of the proof. The first author would like to thank prof.\\ Frank Kutzschebauch for drawing his attention to the article \\cite{FKN_k0} and for many insightful discussions on the subject.\n\n\n\\section{Preliminaries}\\label{sec:prelims}\n\n\\@startsection{subsection}{3{Log surfaces}\\label{sec:log_surfaces}\n\nWe now briefly recall the language of log surfaces. For a complete introduction we refer to \\cite{Fujita-noncomplete_surfaces}.\n\nBy a \\emph{curve} we mean an irreducible, reduced variety of dimension $1$.\n\nLet $X$ be a smooth projective surface. A curve $C\\subseteq X$ is called an \\emph{$n$-curve} if $C\\cong \\P^{1}$ and $C^{2}=n$. In particular, a $(-1)$-curve is an exceptional curve of a blowup, and a $0$-curve is a fiber of a $\\P^{1}$-fibration.\n\nLet $D$ be an effective $\\mathbb{Q}$-divisor on $X$. By a \\emph{component} of $D$ we always mean an irreducible component. The number of components of $D$ is denoted by $\\#D$. The \\emph{branching number} of a component $C\\subseteq D$ is\n\\begin{equation*}\n\\beta_{D}(C)=C\\cdot (D-C).\n\\end{equation*}\nWe say that $C$ is a \\emph{tip} of $D$ if $\\beta_{D_{\\mathrm{red}}}(C)=1$, and \\emph{branching} in $D$ if $\\beta_{D_{\\mathrm{red}}}(C)\\geq 3$. \n\nAssume that $D$ is \\emph{simple normal crossing} (\\emph{snc}), that is, $D$ is reduced, all components of $D$ are smooth and meet transversally, at most two at each point. A $(-1)$-curve $C\\subseteq D$ is called \\emph{superfluous} if $1\\leq \\beta_{D}(C)\\leq 2$ and $C$ meets each component of $D$ at most once: in other words, after contraction of $C$, the image of $D$ remains snc. We say that $D$ is \\emph{snc-minimal} if it contains no superfluous $(-1)$-curves. An \\emph{snc-minimalization} of $D$ is the contraction of all superfluous $(-1)$-curves in $D$ and its images.\n\nAssume further that $D$ has connected support and $\\beta_{D}(C)\\leq 2$ for all components $C$ of $D$. We say that $D$ is a \\emph{chain} if at least one inequality is strict, otherwise we say that $D$ is \\emph{circular}. We order the components $T_{1},\\dots, T_{m}$ of $D$ in such a way that $T_{i}\\cdot T_{i+1}=1$ for $i\\in \\{1,\\dots, m-1\\}$, and $T_{1}$ is a tip of $D$ in case $D$ is a chain. The sequence of integers $(-T_{1}^{2},\\dots, -T_{m}^{2})$ is then called a \\emph{type} of $D$. We often abuse notation and write \n\\begin{equation*}\nD=[-T_{1}^{2},\\dots,-T_{m}^{2}]\\mbox{ if $D$ is a chain and } D=(\\!( -T_{1}^{2},\\dots, -T_{m}^{2})\\!) \\mbox{ if $D$ is circular}.\n\\end{equation*}\nA sequence consisting of an integer $a$ repeated $k$ times will be abbreviated as $(a)_{k}$.\n\nAn effective divisor $D$ is called a \\emph{tree} if $D_{\\mathrm{red}}$ is snc and has no circular subdivisor. A \\emph{fork} is a tree with exactly one branching component and three maximal twigs. A \\emph{rational} tree (chain, fork...) is a tree (chain, fork...) whose components are all rational.\n\n\nWe say that a reduced divisor $D=\\sum D_{i}$ is \\emph{negative definite} if its intersection matrix $[D_{i}\\cdot D_{j}]_{1\\leq i,j\\leq r}$ is.\n\nLet again $D$ be an snc divisor. A chain $T\\subseteq D$ is a \\emph{twig} of $D$ if it contains a tip of $D$ and its components are non-branching in $D$. We order a twig in such a way that its first component is a tip of $D$. A twig is called \\emph{admissible} if all its components are rational and have self-intersection number at most $-2$. An (admissible) twig of $D$ is \\emph{maximal} if it is maximal in the set of (admissible) twigs ordered by inclusion of supports. A \\emph{$(-2)$-twig} is a twig whose components are $(-2)$-curves.\n\nLet $T$ be an admissible twig of $D$. Then $T$ is negative definite, so there is a unique $\\mathbb{Q}$-divisor $\\operatorname{Bk}_{D}(T)\\leq T$, called the \\emph{bark} of $T$, such that for every component $T_{0}$ of $T$ \n\\begin{equation*}\nT_{0}\\cdot\\operatorname{Bk}_{D}(T)=T_{0}\\cdot (K_{X}+D),\n\\end{equation*}\nthat is, $T_{0}\\cdot \\operatorname{Bk}_{D}(T)=-1$ if $T_{0}$ is a tip of $D$ and $T_{0}\\cdot \\operatorname{Bk}_{D}(T)=0$ otherwise. An important result for the theory of peeling asserts that the coefficients of $\\operatorname{Bk}_{D}(T)$ are strictly between $0$ and $1$ \\cite[II.3.3]{Miyan-OpenSurf}.\n\nAssume that $D$ is a connected snc-minimal divisor which is not negative definite. Then $\\operatorname{Bk} D$ is defined as the sum of barks of all maximal admissible twigs of $D$. We put $D^{\\#}=D-\\operatorname{Bk} D$.\n\\smallskip\n\nLet $D$ be an snc divisor on $X$. We say that a blowup $X'\\to X$ at a point $p\\in D$ is \\emph{outer} with respect to $D$ if $p$ is a smooth point of $D$, otherwise it is called \\emph{inner}. We say that a birational map $X\\dashrightarrow Y$ \\emph{touches} a divisor if it is not an isomorphism in any of its neighborhoods.\n\\smallskip\n\n\nA \\emph{log smooth} pair $(X,D)$ consists of a smooth projective surface $X$ and an snc divisor $D$ on $X$. Any smooth surface $S$ admits a \\emph{log smooth completion}, that is, a log smooth pair $(X,D)$ such that $S\\cong X\\setminus D$. We say that such a completion is \\emph{minimal} if it does not dominate birationally any other log smooth completion, or, equivalently, if $D$ is snc-minimal.\n\n\\@startsection{subsection}{3{Affine surfaces whose coordinate rings are factorial and have trivial units}\n\n\nFor a divisor $D$ on a smooth surface $X$, we denote by $\\mathbb{Z}[D]$ the free abelian group whose basis is the set of components of $D$. There is a natural group homomorphism $\\mathbb{Z}[D]\\to\\operatorname{Pic}(X)$. The following lemma is well known, cf.\\ \\cite[1.17]{Fujita-noncomplete_surfaces} or \\cite[Lemma 2.2]{GM_k<2}. \n\n\\begin{lem}[Criterion for $S\\in \\mathcal{S}_{\\kappa}$]\\label{lem:S0}\n\tLet $(X,D)$ be a smooth completion of a smooth affine surface $S$.\n\t\\begin{enumerate}\n\t\t\\item\\label{item:UFD} $\\mathbb{C}[S]$ is a UFD if and only if $\\mathbb{Z}[D]\\to\\operatorname{Pic}(X)$ is surjective.\n\t\t\\item\\label{item:units} $\\mathbb{C}[S]^{*}=\\mathbb{C}^{*}$ if and only if $\\mathbb{Z}[D]\\to\\operatorname{Pic}(X)$ is injective.\n\t\\end{enumerate}\n\tAssume that $S\\in \\mathcal{S}_{\\kappa}$. Then $\\mathbb{Z}[D]=\\operatorname{Pic}(X)=\\operatorname{NS}(X)$ and $q(X)=0$. If $\\kappa<2$ then $X$ is rational.\n\\end{lem}\n\\begin{proof}\n\t\\ref{item:UFD} By \\cite[II.6.2]{Hartshorne_AG}, $\\mathbb{C}[S]$ is factorial if and only if the divisor class group $\\mathrm{Cl}(S)$ of $S$ is trivial. It follows from \\cite[II.6.5(c)]{Hartshorne_AG} that $\\mathrm{Cl}(S)$ is the cokernel of the natural map $\\mathbb{Z}[D]\\to\\mathrm{Cl}(X)$. Eventually, $\\mathrm{Cl}(X)=\\operatorname{Pic}(X)$ because $X$ is smooth \\cite[II.6.16]{Hartshorne_AG}.\n\t\n\t\\ref{item:units} A regular function on $S$ is invertible if and only if it extends to $f\\in \\mathbb{C}(X)$ with all zeros and poles in $D$, that is, with $\\mathrm{div}(f)\\in \\mathbb{Z}[D]$. Two rational functions on $X$ with the same divisor differ by an element of $\\O_{X}(X)^{*}=\\mathbb{C}^{*}$, so $f\\mapsto \\mathrm{div}(f)$ gives an isomorphism of $\\mathbb{C}[S]^{*}\/\\mathbb{C}^{*}$ with the kernel of $\\mathbb{Z}[D]\\to \\operatorname{Pic}(X)$.\n\t\n\tFor the remaining part, we follow \\cite[Lemma 2.2]{GM_k<2}. Assume that $S\\in \\mathcal{S}_{\\kappa}$. Then \\ref{item:UFD}, \\ref{item:units} give $\\mathbb{Z}[D]=\\operatorname{Pic}(X)$. The exponential sequence induces an exact sequence\n\t\\begin{equation*}\n\t0\\to H^{1}(X,\\mathbb{Z})\\to H^{1}(X,\\O_{X})\\to \\operatorname{Pic}(X)\\to H^{2}(X,\\mathbb{Z}).\n\t\\end{equation*}\n\tSince $H^{1}(X,\\O_{X})$ is a vector space over $\\mathbb{C}$, and by \\ref{item:UFD} $\\operatorname{Pic}(X)$ is a finitely generated abelian group, we obtain $q(X)=h^{1}(X,\\O_{X})=0$ and that $\\operatorname{Pic}(X)\\to H^{2}(X,\\mathbb{Z})$ is injective, i.e.\\ $\\operatorname{Pic}(X)=\\operatorname{NS}(X)$. Now to prove that $X$ is rational, it remains to show that $p_{g}(X)=0$ \\cite[IV.6.2]{Hartshorne_AG}. Assume $\\kappa<2$. Because $S$ is affine, $D$ supports an effective ample divisor, so if $|K_{X}|\\neq \\emptyset$ then for $m\\gg 0$ we have $h^{0}(m(K_{X}+D))\\sim m^{2}$, hence $\\kappa=2$; a contradiction.\n\\end{proof}\n\n\\begin{rem}\\label{rem:open_UFD}\n\tLemma \\ref{lem:S0}\\ref{item:UFD} implies that an open affine subset of a smooth factorial surface is factorial, and that the converse holds if and only if its complement is a principal divisor. This is a special case of the Nagata lemma, see e.g.\\ \\cite[Lemma 19.20]{Eisenbud_comm-alg}\n\\end{rem}\n\n\\begin{cor}[$S_{p_1,p_2}\\in \\mathcal{S}_{\\kappa}$]\\label{cor:S_k}\n\tLet $(X,D)$ be the log smooth completion of $S_{p_1,p_2}$ as in Construction \\ref{constr}. Then $\\operatorname{Pic}(X)=\\mathbb{Z}[D]$. In particular, $\\mathbb{C}[S_{p_1,p_2}]$ is a UFD and $\\mathbb{C}[S_{p_1,p_2}]=\\mathbb{C}^{*}$.\n\\end{cor}\n\\begin{proof}\n\tWe use the notation introduced in Construction \\ref{constr}. The group $\\operatorname{Pic}(X)$ is generated by the components of $\\phi^{*}(\\ll_{1,\\infty}+\\ll_{1,0}+\\ll_{2,\\infty}+\\ll_{2,0})_{\\mathrm{red}}=D+A_{1}+A_{2}$, with relations $0=\\phi^{*}(\\ll_{j,0}-\\ll_{j,\\infty})=A_{j}+T_{j}+L_{j,0}-L_{j,\\infty}$, $j\\in \\{1,2\\}$, where $T_{j}$ is the twig of $D$ meeting $L_{j,0}$. Therefore, $\\operatorname{Pic}(X)=\\mathbb{Z}[D+A_{1}+A_{2}]\/\\mathbb{Z}[A_{1}+A_{2}]=\\mathbb{Z}[D]$. The remaining assertion follows from Lemma \\ref{lem:S0}.\n\\end{proof}\n\n\\@startsection{subsection}{3{$\\P^{1}$-fibrations}\n\nA \\emph{fibration} of a smooth surface $S$ is a surjective morphism onto a curve whose general fiber is irreducible and reduced. A curve on $S$ is called \\emph{vertical} (resp.\\ \\emph{horizontal}) if $C\\cdot F=0$ (resp.\\ $C\\cdot F>0$) for a fiber $F$; it is called an \\emph{$n$-section} if $C\\cdot F=n$. Any divisor $D$ decomposes uniquely as $D=D\\vert+D_{\\mathrm{hor}}$, where all components of $D\\vert$ are vertical and all components of $D_{\\mathrm{hor}}$ are horizontal.\n\nA $\\P^{1}$- (resp.\\ $\\mathbb{C}^{1}$-, $\\mathbb{C}^{*}$-) fibration is a fibration whose general fiber is isomorphic to $\\P^{1}$ (resp.\\ $\\mathbb{C}^{1}$, $\\mathbb{C}^{*}:= \\mathbb{C}^{1}\\setminus \\{0\\}$). A fiber not isomorphic to a general one will be called \\emph{degenerate}. \n\nThe surfaces in $\\mathcal{S}_{0}$ do not admit $\\mathbb{C}^{1}$-fibrations by the Iitaka Easy Addition Theorem \\cite[Theorem 11.9]{Iitaka_AG}, but they admit plenty of $\\mathbb{C}^{*}$-fibrations (eg.\\ produced by Lemma \\ref{lem:producing_Cst}\\ref{item:HP_Cst}), which we will use in Proposition \\ref{prop:fibrations}\\ref{item:vf} as first integrals for complete algebraic vector fields, see \\cite[Theorem 1.6]{KKL_avf}. The structure of $\\mathbb{C}^{*}$-fibrations is well described in \\cite[\\S 7]{Fujita-noncomplete_surfaces}, however, for our purposes it will be more convenient to study them directly, by completing them to $\\P^{1}$-fibrations. \n\nLet $F$ be a degenerate fiber of a $\\P^{1}$-fibration of a smooth projective surface. Then $F$ is obtained from a $0$-curve by a sequence of blowups, hence its geometry is easy to understand, see eg.\\ \\cite[\\S 4]{Fujita-noncomplete_surfaces}. In particular, $F$ is a rational tree with no branching $(-1)$-curves; moreover, if a $(-1)$-curve has multiplicity $1$ in $F$ then it is a tip of $F$, and $F$ contains another $(-1)$-curve \\cite[7.3]{Fujita-noncomplete_surfaces}.\n\nThe following observation, see \\cite[7.14(1)]{Fujita-noncomplete_surfaces}, will be used several times.\n\n\\begin{lem}[our $\\mathbb{C}^{*}$-fibrations are untwisted]\\label{lem:untwisted}\n\tLet $(X,D)$ be a log smooth completion of a smooth affine surface $S$ such that $\\mathbb{C}[S]$ is a UFD. Let $H_{1},\\dots, H_{r}$ be all components of $D_{\\mathrm{hor}}$ for some $\\P^{1}$-fibration of $X$. Put $h_j=H_{j}\\cdot F$ for a fiber $F$. Then $\\gcd(h_{1},\\dots, h_{r})=1$.\n\\end{lem}\n\\begin{proof}\n\tBy Tsen's theorem, $X$ contains a $1$-section $H$. By Lemma \\ref{lem:S0}\\ref{item:UFD}, $H\\equiv V+\\sum_{j=1}^{r} a_{j}H_{j}$ for some vertical $V$ and $a_{1},\\dots, a_{r}\\in \\mathbb{Z}$. Intersecting with $F$ gives $1=\\sum_{j=1}^{r} a_{j}h_{j}$, so $\\gcd(h_1,\\dots, h_r)=1$.\n\\end{proof}\n\n\n\n\n\n\\@startsection{subsection}{3{Standard completions and elementary transformations}\\label{sec:standard}\n\nA minimal log smooth completion of an open surface may not be unique. However, \\cite[Definition 2.13]{FKZ-weighted-graphs} distinguishes a class of \\emph{standard completions}, which gives a convenient tool to tell affine surfaces apart. We now recall this definition. A rational chain is \\emph{standard} if it is of type\n\\begin{equation*}\n[(0)_{2k+1}]\\quad\\mbox{or}\\quad[(0)_{2k},a_{1},\\dots, a_{l}]\\quad\\mbox{for some integers }k,l\\geq 0\\mbox{ and }a_{1},\\dots,a_{l}\\geq 2.\n\\end{equation*}\nA rational circular divisor is \\emph{standard} if it is of type\n\\begin{equation*}\n(\\!((0)_{2k},a_{1},\\dots, a_{l})\\!)\\mbox{ or }\n(\\!((0)_{k},a)\\!)\\mbox{ or }\n(\\!((0)_{2k},1,1)\\!)\\quad\\mbox{for some }k,l\\geq 0,\\ a\\geq 0\\mbox{ and }a,a_{1},\\dots,a_{l}\\geq 2.\n\\end{equation*}\nLet now $D$ be any snc divisor, and let $B$ be the sum of all components of $D$ which are either branching or non-rational; so every connected component of $D-B$ is rational, chain or circular. We say that $D$ is standard if every connected component of $D-B$ is standard in the above sense. A log smooth completion $(X,D)$ is \\emph{standard} if $D$ is. Note that the log smooth completions of $S_{p_1,p_2}$ in Figure \\ref{fig:boundaries} are standard.\n\\smallskip\n\nLet $D$ be an snc divisor on $X$, and let $C\\subseteq D$ be a non-branching $0$-curve. If $\\beta_{D}(C)=2$ choose $p\\in C\\cap (D-C)$, otherwise let $p$ be any point of $C$. A \\emph{flow} (or \\emph{elementary transformation}) \\emph{on} $C$ is the birational map $(X,D)\\dashrightarrow (X',D')$, where $D'$ is the reduced total transform of $D$, defined as a blowup at $p$, followed by the contraction of the proper transform of $C$. We remark that in Section \\ref{sec:AAut} we will use the word \\enquote{flow} in its standard meaning, i.e.\\ flow of a vector field: this will lead to no confusion.\n\nLet $\\phi\\colon (X,D)\\dashrightarrow (X',D')$ be a flow on $C\\subseteq D$. Then $\\phi|_{X\\setminus D}\\colon X\\setminus D\\to X'\\setminus D'$ is an isomorphism. In particular, if the components of $D$ generate $\\operatorname{Pic}(X)$ (respectively, are $\\mathbb{Z}$-linearly independent in $\\operatorname{Pic}(X)$), then by Lemma \\ref{lem:S0} the same is true for the components of $D'$ in $\\operatorname{Pic}(X')$. Moreover, $\\phi$ does not touch components of $D$ other than $C$ and the components of $D$ meeting $C$. If $C$ is not a tip of $D$, then $\\phi$ replaces a subchain $[a,0,b]\\subseteq D$, whose middle component is $C$, by $[a+1,0,b-1]\\subseteq D'$.\n\\smallskip\n\nWe are ready to formulate the result used to distinguish the surfaces $S_{p_1,p_2}$ for different $\\{p_1,p_2\\}$.\n\n\\begin{lem}[{\\cite[Corollary 3.36]{FKZ-weighted-graphs}}]\\label{lem:boundaries}\n\tAny smooth affine surface admits a standard log smooth completion. Any two such completions differ by a sequence of flows on some $0$-curves in the boundary.\n\\end{lem}\n\n\\begin{cor}[$S_{p_1,p_2}$ are non-isomorphic]\\label{cor:st}\n\tLet $S_{p_1,p_2}$ be as in Theorem \\ref{THM}. Then\n\t\\begin{enumerate}\n\t\t\\item\\label{item:st_uniq} The standard completion $(X,D)$ of $S_{p_1,p_2}$ is unique up to an isomorphism, with $D$ as in Figure \\ref{fig:boundaries}.\n\t\t\\item\\label{item:distinct} The surfaces $S_{p_1,p_2}$ and $S_{p_1',p_2'}$ are not isomorphic unless $\\{p_1,p_2\\}=\\{p_1',p_2'\\}$.\n\t\t\\item \\label{item:Aut} If $p_1\\neq p_2$ then $\\operatorname{Aut}(S_{p_1,p_2})$ is trivial. Otherwise, $\\operatorname{Aut}(S_{p,p})=\\mathbb{Z}_{2}$ is generated by $(x_1,x_2)\\mapsto (x_2,x_1)$.\n\t\\end{enumerate}\n\\end{cor}\n\\begin{proof}\n\t\\ref{item:st_uniq} By Lemma \\ref{lem:boundaries}, $D$ differs from the one in Figure \\ref{fig:boundaries} by a sequence of flows. Because all twigs of the one in Figure \\ref{fig:boundaries} are admissible, they are not touched by any flow. Since a flow preserves the non-weighted graph and the sum of self-intersection numbers; we only need to check that the latter agree.\n\t\n\tConsider the case when $p_1,p_2\\neq 1$ or $p_1=p_2=1$. Then any such $D$ contains a rational circular divisor $L$ with components, say, $L_{1},L_{2},L_{3},L_{4}$, such that $L_{i}\\cdot L_{i+1}=1$ for $i\\in \\{1,2,3\\}$; if $p_{j}\\neq 1$ for $j\\in \\{1,2\\}$ then $L_{2+j}$ meets a twig of $D$ of type $[(2)_{\\deg p_j}]$, and $L_{1}^{2}=0$, i.e.\\ $L_{1}$ is the $0$-curve on which the last flow is performed. If $p_1,p_2\\neq 1$ then the chain $L_{1}+L_{2}$ is standard, and if $p_1=p_2=1$ then the whole $L$ is standard. In any case, we infer that $L_{2}^{2}=0$. It follows that $L_{3}^{2},L_{4}^{2}<0$: indeed, for $i\\in \\{1,2\\}$, $L_{i+2}$ is vertical for the $\\P^{1}$-fibration of $X$ induced by $|L_{i}|$, and it is not a fiber because otherwise $L_{i+2}=L_{i}$ in $\\operatorname{Pic}(X)$, contrary to Lemma \\ref{lem:S0}\\ref{item:units}. Because the sum $L_{1}^{2}+L_{2}^{2}+L_{3}^{2}+L_{4}^{2}$ is not changed by a flow, we get $L_{3}^{2}+L_{4}^{2}=-2$, so $L_{3}^{2}=L_{4}^{2}=-1$, as claimed.\n\t\n\tIn case $p_j\\neq 1$, $p_{3-j}= 1$ we similarly obtain that $D$ has a rational circular divisor $L_{1}+L_{2}+L_{3}$, where $L_{3}$ meets a twig $[(2)_{\\deg p_j}]$ and $L_{1}$ is a $0$-curve. The chain $L_{1}+L_{2}$ is standard, so $L_{2}^{2}=0$, and $L_{3}^{2}=L_{1}^{2}+L_{2}^{2}+L_{3}^{2}=1$.\n\t\n\t\\ref{item:distinct}, \\ref{item:Aut} Let $S_{p_1,p_2}\\to S_{p_1',p_2'}$ be an isomorphism. By Lemma \\ref{lem:boundaries}, it extends to an isomorphism, say $\\tau\\colon (X_1,D_1)\\to(X_2,D_2)$, between their standard completions, which by \\ref{item:st_uniq} are both as in Construction \\ref{constr}. If they are as in Figure \\ref{fig:Sk0}, we reverse the last step of Construction \\ref{constr} and lift $\\tau$ accordingly, so that $(X_{i},D_{i})$ are as in Figure \\ref{fig:Skl} or \\ref{fig:S00} (but possibly no longer standard). \n\t\n\tFor $i,j\\in \\{1,2\\}$ let $A_{j,i}\\subseteq X_{i}$ be as in Figure \\ref{fig:boundaries}. Because by Lemma \\ref{lem:S0} $\\operatorname{Pic}(X_{i})=\\mathbb{Z}[D_{i}]$, the curve $\\tau(A_{j,1})$ is linearly equivalent to $A_{j,2}$. In fact, $\\tau(A_{j,1})=A_{j,2}$ because $\\tau(A_{j,1})\\cdot A_{j,2}=A_{j,2}^{2}<0$. Therefore, $\\tau$ descends to an automorphism $\\tau_{0}$ of $\\P^{1}\\times \\P^{1}$ fixing $\\ll_{1,0}+\\ll_{1,\\infty}+\\ll_{2,0}+\\ll_{2,\\infty}$ and the points $(0,1)$, $(1,0)$, so $\\tau_{0}(x_1,x_2)=(x_1,x_2)$ or $(x_2,x_1)$. Hence $\\tau$ or $\\tau\\circ \\epsilon$, where $\\epsilon(x_1,x_2)=(x_2,x_1)$, is trivial on each exceptional curve of $\\phi\\colon X\\to \\P^{1}\\times \\P^{1}$. This shows that $\\{p_1,p_2\\}=\\{p_1',p_2'\\}$ and $\\tau=\\mathrm{id}$ or $\\epsilon$.\n\\end{proof}\n\n\\begin{rem}[{$S_{p_1,p_2}$ are stably non-isomorphic, cf.\\ \\cite[Corollary 4.3]{FKN_k0}}]\n\tBecause $\\kappa(S_{p_1,p_2})=0$ by Corollary \\ref{cor:S_kappa} below, \\cite{Iitaka_Fujita-cancellation} implies that $S_{p_1,p_2}\\times \\mathbb{C}^{n}\\not\\cong S_{p'_1,p'_2}\\times \\mathbb{C}^{n}$ for any $n\\geq 0$ and $\\{p_1,p_2\\}\\neq \\{p_1',p_2'\\}$.\n\\end{rem}\n\nWe conclude with a well-known application of flows, which will be useful in the proof of Lemma \\ref{lem:producing_Cst}.\n\n\\begin{lem}[{\\cite[6.13]{Fujita-noncomplete_surfaces}, cf.\\ \\cite[Lemma 4.14]{FZ-deformations}}]\\label{lem:rational-twigs-are-admissible}\n\tLet $(X,D)$ be a log smooth pair with $\\kappa(X\\setminus D)\\geq 0$. Let $T$ be a rational twig of $D$. Then $C^2\\leq -1$ for every component $C$ of $T$. In particular, if $D$ is snc-minimal then all rational twigs of $D$ are admissible.\n\\end{lem}\n\\begin{proof}\n\tLet $T_{k}$ be the $k$-th component of $T$, put $T_0=0$. Suppose $T_{k}^{2}\\geq 0$ for some $k\\in \\{1,\\dots, \\#T\\}$. Write $\\{p\\}=T_{k}\\cap (D-T_{k}-T_{k-1})$, or choose $p\\in T_{k}\\setminus T_{k-1}$ if that intersection is empty. Blowing up $T_{k}^{2}$ times at $p$ and its infinitely near points on the proper transforms of $T_{k}$, we can assume that $T_{k}=[0]$. If $k>1$, then after $\\pm T_{k-1}^{2}$ flows on $T_{k}$, we can assume $T_{k-1}=[0]$, too. Thus by induction, we can assume $k=1$, i.e.\\ $T_{1}=[0]$ is a tip of $D$. Then $|T_{1}|$ induces a $\\P^{1}$-fibration of $X$ which restricts to a $\\mathbb{C}^{1}$-fibration of $X\\setminus D$, so $\\kappa(X\\setminus D)=-\\infty$ by the Iitaka Easy Addition theorem; a contradiction.\n\\end{proof}\n\n\\@startsection{subsection}{3{Graph $3$-manifolds and their normal forms}\\label{sec:graph_manifolds}\n\nAn important topological invariant of an affine surface $S$ is its \\emph{fundamental group at infinity}, $\\pi_{1}^{\\infty}(S):=\\varprojlim \\pi_{1}(S\\setminus K)$, where $\\varprojlim$ runs over all compact $K\\subseteq S$. If $(X,D)$ is a log smooth completion of $S$, then $\\pi_{1}^{\\infty}(S)=\\pi_{1}(M)$, where $M$ is the boundary of a nice tubular neighborhood of $D$, constructed in \\cite{Mumford-surface_singularities} as a plumbed manifold whose graph is the dual graph of $D$, see \\cite[4.6.2]{GS_Kirby}. More precisely, $\\mathrm{Tub}(D)$ is a union, taken over components $C$ of $D$, of disk bundles $\\xi_C$ with Euler numbers $C^2$, where for any component $C'$ of $D$ meeting $C$, a fiber of $\\xi_C$ over $C\\cap C'$ is glued to a neighborhood of $C\\cap C'$ in $C$. \n\nFor a general introduction to $3$-manifolds we refer to \\cite{Hatcher_3M,AFW_3M}. \nWe denote by $\\mathbb{S}^{k}$, $\\mathbb{D}^{k}$ the (real) $k$-dimensional sphere and disk, respectively, and by $\\mathbb{T}^{k}=(\\mathbb{S}^{1})^{k}$ the $k$-dimensional torus. We say that a $3$-manifold is \\emph{prime} if it cannot be written as a connected sum of two $3$-manifolds other than $\\mathbb{S}^{3}$, and \\emph{irreducible} if any embedded $\\mathbb{S}^{2}$ bounds a ball. A prime, oriented $3$-manifold is either irreducible or $\\mathbb{S}^{1}\\times \\mathbb{S}^{2}$.\n\nIf $M$ is prime, and not a lens space (which will be the case for $S=S_{p_{1},p_{2}}$), then $\\pi_{1}(M)$ determines $M$ uniquely, up to a diffeomorphism \\cite[Theorem 2.1.2]{AFW_3M}. Therefore, in this case to distinguish the homeomorphism type of $S$ at infinity, it suffices to describe $\\pi_1(M)$, which, in turn, is given by the graph of $D$, once put in a normal form \\cite[Theorem 4.2]{Neumann-plumbing_graphs}. The definition of normal form is similar to the one of standard boundary, but requires some more flexibility to allow more operations between $3$-manifolds than just blowing up and down. We now recall the definition of graph $3$-manifolds following \\cite{Neumann-plumbing_graphs}. We refer to \\cite{EN_torus-links} for details and relation to the JSJ-decomposition (graph manifolds are exactly those for which the latter consists only of Seifert fibered spaces: we compute it for our $M$ in Proposition \\ref{prop:JSJ}).\n\nA \\emph{graph manifold} is a $3$-manifold which is a union of $\\mathbb{S}^{1}$-bundles over compact surfaces, glued according to some graph, as follows. An $\\mathbb{S}^{1}$-bundle over $C$ with Euler number $e$ is represented by a vertex of weight $e$ and two additional numbers $g$, $r$, where $g$ is the genus of $C$ ($g<0$ if $C$ is non-orientable) and $r=b_{0}(\\d C)$. They are skipped whenever both are zero, i.e.\\ $C\\cong \\P^{1}$. An edge between two vertices $v_{1}$, $v_{2}$ represents gluing of a fiber of $v_{1}$ to a small loop around a point in a base of $v_2$, and vice versa. Multiple edges and loops are allowed. An edge is labeled by \\enquote{$+$} if the gluing respects the chosen orientations, and \\enquote{$-$} if it reverses both of them. Clearly, reversing the sign of all edges adjacent to a single vertex, or of any edge adjacent to a vertex with $g<0$, does not change the manifold. Hence we will label only the edges contained in circular subgraphs with no vertices of $g<0$, see \\cite[p.\\ 304]{Neumann-plumbing_graphs}.\n\nFor example, the boundary $M$ from \\cite{Mumford-surface_singularities}, described above, is a graph manifold associated to the dual graph of $D$, with all edges labeled with \\enquote{$+$}. For such graph, we use the notions of branching numbers, chains, twigs as in Section \\ref{sec:log_surfaces}, a vertex being \\enquote{rational} if $g=r=0$.\n\nA connected, rational graph $\\Gamma$ is \\emph{normal} if the following three conditions hold. First, all non-branching vertices of $\\Gamma$ have weights at most $-2$. Second, if a vertex with $\\beta=3$ meets two twigs of type $[2]$ then $\\Gamma$ is a fork. Third, if $\\Gamma=(\\!( (2)_{k} )\\!)$ for some $k> 0$ then at least two edges are labeled with \\enquote{$-$}. To extend this definition to non-rational graphs, one needs to exclude some additional special cases, see \\cite[\\S 4]{Neumann-plumbing_graphs}. We will not need this extension. In general, a graph is \\emph{normal} if all its connected components are.\n\nFor a graph $3$-manifold $N$, \\cite[Theorem 4.1]{Neumann-plumbing_graphs} gives an algorithm to reduce its graph to a normal form, which we denote by $\\Gamma(N)$. This algorithm uses certain operations R0--R8 defined in Proposition 2.1 loc.\\ cit, which do not change the graph manifold. Among these, R1 is blowing down a vertex with weight $\\pm 1$, and R3 is an \\enquote{absorption} of a cylinder $(\\mathbb{D}^{1}\\times \\mathbb{S}^{1})\\times \\mathbb{S}^{1}$ corresponding to a rational vertex of weight $0$ and $\\beta=2$. The graph $\\Gamma(-N)$ can be directly computed from $\\Gamma(N)$ using \\cite[Theorem 7.1]{Neumann-plumbing_graphs}.\n\n\nNow we can state the main result of \\cite{Neumann-plumbing_graphs} used to distinguish $\\pi_{1}^{\\infty}(S_{p_1,p_2})$ for different degrees of $p_{j}$.\n\n\\begin{lem}[{\\cite[Theorems 4.2, 4.3]{Neumann-plumbing_graphs}}]\\label{lem:graphs}\n\tLet $N,N'$ be graph $3$-manifolds.\n\t\\begin{enumerate}\n\t\t\\item\\label{item:Gamma} $\\Gamma(N)=\\Gamma(N')$ if and only if $N$ and $N'$ are orientation-preserving diffeomorphic.\n\t\t\\item\\label{item:irr} If $\\Gamma$ is connected then $N$ is prime.\n\t\\end{enumerate}\n\\end{lem}\n\n\n\t \t \\begin{figure}[ht]\t\n\t\\begin{tabular}{cc}\n\t\t\\begin{subfigure}[t]{0.45\\textwidth}\n\t\t\t\\centering\t \t \t\t\t\n\t\t\t\\includegraphics[scale=0.3]{figures\/g_1}\n\t\t\t\\caption{Graph of $M$ as in \\cite{Mumford-surface_singularities}, see Figure \\ref{fig:Skl}}\n\t\t\t\\label{fig:g_1}\n\t\t\\end{subfigure}\t\n\t\t&\n\t\t\\begin{subfigure}[t]{0.45\\textwidth}\n\t\t\t\\centering\t \t \t\t\t\n\t\t\t\\includegraphics[scale=0.3]{figures\/g_2}\n\t\t\t\\caption{$\\Gamma(M)$ for $p_{1},p_{2}\\neq 1$}\n\t\t\t\\label{fig:g_2}\n\t\t\\end{subfigure}\n\t\t\\\\\n\t\t\\begin{subfigure}[t]{0.45\\textwidth}\n\t\t\t\\centering\t \t \t\t\t\n\t\t\t\\includegraphics[scale=0.35]{figures\/g_3}\n\t\t\t\\caption{$\\Gamma(M)$ for $p_{1}=1$, $p_{2}\\neq 1$}\n\t\t\t\\label{fig:g_3}\n\t\t\\end{subfigure}\t\n\t\t&\n\t\t\\begin{subfigure}[t]{0.45\\textwidth}\n\t\t\t\\centering\n\t\t\t\\includegraphics[scale=0.3]{figures\/g_4}\n\t\t\t\\caption{$\\Gamma(-M)$ for $p_{1},p_{2}=1$, see Proposition \\ref{prop:JSJ}\\ref{item:Moser}}\n\t\t\t\\label{fig:g_4}\n\t\t\\end{subfigure}\n\t\\end{tabular}\n\t\\caption{Graphs of $M=\\d\\mathrm{Tub}(D)$ in the proof of Corollary \\ref{cor:not_homeo}}\n\t\\label{fig:graphs}\n\\end{figure}\n\n\\begin{cor}\\label{cor:not_homeo}\n\tThe surfaces $S_{p_{1},p_{2}}$ and $S_{p_{1}',p_{2}'}$ are not homeomorphic for $\\{\\deg p_{1},\\deg p_{2} \\}\\neq \\{\\deg p_{1}',\\deg p_{2}' \\}$.\n\\end{cor}\n\\begin{proof}\n\tFor $S_{p_{1},p_{2}}$ and $S_{p_{1}',p_{2}'}$, $\\{\\deg p_{1},\\deg p_{2}\\}\\neq\\{\\deg p_{1}',\\deg p_{2}'\\}$ let $M$, $M'$ be a boundary of tubular neighborhoods of respective $D\\subseteq X$ as in Construction \\ref{constr}. Such $M$ is a graph manifold represented by Figure \\ref{fig:g_1}. Absorption (move R3 in \\cite{Neumann-plumbing_graphs}) of a vertex corresponding to $L_{1,\\infty}$ reduces it to a graph in Figure \\ref{fig:g_2}. It is in normal form if $p_{1},p_{2}\\neq 1$. If, say, $p_{1}=1$ then the normal form is obtained by blowing down (move R1) the vertex corresponding to $L_{1,0}$. If $p_{2}\\neq 1$, then the resulting graph in Figure \\ref{fig:g_3} is normal, otherwise it needs to be replaced by a standard graph of a Seifert fiber space (move R7). In our case this graph is a graph $E_{9}$, which after orientation reversing gives the one in Figure \\ref{fig:g_4}, cf.\\ Proposition \\ref{prop:JSJ}\\ref{item:Moser} and \\cite[p.\\ 309]{Neumann-plumbing_graphs}. \n\t \t \t\n\tHence $\\Gamma(M)\\neq \\Gamma(M')$, so by Lemma \\ref{lem:graphs}\\ref{item:Gamma} $M$ is not orientation-preserving diffeomorphic to $M'$. Reversing orientation in Figures \\ref{fig:graphs}\\subref{fig:g_2}-\\subref{fig:g_3} inverts the edge signs and replaces a twig $[(2)_{d}]$ with $[d+2]$, so the resulting graph is different from the previous ones. Therefore, $M\\not\\cong_{\\mathrm{diffeo}}M'$, so $M\\not\\cong_{\\mathrm{homeo}}M'$, see \\cite[1.1]{AFW_3M}. Because $\\Gamma(M)$ is connected and not a rational chain, $M$ is prime, and not a lens space (see \\cite[6.1]{Neumann-plumbing_graphs} for lens space graphs), so by \\cite[2.1.2]{AFW_3M} $\\pi_{1}(M)\\not\\cong \\pi_{1}(M')$. The result follows because $\\pi_{1}^{\\infty}(S_{p_{1},p_{2}})=\\pi_{1}(M)$ \\cite{Mumford-surface_singularities}.\n\\end{proof}\n\n\n\n\\@startsection{subsection}{3{Almost minimal models}\\label{sec:peeling}\n\nWe now recall the construction of an almost minimal model for a log smooth pair $(X,D)$. We consider only the case when $X\\setminus D$ is affine, for a complete treatment we refer to \\cite[\\S 3]{Miyan-OpenSurf}, see p.\\ 107 loc.\\ cit.\\ for a summary. The exposition here is based on \\cite[\\S 3]{Palka-minimal_models}.\n\nThe motivation for this construction to get an explicit description of a log MMP run for the pair $(X,D)$, such that the log terminal singularities on the image of $(X,D)$ are introduced as late as possible: in fact, \\emph{an almost minimal model} is a log resolution of the minimal model in the sense of Mori. Having understood the latter, one can use this theory to reconstruct the initial pair $(X,D)$. \n\nThe almost minimal model will be created by iterating birational morphisms of the following type\n\n\\begin{dfn}[{half-point attachment, cf.\\ \\cite{Fujita-noncomplete_surfaces}}]\\label{def:hp}\nLet $(X,D)$ be a log smooth pair. Assume that\n\\begin{equation}\\label{eq:hp}\nA\\subseteq X\\mbox{ is a }(-1)\\mbox{-curve such that } A\\not\\subseteq D\\mbox{ and } A\\cdot D=1.\n\\end{equation}\nThe composition of the contraction of $A$ and new superfluous $(-1)$-curves in the subsequent images of $D$:\n\\begin{equation*}\n\\phi_{A}\\colon (X,D)\\to (X',D'),\n\\end{equation*} \nwhere $D'=(\\phi_{A})_{*}D$, is called a \\emph{half-point attachment} (of $A$).\n\\end{dfn}\n\\begin{rem*}\n\tIn general, $\\phi_{A}$ is \\emph{not} uniquely determined by $A$. Indeed, it may happen that after some contraction within $\\phi_{A}$, there appear two superfluous $(-1)$-curves. For example, if $A$ meets the middle component $T'$ of a subchain $T=[2,2,2]$ of $D$, whose components are non-branching in $D$, then after the contraction of $A+T'$, both components of the image of $T$ are superfluous $(-1)$-curves in the image of $D$, but only one of them can be contracted.\n\t\n\tHowever, such ambiguity never happens if $\\mathbb{C}[X\\setminus D]$ is a UFD. Indeed, Lemma \\ref{lem:producing_Cst}\\ref{item:HP_C} shows that in this case $\\phi_{A}$ contracts exactly $A$ and a maximal $(-2)$-twig of $D$ meeting $A$ (if such occurs).\n\\end{rem*}\n\n\n\\begin{lem}\\label{lem:HPD_k}\n\tLet $(X,D)\\to (X',D')$ be a half-point attachment. Then for all $m\\geq 0$ we have $h^{0}(m(K_{X'}+D'))=h^{0}(m(K_{X}+D))$. In particular, $\\kappa(X'\\setminus D')=\\kappa(X\\setminus D)$.\n\\end{lem}\n\\begin{proof}\n\tLet $A$ be as in \\eqref{eq:hp}. We have $A\\cdot(K_{X}+D+A)=-1<0$, so for any $m>0$, $mA$ is in the fixed locus of $|m(K_{X}+D+A)|$. Therefore, $|m(K_{X}+D)| \\cong |m(K_{X}+D+A)| \\cong |m(K_{X'}+D')|$ because $X\\to X'$ contracts only components of $D+A$.\n\\end{proof}\n\n\\begin{cor}\\label{cor:S_kappa}\n\tThe surfaces $S_{p_1,p_2}$ from Theorem \\ref{THM} are in $\\mathcal{S}_{0}$.\n\\end{cor}\n\\begin{proof}\n\tBy Corollary \\ref{cor:S_k}, $S_{p_1,p_2}\\in \\mathcal{S}_{\\kappa}$ for some $\\kappa$. Construction \\ref{constr} provides a log smooth completion of $S_{p_1,p_2}$ obtained from a log smooth completion of $\\mathbb{C}^{*}\\times \\mathbb{C}^{*}$ by two half-point attachments. Hence by Lemma \\ref{lem:HPD_k}, $\\kappa=\\kappa(S_{p_1,p_2})=\\kappa(\\mathbb{C}^{*}\\times \\mathbb{C}^{*})=0$.\n\\end{proof}\n\nWe now state the main result of the theory of peeling. Recall that $D^{\\#}=D-\\operatorname{Bk} D$, and $\\operatorname{Bk} D$ is a $\\mathbb{Q}$-effective divisor with coefficients in $(0,1)$, supported on the twigs of $D$.\n\n\\begin{prop}[Construction of an almost minimal model]\\label{prop:MMP}\n\tLet $(X,D)$ be a minimal log smooth completion of a smooth affine surface. Then there is a birational morphism\n\t\t\\begin{equation}\\label{eq:MMP}\n\t\t\\psi\\colon (X,D)\\toin{\\psi_{1}}(X_{1},D_{1})\\toin{\\psi_{2}}\\dots \\toin{\\psi_{n}} (X_{n},D_{n})\n\t\t\\end{equation}\n\tsuch that, for all $i\\in \\{1,\\dots, n\\}$:\n\t\\begin{enumerate}\n\t\t\\item\\label{item:psi_hpd} $\\psi_{i}$ is a half-point attachment.\n\t\t\\item\\label{item:X_i-smooth} $D_{i}=(\\psi_{i})_{*}D_{i-1}$ is snc-minimal, $(X_{i},D_{i})$ is log smooth and $X_{i}\\setminus D_{i}$ is an affine open subset of $X\\setminus D$.\n\t\t\\item\\label{item:psi_kappa} $h^{0}(m(K_{X_{i}}+D_{i}))=h^{0}(m(K_{X}+D))$ for all $m\\geq 0$. In particular, $\\kappa(X_{i}\\setminus D_{i})=\\kappa(X\\setminus D)$.\n\t\t\\item\\label{item:psi_min} If $\\kappa(X\\setminus D)\\geq 0$ then $K_{X_{n}}+D_{n}^{\\#}$ is nef. Otherwise, either $X_{n}$ admits a $\\P^{1}$-fibration with $(K_{X_{n}}+D_n^{\\#})\\cdot F<0$ for a fiber $F$; or $-(K_{X_{n}}+D_{n})$ becomes ample after the contraction of $\\operatorname{Bk} D_n$.\n\t\\end{enumerate}\n\\end{prop}\n\\begin{proof}[Sketch of a proof]\n\tLet $\\alpha\\colon (X,D)\\to (Y,D_{Y})$ be the contraction of $\\operatorname{Supp}\\operatorname{Bk} D$, i.e.\\ of all admissible twigs of $T$. Then $\\alpha^{*}(K_{Y}+D_{Y})=K_{X}+D^{\\#}$. Indeed, for every component $T_{0}$ of $\\operatorname{Bk} D$, we have $T_{0}\\cdot (K_{X}+D-\\alpha^{*}(K_{Y}+D_{Y}))=T_{0}\\cdot(K_{X}+T_{0})+\\beta_{D}(T_{0})=-2+\\beta_{D}(T_{0})=T_{0}\\cdot \\operatorname{Bk} D$ by definition of bark, so $K_{X}+D-\\alpha^{*}(K_{Y}+D_{Y})=\\operatorname{Bk} D$ because $\\operatorname{Bk} D$ is negative definite.\n\t\n\tIn particular, $(Y,D_{Y})$ is log terminal. General theorems of Mori theory imply that either $(Y,D_{Y})$ is minimal, so \\ref{item:psi_min} holds for $n=0$, or there is a log extremal curve $\\ll\\subseteq Y$ such that $\\ll^{2}<0$ and $\\ll\\cdot (K_{Y}+D_{Y})<0$. Put $A=\\alpha^{-1}_{*}\\ll$. We claim that $A$ satisfies \\eqref{eq:hp}.\n\t\n\tWe have $0>\\ll\\cdot (K_{Y}+D_{Y})=A \\cdot(K_{X}+D^{\\#})$. Suppose $A\\subseteq D$. Then $0>-2+\\beta_{D}(A)-A\\cdot \\operatorname{Bk} D$, hence $A$ is a tip of $D-\\operatorname{Supp}\\operatorname{Bk} D$ and meets at most one component of $\\operatorname{Bk} D$. This means that $A$ is a component of an admissible twig of $D$, a contradiction with the definition of $\\operatorname{Bk} D$.\n\t\n\tTherefore, $A\\not \\subseteq D$, so $A\\cdot D^{\\#}>0$. Thus $A\\cdot K_{X}<0$, which together with $A^{2}<0$ implies that $A$ is a $(-1)$-curve. Moreover, $A\\cdot D^{\\#}<-A\\cdot K_{X}=1$, so $A$ meets $D$ only on $\\operatorname{Bk} D$. Now a computation \\cite[II.3.7.1(2)]{Miyan-OpenSurf}, which is rather complicated, but uses nothing more than negative definiteness of $A+\\operatorname{Bk} D$, shows that $A$ meets each connected component of $D$ at most once. Because $X\\setminus D$ is affine, it follows that $A\\cdot D=1$, so $A$ satisfies \\eqref{eq:hp}, as claimed.\n\t\n\tWe now put $\\psi_{1}=\\phi_{A}\\colon (X,D)\\to(X_{1},D_{1})$, for some choice of $\\phi_{A}$ such that $D_{1}$ is snc-minimal. Then for $i=1$, \\ref{item:psi_hpd},\\ref{item:X_i-smooth} hold by definition and \\ref{item:psi_kappa} holds by Lemma \\ref{lem:HPD_k}. Next, we replace $(X,D)$ by $(X_{1},D_{1})$ and repeat the procedure, which eventually ends since $\\rho(X)$ drops after each step.\n\\end{proof}\n\n\n\\section{Proof of Theorem \\ref{THM}}\\label{sec:proof}\n\\setcounter{claim}{0}\n\nRecall that $S_{p_1,p_2}\\in \\mathcal{S}_{0}$ by Corollary \\ref{cor:S_kappa}, and $S_{p_1,p_2}\\not\\cong S_{p_1',p_2'}$ for $\\{p_1,p_2\\}\\neq \\{p_1',p_2'\\}$ by Corollary \\ref{cor:st}\\ref{item:distinct}. It remains to show that any $S\\in \\mathcal{S}_{0}$ is isomorphic to some $S_{p_1,p_2}$. Let $(X,D)$ be a minimal log smooth completion of $S$. The idea of the proof is to view $\\phi$ from Construction \\ref{constr} as some almost minimalization \\eqref{eq:MMP} of $(X,D)$. By Proposition \\ref{prop:MMP} each intermediate pair $(X_{i},D_{i})$ in \\eqref{eq:MMP} is a log smooth completion of an affine open subset of $X\\setminus D$, of the same Kodaira dimension. Hence under the assumptions of Theorem \\ref{THM}, $\\kappa(X_{i}\\setminus D_{i})=0$ and $\\mathbb{C}[X_{i}\\setminus D_{i}]$ is a UFD by Remark \\ref{rem:open_UFD}. \n\nTo reconstruct $(X,D)$ we need to understand each half-point attachment $\\psi_{i+1}\\colon (X_{i},D_{i})\\to(X_{i+1},D_{i+1})$. This is done in Lemma \\ref{lem:producing_Cst}, which is a version of \\cite[Lemma III.4.4.3]{Miyan-OpenSurf}, with almost the same proof.\n\n\\begin{lem}\\label{lem:producing_Cst}\n\tLet $S$ be a smooth affine surface. Assume that $\\kappa(S)=0$ and $\\mathbb{C}[S]$ is UFD. Let $(X,D)$ be a minimal log smooth completion of $S$ and let $\\phi_{A}$ be a half-point attachment of some $A\\subseteq X$. Then\n\t\\begin{enumerate}\n\t\t\\item\\label{item:HP_Cst} $A\\cap S\\cong \\mathbb{C}^{1}$ is a smooth fiber of some $\\mathbb{C}^{*}$-fibration of $S$\n\t\t\\item\\label{item:HP_C} $\\operatorname{Exc} \\phi_{A}=[1,(2)_{k}]$, for some $k\\geq 0$, is a twig of $D+A$. Its image does not lie on a rational twig of $(\\phi_A)_{*}D$.\n\t\\end{enumerate}\n\t\\begin{figure}[htbp]\n\t\t\\begin{tabular}{ccc}\n\t\t\t\\begin{subfigure}[t]{0.4\\textwidth}\n\t\t\t\t\\includegraphics[scale=0.34]{figures\/HP}\n\t\t\t\\end{subfigure}\n\t\t\t&\n\t\t\t\\raisebox{1cm}{$\\xrightarrow{\\quad \\phi_{A}\\quad}$}\n\t\t\t&\n\t\t\t\\begin{subfigure}[t]{0.2\\textwidth}\n\t\t\t\t\\includegraphics[scale=0.34]{figures\/HP_1}\n\t\t\t\\end{subfigure}\n\t\t\\end{tabular}\n\t\t\\caption{A half-point attachment to $S$ if $\\mathbb{C}[S]$ is a UFD and $\\kappa(S)=0$, see Lemma \\ref{lem:producing_Cst}.}\n\t\t\\label{fig:HP}\n\t\\end{figure}\n\\end{lem}\n\\begin{proof}\n\\ref{item:HP_Cst} Because $\\mathbb{C}[S]$ is a UFD, $A|_{S}$ is a divisor of zeros of some $f\\in \\mathbb{C}[S]$. Let $F_{\\mathrm{aff}}$ be some general fiber of $f$. It is reduced and irreducible because $A|_{S}$ is. Put $S'=S\\setminus A$, $U'=f(S')\\subseteq \\mathbb{C}^{*}$. By Lemma \\ref{lem:HPD_k} $\\kappa(S')=\\kappa(S)=0$, so the Kawamata addition theorem \\cite[Theorem 11.15]{Iitaka_AG} gives $0=\\kappa(S')\\geq \\kappa(F_{\\mathrm{aff}})+\\kappa(U')\\geq \\kappa(F_{\\mathrm{aff}})+\\kappa(\\mathbb{C}^{*})=\\kappa(F_{\\mathrm{aff}})\\geq 0$, where the last inequality follows from the Iitaka Easy Addition theorem. Hence $\\kappa(F_{\\mathrm{aff}})=0$, which implies that $\\kappa(F_{\\mathrm{aff}})\\cong \\mathbb{C}^{*}$ because $F_{\\mathrm{aff}}$ is smooth and affine. Thus $f$ is a $\\mathbb{C}^{*}$-fibration of $S$, with $A|_{S}$ as one of its fibers.\n\n\\ref{item:HP_C} Let $T$ be a maximal admissible twig of $D$ meeting $A$ in its first tip; put $T=0$ if there is no such. Since $S$ is affine, $D\\neq T$. Let $C$ be the component of $D-T$ meeting $A+T$.\nLet $\\tau\\colon \\hat{X}\\to X$ be a minimal sequence of blowups over $D$ such that the $\\mathbb{C}^{*}$-fibration from \\ref{item:HP_Cst} extends to a $\\P^{1}$-fibration of $\\hat{X}$; let $F$ be a fiber containing $\\hat{A}:= \\tau^{-1}_{*}A$. Put $\\hat{D}=(\\tau^{*}D)_{\\mathrm{red}}$. Note that $\\tau$ does not touch $\\hat{A}$: indeed, otherwise the only base point of $\\tau^{-1}$ is the unique common point of $A$ and $D$, so $F\\subseteq \\tau^{*}A$ is negative definite, which is false.\n\nIf a $(-1)$-curve $L\\subseteq \\hat{D}\\vert$ is superfluous in $\\hat{D}$, then $L$ is not a component of a rational twig of $\\hat{D}$. Indeed, $L\\not\\subseteq \\operatorname{Exc}\\tau$ since $\\tau$ is minimal, and $\\tau(L)^{2}\\neq -1$ since $D$ is snc-minimal; hence $\\tau(L)^{2}\\geq 0$, and the claim follows from Lemma \\ref{lem:rational-twigs-are-admissible}. Let $\\sigma$ be the contraction of all vertical superfluous $(-1)$-curves in $\\hat{D}$ and its images. By our claim, $\\sigma$ does not touch rational twigs of $\\hat{D}$. Put $D'=\\sigma_{*}\\hat{D}$, $F'=\\sigma_{*}F$, $A':= \\sigma_{*}\\hat{A}$.\n\nIf $\\sigma$ touches $\\hat{A}$ then $T=0$ and $\\tau^{-1}_{*}C\\subseteq \\operatorname{Exc}\\sigma$ does not lie in a rational twig of $\\hat D$, so the result follows. Assume $\\sigma$ does not touch $\\hat A$, so $A'$ is a $(-1)$-curve of multiplicity one in a fiber $F'$. Now, $F'-A'$ contains another $(-1)$-curve, say $C'$. By definition of $\\sigma$, $\\beta_{D'}(C')\\geq 3$. Since $\\beta_{F'_{\\mathrm{red}}}(C')\\leq 2$, $C'$ meets $D'_{\\mathrm{hor}}$. By Lemma \\ref{lem:untwisted}, $D'_{\\mathrm{hor}}$ consists of two $1$-sections, so $C'$ has multiplicity one in $F'$. Hence $\\beta_{F'_{\\mathrm{red}}}(C')=1$, so $C'\\cdot D'_{\\mathrm{hor}}=2$, which implies that $(F'-C')\\cdot D'_{\\mathrm{hor}}=0$, and that $C'$ is a unique $(-1)$-curve in $F'-A'$. Therefore, $F'=[1,(2)_{k},1]$ for some $k\\geq 0$. In other words, $T':= F'-A'-C'=[(2)_{k}]$ is zero or a maximal twig of $D'$. Applying Lemma \\ref{lem:rational-twigs-are-admissible} to image of $D'$ after contraction of $A'+T'$, we infer that $C'$ does not lie in a rational twig of $D'-T'$. \n\nSince $\\sigma$ does not touch rational twigs of $D'$, $\\hat{T}:= \\sigma^{-1}_{*}T'$ is zero or a maximal twig of $\\hat{D}$; and the component $\\hat{C}:= \\sigma^{-1}_{*}C'$ meeting it does not lie in a rational twig of $\\hat{D}-\\hat{T}$. Thus $\\tau_{*}\\hat{T}=T$, $\\tau_{*}\\hat{C}=C$: indeed, otherwise $\\tau$ touches $\\hat{C}$, so $C^{2}\\geq 0$, and $C$ lies in an admissible twig of $D$, which is impossible by Lemma \\ref{lem:rational-twigs-are-admissible}. Therefore, $T=[(2)_{k}]$, and $C$ does not lie in any rational twig of $D$, as claimed.\n\\end{proof}\n\n\nOur next lemma is a standard result concerning curves of canonical type, see eg.\\ \\cite[I.3.3.1]{Miyan-OpenSurf}. It will give a description of the almost minimal model of $(X,D)$, or, more precisely, the pair obtained from $(X,D)$ after all possible half point attachments, called a \\emph{strongly minimal model} in \\cite[II.4.9]{Miyan-OpenSurf}.\n\nRecall that $\\mathbb{F}_{m}$ for an integer $m\\geq 0$ denotes the $m$-th Hirzebruch surface $\\P(\\O_{\\P^{1}}\\oplus \\O_{\\P^{1}}(m))$.\n\n\n\\begin{lem}[strongly minimal models]\\label{lem:O}\n\tLet $S$ be a smooth rational affine surface such that $\\mathbb{C}[S]$ is a UFD. Let $(X,D)$ be a minimal log smooth completion of $S$. Assume that $K_{X}+D\\equiv 0$ and that all $(-1)$-curves on $X$ are contained in $D$. Then one of the following holds:\n\t\t\\begin{enumerate}\n\t\t\t\\item\\label{item:Fm_4} $X=\\mathbb{F}_{m}$ for some $m\\neq 1$ and $D=\\lc0,m,0,-m)\\!)$,\n\t\t\t\\item\\label{item:Fm_3} $X=\\mathbb{F}_{m}$ for some $m\\neq 1$ and $D=\\lc0,m,-m-2 )\\!)$,\n\t\t\t\\item\\label{item:triangle} $X=\\P^{2}$ and $D=(\\!( -1,-1,-1 )\\!)$, i.e.\\ $D$ is a triangle,\n\t\t\t\\item\\label{item:conic} $X=\\P^{2}$ and $D=(\\!( -1,-4)\\!) $, i.e.\\ $D$ is a sum of a line and a conic which meet at two points.\n\t\t\\end{enumerate}\n\\end{lem}\n\\begin{proof}\n\tLet $\\tau\\colon (X,D)\\to (X',D')$ be a contraction of some $(-1)$-curves in $D$ and its images, such that all singularities of $D'$ are ordinary double points. Since $D$ is snc-minimal, $\\operatorname{Bs}\\tau^{-1}\\subseteq \\operatorname{Sing} D$. Let $A'$ be a $(-1)$-curve in $X'$. Then $A'\\cdot D'=-A'\\cdot K_{X'}=1$. If $A'\\not\\subseteq D'$ then $A'\\cap \\operatorname{Sing} D'=\\emptyset$, so $\\tau$ does not touch $A:=\\tau^{-1}_{*}A$. Therefore, $A\\subseteq X$ is a $(-1)$-curve not contained in $D$, contrary to our assumption. Hence $A'\\subseteq D'$, so $\\beta_{D'}(A')=A'\\cdot D'-A^2=2$. Since no two components of $D'$ are tangent, $\\#A'\\cap (D'-A')=2$. After the contraction of $A'$, the singularities of the image of $D'$ are still at most ordinary double points. \n\t\n\tThus we can assume that $X'$ has no $(-1)$-curves, i.e.\\ $X'=\\P^2$ or $\\mathbb{F}_{m}$ for some $m\\neq 1$. By Lemma \\ref{lem:S0}\\ref{item:UFD}, the components of $D$ generate $\\operatorname{Pic}(X)$, so the components of $D'$ generate $\\operatorname{Pic}(X')$, too. \n\t\n\tAssume first that $X'=\\P^{2}$. Then $D'\\equiv -K_{\\P^{2}} \\equiv 3\\cdot [\\mbox{line}]$. Because the components of $D'$ generate $\\operatorname{Pic}(\\P^{2})$, $D'$ is reducible. In particular, $D'$ has no singular component, so $\\tau=\\mathrm{id}$ and $D$ is as in \\ref{item:triangle} or \\ref{item:conic}.\n\t\n\tAssume now that $X'=\\mathbb{F}_{m}$. Then $2=\\rho(X')\\leq \\#D'$. Because $X'\\setminus D'$ is affine, $D'$ is connected, so $\\beta_{D'}(C)\\geq 1$ for every component $C$ of $D'$. Hence $0=C\\cdot (K_{X'}+D')=2p_{a}(C)-2+\\beta_{D'}(C)\\geq 2p_{a}(C)-1$, It follows that $C$ is smooth, so $\\tau=\\mathrm{id}$. Moreover, $C$ is rational and $\\beta_{D}(C)=2$, so $D$ is circular. \n\t\n\tDenote by $F,T\\in\\operatorname{Pic}(\\mathbb{F}_{m})$ the classes of the fiber of $\\mathbb{F}_{m}$ and of the section with $T^{2}=-m$, which is unique if $m\\neq 0$. We have $F\\cdot D=-F\\cdot K_{\\mathbb{F}_{m}}=2$, so by Lemma \\ref{lem:untwisted} $D_{\\mathrm{hor}}$ consists of two $1$-sections, say $H_{i}\\equiv a_{i}F+T$ for $i\\in\\{1,2\\}$. We can assume $a_1\\geq a_2$. Note that either $a_{i}=0$ or $0\\leq H_{i}\\cdot T=a_{i}-m$, so $a_{i}\\geq m$. Hence $a_1\\geq m$ and either $a_2\\geq m$ or $H_2\\equiv T$. Let $v$ be the number of fibers in $D$. Because all fibers are disjoint and meet $H_{1}$ and $H_{2}$, from the fact that $\\beta_{D}(H_j)=2$, $j\\in \\{1,2\\}$ we infer that \n\t\\begin{equation}\\label{eq:Hirz}\n\t2-v=H_{1}\\cdot H_{2}=a_{1}+a_{2}-m.\n\t\\end{equation}\n\tSuppose $v=0$. Then because $H_{1},H_{2}$ generate $\\operatorname{Pic}(\\mathbb{F}_{m})=\\mathbb{Z}[F+T]$, we have $\\det \\left[ \\begin{smallmatrix} a_{1} & 1 \\\\ a_{2} & 1 \\end{smallmatrix}\\right]=\\pm 1$, so $a_{1}-a_{2}=1$. Now \\eqref{eq:Hirz} gives $a_{2}=\\tfrac{1}{2}(m+1)$, so $2\\nmid m$, hence $m\\geq 3$ and $0\\neq a_2}[d]^{\\sigma'} & (Z,D_{Z}) \\ar@{-->}[d]^{\\sigma} & \\\\\t\n\t\t(X',D') \\ar[r]^{\\phi'} & (Z',D_{Z'}) \\ar[r]^{\\phi''} & (Z'',D_{Z''})\n\t}\n\t\\end{equation*}\n\twhere $D'$, $D_{Z'}$, $D_{Z''}$ are reduced total transforms of $D$, and $(X',D')$ is another log smooth completion of $S$. Eventually, we replace $\\phi$ with $\\phi''\\circ \\phi'$, which is another sequence of half-point attachments. \n\t\n\tConsider case \\ref{item:conic}, i.e.\\ $r=1$, $D_{Z}$ is a sum of a line $L$ and a conic $C$, and $\\#L\\cap C=2$. We have $R_{1}=C-2L$, so by \\eqref{eq:det} $b_{1}\\in C$. Define $\\sigma$ as a blow up at a point of $C\\cap L$ followed by a flow (see Section \\ref{sec:standard}) on the proper transform of $L$. Then $(Z',D_{Z}')$ is as in case \\ref{item:Fm_3}, so we put $\\phi''=\\mathrm{id}$.\n\t\n\tConsider case \\ref{item:triangle} and let $L\\subseteq D_{Z}$ be the line not containing $b_{1},b_{2}$. As before, we define $\\sigma$ as a blowup at $L\\cap (D_{Z}-L)$, followed by a flow on the proper transform of $L$. Then $(Z',D_{Z}')$ is as in case \\ref{item:Fm_4}.\n\t\n\tConsider case \\ref{item:Fm_3}. Write $D_{Z}=F+T+H$, where $F^{2}=0$, $T^{2}=-m$ and $H^{2}=m+2$. We can assume that $b_{1}\\in F$: indeed, otherwise after $\\pm C_{1}^{2}$ flows on $F$ we obtain a pair with that property. Now $r=1$ and $R_{1}=(m+1)F+T-H$, so \\eqref{eq:det} gives $m=0$. Let $\\sigma$ be a flow on $T$ such that $(\\sigma_{*}F)^{2}=1$, followed by a blowup at $\\sigma(F\\cap H)$. Let now $\\phi''$ be the contraction of a proper transform of a member of $|T|$ passing through $F\\cap H$. The resulting pair is as in \\ref{item:Fm_4}.\n\t\n\tEventually, consider case \\ref{item:Fm_4}. Write $D_{Z}=F_{1}+T+F_{2}+H$, where $F_{1}^{2}=F_{2}^{2}=F_{1}\\cdot F_{2}=0$ and as before $T^{2}=-m$, $b_{1},b_{2}\\not\\in T$. We have $r=2$ and $R_{1}=F_{1}-F_{2}$, $R_{2}=mF_{2}+T-H$, so \\eqref{eq:det} implies (after renaming $b_{i}$, $F_{i}$ if necessary) that $b_{1}\\in F_{1}$, $b_{2}\\in H+T$. After $m$ flows on $F_{2}$ we can assume that $m=0$, so $Z\\cong \\P^{1}\\times \\P^{1}$. We can choose coordinates on $Z$ such that $F_{1}=\\ll_{1,0}$, $F_{2}=\\ll_{1,\\infty}$, $b_2\\in H=\\ll_{2,0}$, $T=\\ll_{2,\\infty}$, $b_{1}=(0,1)$, $b_{2}=(1,0)$. Now by Lemma \\ref{lem:producing_Cst}\\ref{item:HP_C}, $\\phi\\colon (X,D)\\to (Z,D_{Z})$ is as in Construction \\ref{constr}.\n\\end{proof}\n\n\n\\section{Diffeomorphism type of $S_{p_{1},p_{2}}$}\\label{sec:diff}\n\nIn this section we prove Theorem \\ref{thm:diff}. To this end, we need to translate Construction \\ref{constr} to the language of Kirby calculus, where blowing up corresponds to attaching handles. We give a quick overview of handlebodies and Kirby diagrams in Section \\ref{sec:Kirby_overview}. In Proposition \\ref{prop:handles} we show that Construction \\ref{constr} amounts to attaching two $2$-handles to $\\mathbb{T}^{2}\\times \\mathbb{D}^{2}$, whose interior is identified with $\\mathbb{C}^{*}\\times \\mathbb{C}^{*}$, see Figure \\ref{fig:handles}. Theorem \\ref{thm:diff} follows by sliding its $2$-handle over the attached ones. The boundary $M$ of the obtained $4$-manifold is a $0$-surgery on the knot $K$ from Figure \\ref{fig:knot}. It is interesting to note that this is one of the few \\emph{exceptional} surgeries on $K$, classified in \\cite{BW_surgery}. More precisely, while all but finitely many surgeries on $K$ are hyperbolic, $M$ contains incompressible tori. Decomposition of $M$ along these tori is described in Proposition \\ref{prop:JSJ}. While interesting in its own right, it gives another way to distinguish $S_{p_1,p_2}$, for different $\\{\\deg p_{1}, \\deg p_{2}\\}$. \n\\smallskip\n\nWe begin with the following observation.\n\\begin{prop}\\label{prop:htp}\n\tFor every $p_{1},p_{2}$, $S_{p_1,p_2}$ is homotopically equivalent to $\\mathbb{S}^{2}$. \n\\end{prop}\n\\begin{proof}\t\n\tViewing $S=S_{p_1,p_2}$ as an iterated affine modification of $\\mathbb{C}^{2}$, we infer from \\cite[Lemma 3.4]{Kal_exotic-measures} that $\\pi_{1}(S)=\\pi_{1}(\\mathbb{C}^{2})=\\{1\\}$. This modification replaces $\\{x_1x_2=0\\}\\subseteq \\mathbb{C}^{2}$, which is contractible, by the affine part of $A_1+A_2$, which is a disjoint union of two copies of $\\mathbb{C}^{1}$. Hence $e_{\\mathrm{top}}(S)=e_{\\mathrm{top}}(\\mathbb{C}^{2})-1+2=2$. Because $S$, being Stein, has a homotopy type of a CW-complex of (real) dimension two \\cite[11.2.6]{GS_Kirby}, it follows that $H_{i}(S,\\mathbb{Z})=\\mathbb{Z}$ for $i=2$ and $0$ for $i>2$. Thus $S$ has weak homotopy type of $\\mathbb{S}^{2}$. Applying the Whitehead theorem to the continuous map $\\mathbb{S}^{2}\\to S$ given by a generator of $\\pi_{2}(S)=H_{2}(S;\\mathbb{Z})=\\mathbb{Z}$, we infer that $\\mathbb{S}^{2}\\simeq_{\\mathrm{htp}} S$, as claimed.\n\\end{proof}\n\n\\@startsection{subsection}{3{Overview of Kirby calculus}\\label{sec:Kirby_overview}\n\nWe now briefly recall the language of Kirby diagrams of $4$-manifolds, for a complete treatment see \\cite[\\S 4-5]{GS_Kirby}. A ($4$-dimensional) $k$-handle $h$ for $k\\in \\{0,\\dots, 4\\}$ is a copy of $\\mathbb{D}^{k}\\times \\mathbb{D}^{4-k}$ attached to a boundary of a $4$-manifold $V$ via an embedding $\\phi\\colon \\d\\mathbb{D}^{k}\\times \\mathbb{D}^{4-k}\\to \\d V$. The images of $\\d\\mathbb{D}^{k}\\times 0$ and $0\\times \\d \\mathbb{D}^{4-k}$ are called the \\emph{attaching} and \\emph{belt} spheres of $h$, the image of $\\mathbb{D}^{k}\\times 0$ is called the \\emph{core} of $h$, see \\cite[Fig.\\ 4.1]{GS_Kirby}. Smoothing corners gives a $4$-manifold $V\\cup_{\\phi}h$ which depends only on the attaching sphere of $h$ and, if $k=2$, on the integer called \\emph{framing} of $h$, which equals the linking number of its attaching circle and its push-off along a transverse vector field \\cite[4.1, 4.5]{GS_Kirby}. For example, blowing up is the same as attaching a $2$-handle with framing $-1$ \\cite[p.\\ 150]{GS_Kirby}.\n\nAny $4$-manifold $V$ can be obtained from $\\mathbb{D}^{4}$ by attaching handles of increasing index. There are two ways of modifying such handle-decomposition without changing the diffeomorphism type of $V$. First, if the attaching sphere of a $k$-handle meets the belt sphere of a $(k-1)$-handle transversally in a single point, these two handles can be \\emph{canceled} \\cite[4.2.9]{GS_Kirby}. Second, a $2$-handle $h$ can be \\emph{slid over} a $2$-handle $h_{0}$ by pushing the attaching circle of $h$ through the belt circle of $h_0$ \\cite[4.2.10]{GS_Kirby}. We use the same letter to denote a handle before and after the slide.\n\nA handle-decomposition of $V$ is encoded by a \\emph{Kirby diagram}, which is a (decorated) link in $\\mathbb{S}^{3}$, i.e.\\ in the boundary of the initial $0$-handle. A $2$-handle is represented by its attaching circle with a framing coefficient. Attaching a $1$-handle to $V$ is the same as drilling a tubular neighborhood of a properly embedded $\\mathbb{D}^{2}\\subseteq V$ \\cite[p.\\ 168]{GS_Kirby}, so it is represented by a circle $\\d\\mathbb{D}^{2}\\subseteq \\d V$, with a dot to distinguish it from a $2$-handle. All dotted circles in a Kirby diagram form an unlink. Once $1$- and $2$-handles are attached, there is usually (e.g.\\ if $V$ is closed or $\\d V$ is connected and $V$ is simply connected, which is our case) a unique way to attach the remaining ones \\cite[p.\\ 148]{GS_Kirby}, so they are not drawn. We use the same letters for handles and corresponding knots in the diagram. \n\nFor example, a diagram consisting of one unknot with coefficient $e$ represents a $\\mathbb{D}^{2}$-bundle over $\\mathbb{S}^{2}$ of Euler number $e$ \\cite[Fig.\\ 4.20]{GS_Kirby}. The diagram in Figure \\ref{fig:handles} without the leftmost and rightmost circle represents $\\mathbb{T}^{2}\\times \\mathbb{D}^2$, see \\cite[Fig.\\ 4.36]{GS_Kirby}. Indeed, attaching the $1$-handles gives $(\\mathbb{T}^{2}\\setminus \\{\\mbox{disk}\\}) \\times \\mathbb{D}^{2}$, whose fundamental group is freely generated by loops based in the initial $0$-handle and going along the cores of the $1$-handles. The $2$-handle, attached along their commutator, caps off the puncture.\n\nHandle cancellation and sliding is represented on the Kirby diagram as follows. First, if a dotted circle meets only one non-dotted circle, once, then the corresponding $1$- and $2$-handle cancel \\cite[Fig.\\ 5.38]{GS_Kirby}. Second, sliding a $2$-handle $t$ over $t_0$ replaces $t$ by its band-sum with a parallel copy $\\tilde{t}_0$ of $t_0$ \\cite[Fig.\\ 5.4]{GS_Kirby}. More precisely, choose an arc in $t$ and $\\tilde{t}_0$ and connect their endpoints such that the obtained \\enquote{square} bounds a band which is disjoint from the rest of the diagram. Recall that $\\ell k(t_0,\\tilde{t}_0)$ is the framing of $t_0$. The new framing of $t$ is the sum of framings of $t$ and $t_0$ plus $2\\ell k(t,t_0)$ \\cite[p.\\ 142]{GS_Kirby}. It is important to remember that the orientations of $t$ and $\\tilde{t}_0$ should match. We denote by $-t$ the knot $t$ taken with opposite orientation. The components of a standard Hopf link drawn as in the left of Figure \\ref{fig:added_torus} are always meant to have counterclockwise orientation and linking number $+1$. \n\n\\@startsection{subsection}{3{Handle decompositions and proof of Theorem \\ref{thm:diff}}\\label{sec:diff_proof}\n\n\\begin{notation}\\label{not:d}\n\tLet $S=S_{p_{1},p_{2}}$ be as in Construction \\ref{constr}. For $j\\in \\{1,2\\}$ put $d_{j}=\\deg p_{j}+1$, $T_{j,d_j}=L_{j,0}$, and denote by $T_{j,i}$, $i\\in \\{1,\\dots, d_{j}-1\\}$ the $i$-th component of the twig $T_{j}$ meeting $A_{j}$, see Figure \\ref{fig:Skl}. Let $\\mathrm{Tub}(D)$ be the tubular neighborhood of $D$ in $X$ constructed in \\cite{Mumford-surface_singularities}. Put $M=\\d \\mathrm{Tub}(D)$ as in Section \\ref{sec:graph_manifolds} and $V=X\\setminus \\mathrm{int}\\, \\mathrm{Tub}(D)$. Clearly, $S\\simeq_{\\mathrm{diffeo}} \\mathrm{int}\\, V$. We may and will assume $d_1\\leq d_2$.\n\\end{notation}\n\n\n\\begin{prop}\\label{prop:handles}\n\tThe $4$-manifold $V$ is obtained from $\\mathbb{T}^{2} \\times \\mathbb{D}^2$ by attaching two $2$-handles with framings $-d_{1}$, $-d_{2}$ along the standard generators of $\\pi_{1}(\\mathbb{T}^{2})$. Its Kirby diagram is given by Figure \\ref{fig:handles}.\n\\end{prop}\n\\begin{proof} \n\tLet $\\phi$ be as in Construction \\ref{constr}. Let $\\mathbb{D}\\subseteq \\mathbb{C}^{2}\\subseteq \\P^1\\times \\P^1$ be a closed disk of radius $2$ centered at $(0,0)$, so that the base points $(0,1)$, $(1,0)$ of $\\phi^{-1}$ belong to $\\mathrm{int}\\, \\mathbb{D}$. Put $X_0:= \\phi^{-1}(\\mathbb{D})\\subseteq X$, $\\phi_{0}:= \\phi|_{X_0}\\colon X_0\\to \\mathbb{D}$ and $\\ll_{i,j}'=\\ll_{i,j}\\cap \\mathbb{D}$. Figure \\ref{fig:link} shows $\\d \\phi_{0}^{-1}(\\ll_{1,0}' \\cup \\ll_{2,0}')$, where the attaching circles of the $2$-handles are drawn together with their framings.\n\t \t \\begin{figure}[ht]\n\t \t \t\\includegraphics[scale=0.35]{figures\/link}\n\t \t \t\\caption{Underlying disks of $L_{1,0}$, $L_{2,0}$ and handles attached to $\\mathbb{C}^{2}$ by $\\phi_{0}$.}\n\t \t \t\\label{fig:link}\n\t \t \\end{figure}\n\t \t \t \n\t To construct $V$, we need to remove from $X_0$ a tubular neighborhood of $\\phi_{0}^{-1}(\\ll_{1,0}' \\cup \\ll_{2,0}')\\setminus (A_{1}\\cup A_{2})$. To do this, we first remove from the $0$-handle all disks whose boundaries are drawn in Figure \\ref{fig:link} in bold, and then we remove the cores of the corresponding $2$-handles (everything suitably thickened).\n\t \n\t By \\cite[p. 214]{GS_Kirby}, the first step gives a Kirby diagram obtained from Figure \\ref{fig:link} as follows. Replace each circle corresponding to $T_{j,i}$ for $j\\in \\{1,2\\}$, $i\\in \\{1,\\dots, d_{j}\\}$ by a dotted circle $t_{j,i}'$, and each twist corresponding to $T_{j,i+1}\\cap T_{j,i}$ (resp.\\ $L_{0,1}\\cap L_{0,2}$) by a $0$-framed $2$-handle $h_{j,i}$ (resp.\\ $h$), linked with the dotted circles as in Figure \\ref{fig:added_torus}. We denote by $a_{j}$ the $2$-handle corresponding to $A_{j}$. \n\t \t\n\t \t\n\t \t \\begin{figure}[ht]\n\t \t \t\\includegraphics[scale=0.35]{figures\/added_torus}\n\t \t \t\\caption{Intersection of two removed disks, see \\cite[Fig.\\ 6.27]{GS_Kirby}.}\n\t \t \t\\label{fig:added_torus}\n\t \t \\end{figure}\n\tBy \\cite[p.\\ 224]{GS_Kirby}, the next step amounts to attaching a $2$-handle along a parallel copy of each removed circle, together with a certain $3$-handle. In our case, for each $T_{j,i}$, $j\\in \\{1,2\\}$, $i\\in \\{1,\\dots, d_{j}-1\\}$ we attach a $(-2)$-framed $2$-handle $t_{j,i}$ such that $\\ell k(t_{j,i},t_{j,i}')=-2$ and $\\ell k(t_{j,i},t_{j,i+1}')=\\ell k(t_{j,i},t_{j,i-1}')=1$, where we put $t_{j,0}'=a_{j}$. The resulting Kirby diagram is shown in Figure \\ref{fig:V} (where we only draw the part corresponding to $L_{1,0}+L_{2,0}+T_{2}+A_{2}$ and skip the subscript \\enquote{$2,$}: the other part is analogous). \n\\begin{figure}[ht]\n\t\\includegraphics[scale=0.32]{figures\/V_half}\n\t\\caption{Right half of the Kirby diagram for $V=X\\setminus \\mathrm{Tub}(D)$.}\n\t\\label{fig:V}\n\\end{figure}\n\t\n\tWe will simplify this diagram by handle slides. First, we slide $t_{j,1}$ over $a_{j}$, see Figure \\ref{fig:first_slide}. After this slide, $t_{j,1}$ unlinks from $a$, meets $t_{j,1}'$ once with $\\ell k(t_{j,1},t_{j,1}')=-1$ and changes framing to $-1$, while the rest of the diagram does not change. \n\t\\begin{figure}[htbp]\n\t\t\\begin{tabular}{ccccc}\n\t\t\t\\begin{subfigure}[t]{0.22\\textwidth}\n\t\t\t\t\\includegraphics[scale=0.3]{figures\/ta_1}\n\t\t\t\\end{subfigure}\n\t\t\t&\n\t\t\t\\raisebox{1.2cm}{\\scalebox{1.5}{$\\rightsquigarrow$}}\n\t\t\t&\n\t\t\t\\begin{subfigure}[t]{0.22\\textwidth}\n\t\t\t\t\\includegraphics[scale=0.3]{figures\/ta_2}\n\t\t\t\\end{subfigure}\n\t\t\t&\n\t\t\t\\raisebox{1.2cm}{\\scalebox{1.5}{$\\sim$}}\n\t\t\t&\n\t\t\t\\begin{subfigure}[t]{0.22\\textwidth}\n\t\t\t\t\\includegraphics[scale=0.3]{figures\/ta_3}\n\t\t\t\\end{subfigure}\n\t\t\\end{tabular}\n\t\t\\caption{Sliding $t_{j,1}$ over $a_j$.}\n\t\t\\label{fig:first_slide}\n\t\\end{figure}\t\t\n\tApplying Lemma \\ref{lem:slides}\\subref{fig:ind_0} below inductively to pairs $(t_0,t)=(t_{j,i},t_{j,i+1})$ for $i=1,\\dots,d_j-1$, we obtain a diagram in Figure \\ref{fig:V_disjoint}. \n\t \t \\begin{figure}[ht]\n\t \t \t\\includegraphics[scale=0.32]{figures\/V_disjoint}\n\t \t \t\\caption{Right half of the Kirby diagram after sliding $t_{i}$ over $t_{i-1}$ for $i=1,2,\\dots, d_{2}-1$}\n\t \t \t\\label{fig:V_disjoint}\n\t \t \\end{figure}\t\n \t \n\tNow by Lemma \\ref{lem:slides}\\subref{fig:sliding_torus} each $h_{i,j}$ can be slid to a $0$-framed unknot disjoint from the rest of the diagram. Since $\\pi_{1}(V)=\\{1\\}$ by Proposition \\ref{prop:htp}, \\cite[p.\\ 148]{GS_Kirby} shows that all $h_{j,i}$'s cancel with all $3$-handles (recall that there are $d_1+d_2-2$ of them, attached while removing the cores of $T_{j,i}$'s). Applying Lemma \\ref{lem:slides}\\subref{fig:a_0} inductively to pairs $(a,t)=(t_{j,i},t_{j,i+1})$, $i=1,\\dots,d_j-1$ gives Figure \\ref{fig:handles}.\n\\end{proof}\n\nIn the above proof, we used the following standard exercise in handle slides, cf.\\ \\cite[p.\\ 142]{GS_Kirby}.\n\n\n\\begin{lem}\\label{lem:slides}\n\tLet $\\ll_{1}$, $\\ll_{2}$ be links in the left and right of one of the Figures \\ref{fig:slides}\\subref{fig:ind_0}-\\subref{fig:a_0}. Define $\\ll\\subseteq \\ll_1$ as $t_0$ in \\subref{fig:ind_0}, $t+h$ in \\subref{fig:sliding_torus} and $t+c_0+a$ in \\subref{fig:a_0}. Let $\\mathscr{k}$ be a Kirby diagram containing $\\ll_{1}$ in such a way that $\\ll$ is disjoint from $\\mathscr{k}-\\ll_{1}$. Then $\\mathscr{k}$ is equivalent to a diagram obtained from $\\mathscr{k}$ by replacing $\\ll_{1}$ with $\\ll_{2}$.\n\\end{lem}\n\t \t \\begin{figure}[ht]\n\t \t \t\\begin{tabular}{cc}\n\t \t \t\t\\begin{subfigure}[t]{0.55\\textwidth}\n\t \t \t\t\t\\includegraphics[scale=0.17]{figures\/ind}\n\t \t \t\t\t\\caption{Sliding $t$ over $t_0$}\n\t \t \t\t\t\\label{fig:ind_0}\n\t \t \t\t\\end{subfigure}\t\n\t \t \t\t&\n\t \t \t\t\\setcounter{subfigure}{2}\n\t \t \t\t\\begin{subfigure}[t]{0.45\\textwidth}\n\t \t \t\t\t\\includegraphics[scale=0.27]{figures\/a}\n\t \t \t\t\t\\caption{Sliding $t$ over $a$ and canceling $(a,c_0)$}\n\t \t \t\t\t\\label{fig:a_0}\n\t \t \t\t\\end{subfigure}\n\t \t \t\t\\\\\n\t \t \t\t\\multicolumn{2}{c}{\n\t \t \t\t\\setcounter{subfigure}{1}\n\t \t \t\t\t\\begin{subfigure}[t]{\\textwidth}\n\t \t \t\t\t\t\\centering\n\t \t \t\t\t\t\\includegraphics[scale=0.35]{figures\/sliding_torus}\n\t \t \t\t\t\t\\caption{After adding and subtracting $t$, a $2$-handle $h$ like in Figure \\ref{fig:added_torus} is ready to be canceled with a $3$-handle}\n\t \t \t\t\t\t\\label{fig:sliding_torus}\n\t \t \t\t\t\\end{subfigure}\n\t \t \t\t}\n\t \t \t\\end{tabular}\n\t \t \t\\caption{Lemma \\ref{lem:slides}: handle slides applied inductively in the proof of Proposition \\ref{prop:handles}}\n\t \t \t\\label{fig:slides}\n\t \t \\end{figure}\n\\begin{proof}\n\t\\subref{fig:ind_0}\n\tIn order to have a clear picture, we draw $t_0$ rotated around a horizontal line, so that both $t_0$ and $t$ go under the bottom part of $c_0$, see Figure \\ref{fig:ind}. We draw a parallel copy of $t_0$ in such a way that it twists over $t_0$ near that crossing, and cut this twist to make a band-sum with $t$. Now the part of $t$ on the right of the vertical diameter of $c$ can be isotoped to the latter, which proves the claim. \n\t\\\n\t\\begin{figure}[ht]\n\t\t\\begin{subfigure}{\\textwidth}\n\t\t\t\\begin{tabular}{ccccc}\n\t\t\t\t\\begin{subfigure}[t]{0.3\\textwidth}\n\t\t\t\t\t\\includegraphics[scale=0.19]{figures\/ind_1}\n\t\t\t\t\\end{subfigure}\n\t\t\t\t&\n\t\t\t\t\\raisebox{1.2cm}{\\scalebox{1.5}{$\\rightsquigarrow$}}\n\t\t\t\t&\n\t\t\t\t\\begin{subfigure}[t]{0.27\\textwidth}\n\t\t\t\t\t\\includegraphics[scale=0.19]{figures\/ind_2}\n\t\t\t\t\\end{subfigure}\n\t\t\t\t&\n\t\t\t\t\\raisebox{1.2cm}{\\scalebox{1.5}{$\\sim$}}\n\t\t\t\t&\n\t\t\t\t\\begin{subfigure}[t]{0.27\\textwidth}\n\t\t\t\t\t\\includegraphics[scale=0.19]{figures\/ind_3}\n\t\t\t\t\\end{subfigure}\n\t\t\t\\end{tabular}\n\t\t\t\\caption{Sliding $t$ over $t_0$}\n\t\t\t\\label{fig:ind}\n\t\t\\end{subfigure}\n\t\\smallskip\n\n\t\\begin{subfigure}{\\textwidth}\n\t\t\\begin{tabular}{ccccc}\n\t\t\t\\begin{subfigure}[t]{0.27\\textwidth}\n\t\t\t\t\\includegraphics[scale=0.35]{figures\/st_1}\n\t\t\t\\end{subfigure}\n\t\t\t&\n\t\t\t\\raisebox{1cm}{\\scalebox{1.5}{$\\rightsquigarrow$}}\n\t\t\t&\n\t\t\t\\begin{subfigure}[t]{0.27\\textwidth}\n\t\t\t\t\\includegraphics[scale=0.35]{figures\/st_2}\n\t\t\t\\end{subfigure}\n\t\t\t&\n\t\t\t\\raisebox{1cm}{\\scalebox{1.5}{$\\sim$}}\n\t\t\t&\n\t\t\t\\raisebox{-.2cm}{\n\t\t\t\t\\begin{subfigure}[t]{0.27\\textwidth}\n\t\t\t\t\t\\includegraphics[scale=0.35]{figures\/st_3}\n\t\t\t\t\\end{subfigure}\n\t\t\t}\n\t\t\\end{tabular}\n\t\t\\caption{After sliding over $t$, $h$ becomes its parallel copy.}\n\t\t\\label{fig:st_proof}\n\t\\end{subfigure}\n\t\\smallskip\n\t\n\t\\begin{subfigure}{\\textwidth}\n\t\t\\begin{tabular}{ccccc}\n\t\t\t\\begin{subfigure}[t]{0.26\\textwidth}\n\t\t\t\t\\includegraphics[scale=0.3]{figures\/a_1}\n\t\t\t\\end{subfigure}\n\t\t\t&\n\t\t\t\\raisebox{1.2cm}{\\scalebox{1.5}{$\\rightsquigarrow$}}\n\t\t\t&\n\t\t\t\\begin{subfigure}[t]{0.28\\textwidth}\n\t\t\t\t\\includegraphics[scale=0.3]{figures\/a_2}\n\t\t\t\\end{subfigure}\n\t\t\t&\n\t\t\t\\raisebox{1.2cm}{\\scalebox{1.5}{$\\rightsquigarrow$}}\n\t\t\t&\n\t\t\t\\begin{subfigure}[t]{0.27\\textwidth}\n\t\t\t\t\\includegraphics[scale=0.3]{figures\/a_3}\n\t\t\t\\end{subfigure}\n\t\t\\end{tabular}\n\t\t\\caption{Canceling the rightmost $1\/2$-handle pair.}\n\t\t\\label{fig:sliding_a}\n\t\\end{subfigure}\n\t\t\\caption{Proof of Lemma \\ref{lem:slides}.}\n\t\t\\label{fig:sliding_lemma}\n\\end{figure}\n\n\t\n\t\\subref{fig:sliding_torus}\t As in \\subref{fig:ind_0}, we draw $t$ rotated so that both $h$ and $t$ go under the bottom part of $c_2$, and do the band-sum there, see Figure \\ref{fig:st_proof}. Then the bottom part of $h$ untangles from $c_1$, $c_2$, and becomes a parallel copy of $t$. Sliding it over $-t$ unlinks it from the rest of the diagram.\n\t\n\t\\subref{fig:a_0} After sliding $t$ over $a$, it becomes disjoint from $c_0$ and meets $c$ as before, see Figure \\ref{fig:sliding_a}, where the box stands for $k$ twists. Moreover, $c_0$ now meets only $a$, once, so these handles cancel.\t\n\\end{proof}\t \t \n\t \t \n\n\\begin{proof}[Proof of Theorem \\ref{thm:diff}]\n\tRecall that by Proposition \\ref{prop:handles}, $S$ is diffeomorphic to an interior of a $4$-manifold with Kirby diagram as in Figure \\ref{fig:handles}, and by Corollary \\ref{cor:not_homeo} such $S$'s are not homeomorphic for different $\\{d_1,d_2\\}$. Therefore, it remains to prove that Kirby diagrams in Figures \\ref{fig:handles} and \\ref{fig:knot} are equivalent.\n\t\n\\begin{figure}[ht]\t\n\\begin{subfigure}{.8\\textwidth}\n\t\\begin{tabular}{ccc}\n\t\t\\begin{subfigure}[t]{0.5\\textwidth}\n\t\t\t\\includegraphics[scale=0.33]{figures\/h_1}\n\t\t\\end{subfigure}\n\t\t&\n\t\t\\raisebox{1.2cm}{\\scalebox{1.5}{$\\sim$}}\n\t\t&\n\t\t\\begin{subfigure}[t]{0.5\\textwidth}\n\t\t\t\\includegraphics[scale=0.33]{figures\/h_12}\n\t\t\\end{subfigure}\n\t\\end{tabular}\n\t\\caption{Adding $a_2-a_1$ to $h$.}\n\t\\label{fig:h}\n\\end{subfigure}\n\\begin{subfigure}{.8\\textwidth}\n\t\\begin{tabular}{ccc}\n\t\t\\begin{subfigure}[t]{0.5\\textwidth}\n\t\t\t\\includegraphics[scale=0.33]{figures\/h_2}\n\t\t\\end{subfigure}\n\t\t&\n\t\t\\raisebox{1.2cm}{\\scalebox{1.5}{$\\sim$}}\n\t\t&\n\t\t\\begin{subfigure}[t]{0.5\\textwidth}\n\t\t\t\\includegraphics[scale=0.33]{figures\/h_3}\n\t\t\\end{subfigure}\n\t\\end{tabular}\n\t\\caption{Subtracting $a_2-a_1$ from $h$ and canceling pairs $(c_j,a_j)$.}\n\t\\label{fig:hn}\n\t\\end{subfigure}\n\\caption{Proof of Theorem \\ref{thm:diff}\n\\label{fig:proof_diff}\n\\end{figure}\t\t\n\n\tLet $h$, $a_{j}$ be the $0$- and $(-d_{j})$-framed $2$-handles in Figure \\ref{fig:handles}, and let $c_{j}$ be the $1$-handle meeting $a_{j}$. We will add and subtract $a_2-a_1$ from $h$ so that the latter untangles from $c_1$, $c_2$. Then we cancel pairs $(c_j,a_j)$. This is done in Figure \\ref{fig:proof_diff}, which we now explain; see also \\cite[5.1.8(b)]{GS_Kirby} for the case $d_1=d_2=1$.\n\t\n\tWe start with Figure \\ref{fig:handles}, but draw it with $a_2$ rotated along a horizontal axis. Next, we draw parallel copies of $-a_{1}$ and $a_{2}$, where a box stands for $d_{j}$ twists. Then we band sum $h$ with those copies near the points where $h$ and $a_{j}$ go under $c_{j}$, see Figure \\ref{fig:h}. This move resolves nearby crossings of $h$ and $c_j$, leaving $\\ell k(h,c_j)=-1$. Now we rotate $a_2$ again (this changes the sign in the box) and repeat the above move, but with reversed orientation of $a_j$, i.e.\\ we add $a_1$ and subtract $a_2$. As before, the crossings of $h$ with $c_j$ get resolved, so these $1$-handles cancel with $a_j$, see Figure \\ref{fig:hn}. After the first slide, $h$ twisted along $a_j$ in the boxes, so now it twists there over itself. A look at Figure \\ref{fig:hn} shows that we do obtain a knot as in Figure \\ref{fig:knot}. The framing of $h$ remains $0$ since we added and subtracted the same thing. \t\n\\end{proof}\n\n\\@startsection{subsection}{3{Computation of $\\pi_{1}^{\\infty}(S_{p_{1},p_{2}})$ and a relation with knot surgeries}\\label{sec:pi_1}\n\nProposition \\ref{prop:handles} allows us to compute the fundamental group of $V$ (which is trivial by Proposition \\ref{prop:htp}), and, more interestingly, that of $M$, which equals $\\pi_{1}^{\\infty}(S)$, see \\cite{Mumford-surface_singularities}.\n\n\\begin{prop}\\label{prop:pi_1}\nThe fundamental group at infinity of $S$ has a presentation\n\\begin{equation}\\label{eq:pi_1}\n\\pi_{1}^{\\infty}(S)=\n\\langle \\delta_{1},\\delta_{2},\\lambda\\ | \\ \\delta_{1}=[\\gamma_2,\\lambda^{-1}],\\ \\delta_{2}=[\\gamma_1,\\lambda],\\ [\\gamma_1,\\gamma_2]=1\\rangle, \\quad \\mbox{ where } \\gamma_{j}=\\delta_{j}^{d_j}.\n\\end{equation}\nMoreover, $\\gamma_j$ for $j\\in \\{1,2\\}$ is the attaching circle of the $(-d_j)$-framed $2$-handle in Figure \\ref{fig:handles}, and $\\lambda$ is the belt sphere of the $0$-framed one.\n\\end{prop}\n\\begin{proof}\n\tLet $\\gamma_{1},\\gamma_{2},\\lambda\\in \\pi_{1}(\\d (\\mathbb{T}^{2}\\times \\mathbb{D}^{2}))=\\pi_{1}(\\mathbb{T}^{3})$ be the standard generators. Proposition \\ref{prop:handles} shows that $M$ is obtained from $\\mathbb{T}^{3}$ by a surgery along $\\gamma_{1},\\gamma_{2}$ with coefficients $d_{1}$, $d_{2}$. The meridians of deleted tori are $\\mu_{1}:=[\\gamma_{2},\\lambda^{-1}], \\mu_{2}:= [\\gamma_{1},\\lambda]$. Now by van Kampen theorem, each surgery replaces the relation $\\mu_j=1$ in $\\pi_{1}(\\mathbb{T}^{3})$ by $\\mu_j=\\delta_j$, where $\\delta_j\\in \\pi_{1}(M)$ is the corresponding attaching circle, satisfying $\\delta_{j}^{d_j}=\\gamma_{j}$.\n\\end{proof}\n\n\\begin{rem}\\label{rem:pi_1-Mumford}\n\tThe presentation \\eqref{eq:pi_1} can also be obtained from the plumbing construction in \\cite{Mumford-surface_singularities}, see Section \\ref{sec:graph_manifolds}. If $D$ is a tree, \\cite{Mumford-surface_singularities} shows how to calculate $\\pi_{1}(M)$ in terms of fibers of the plumbed $\\mathbb{S}^{1}$-bundles, using van Kampen theorem. For our $D$, one needs to use the van Kampen theorem for groupoids \\cite{Brown_van-Kampen}, say, to glue $\\d \\mathrm{Tub}(L_{1,\\infty}+L_{2,\\infty})$ to the rest of $\\d \\mathrm{Tub}(D)$ along two disjoint plumbing tori, each containing a base point. This gives the additional generator $\\lambda$, which is the sum of paths joining the two base points in each of the glued pieces. In this interpretation, $\\gamma_{j}$ is the fiber over $L_{j,0}$. Indeed, in the proof of Proposition \\ref{prop:handles} it was drawn parallel to the boundary of a disk corresponding to $T_{j,d_j-1}$, which goes once around $L_{j,0}$. In turn, $\\delta_{j}$ is a fiber over $T_{j,1}$, that is, a singular fiber of a lens space $L(1,d_{j})$ attached to $\\d \\mathrm{Tub}(L_{j,0})$.\n\\end{rem}\n\nIn the proof of Corollary \\ref{cor:not_homeo}, we have essentially seen the JSJ decomposition of $M$. We will now show how to see it from the above description of $M$, as a $0$-surgery on $K_{[2d_1,2d_2]}$. In order to settle the notation, we recall the basic definitions following \\cite{Hatcher_3M}.\n\nLet $N$ be a $3$-manifold. A surface $\\Sigma$ in $N$ is \\emph{incompressible} if every loop in $\\Sigma$ bounding a disk in $N$ bounds a disk in $\\Sigma$, too. We write $N|\\Sigma$ for a $3$-manifold (with boundary) obtained from $N$ by removing $\\Sigma$ together with some tubular neighborhood. \nAssume that $N$ is irreducible (see Section \\ref{sec:graph_manifolds}). Then $N$ has a collection $\\Sigma$ of incompressible tori such that each component of $N|\\Sigma$ is either a Seifert fiber space or has no incompressible tori. A minimal such $\\Sigma$ (in the sense of inclusion) is unique up to an isotopy \\cite[Theorem 1.9]{Hatcher_3M}, and $N|\\Sigma$ is called a \\emph{JSJ-decomposition} of $N$. \n\nWe now recall the notation for Seifert fibered spaces following \\cite[\\S 2.1]{Hatcher_3M}. Let $B$ be a compact surface of genus $g$ with $r$ boundary components, and let $M_0\\to B_0$ be an $\\mathbb{S}^{1}$-bundle, where $B_0$ is $B$ with $k$ disks $D_{1},\\dots, D_{k}$ removed. The preimage $T_{i}\\subseteq M_{0}$ of $\\d D_{i}$ is a torus: we attach to it a solid torus $\\mathbb{D}^{2}\\times \\mathbb{S}^{1}$ in such a way that a meridian $0\\times \\mathbb{S}^{1}$ is attached to a loop in $T_{i}$ whose preimage in its universal cover is a line of slope $\\tfrac{p_{i}}{q_{i}}$. The resulting manifold is a \\emph{Seifert fibered space} $M(g,r;\\tfrac{p_{1}}{q_{1}},\\dots, \\tfrac{p_{k}}{q_k})$. It is a graph manifold. Its graph has one vertex corresponding to $M_0$ and $k$ maximal twigs, whose type is a continued fraction expansion of $\\tfrac{p_i}{q_i}$. This graph is normal whenever a Seifert fibration is unique, see \\cite[5.7]{Neumann-plumbing_graphs}. Few important exceptions, including lens spaces $\\mathbb{S}^3\/\\mathbb{Z}_{p}$, are listed in \\cite[Theorem 2.3]{Hatcher_3M}. Their normal graphs are described in \\cite[6.1]{Neumann-plumbing_graphs}.\n\nLet again $M=\\d \\mathrm{Tub}(D)$ be as in Notation \\ref{not:d}. The following result can be inferred from the computation of $\\Gamma(M)$ in Figure \\ref{fig:graphs} using the algorithm in \\cite[\\S 5]{Neumann-plumbing_graphs}. Nonetheless, we sketch a direct argument using the geometric description of incompressible tori in $M$ deduced in \\cite[Lemma 2.2]{BW_surgery} from \\cite{HT_inc}.\n\n\n\n\\begin{prop}[JSJ decomposition of $M$]\\label{prop:JSJ}\n\tLet $\\Sigma_{1},\\Sigma_{2}\\subseteq M$ be the plumbing tori at $L_{1,0}\\cap L_{2,0}$ and $L_{1,0}\\cap L_{2,\\infty}$, see Section \\ref{sec:graph_manifolds}. Then the JSJ decomposition of $M$ is\n\t\\begin{enumerate}\n\t\t\\item\\label{item:d_1>1} $M|(\\Sigma_{1}\\sqcup \\Sigma_2)=M(2,0;\\tfrac{1}{d_1})\\sqcup M(2,0;\\tfrac{1}{d_2})$ if $d_{1},d_{2}\\neq 1$,\n\t\t\\item\\label{item:d_1=1} $M|\\Sigma_1=M(2,0;\\tfrac{1}{d_2})$ if $d_{1}=1$ and $d_2>1$,\n\t\t\\item\\label{item:Moser} $M=M(0,0;\\tfrac{1}{2},\\tfrac{1}{3},\\tfrac{1}{6})$ if $d_1=d_2=1$.\n\t\\end{enumerate}\n\tLet $\\gamma_{j},\\delta_{j}$ be as in Proposition \\ref{prop:pi_1}. Then $\\delta_{j}$ is the singular fiber of $M(2,0;\\tfrac{1}{d_j})$ and $\\gamma_{j}$ is a general one.\n\\end{prop}\n\\begin{proof}\n\tBy Theorem \\ref{thm:diff}, $M$ is a $0$-surgery on $\\mathbb{S}^{3}$ along $K=K_{[2d_1,2d_2]}$. Because $K_{[2,2]}$ is just a trefoil, \\ref{item:Moser} follows from \\cite[Proposition 3.1]{Moser_knots} which completely describes surgeries on torus knots. Assume $d_2>1$. If $d_1>1$, too, then cutting along $\\Sigma_1\\sqcup \\Sigma_2$ in case \\ref{item:d_1>1} corresponds to removing the edges of the circular subgraph in Figure \\ref{fig:g_2}, and decorating the leftmost vertices with $r=2$, i.e.\\ puncturing the base spheres twice. If $d_1=1$ then cutting along $\\Sigma_1$ corresponds to removing the loop in Figure \\ref{fig:g_3} and adding boundary as before. This shows the equalities in \\ref{item:d_1>1}--\\ref{item:d_1=1}, and proves that $M$ and, if $d_1=1$, $M|\\Sigma_i$ for $i\\in \\{1,2\\}$ are not Seifert fibered: indeed, $\\Gamma(M|\\Sigma_i)$ is not one of the graphs in \\cite[5.7, 6.1]{Neumann-plumbing_graphs}. It remains to show that $\\Sigma_{1},\\Sigma_{2}\\subseteq M$ are incompressible.\n\t\n\tThe $3$-manifold $M$ is irreducible by \\cite[Theorem 2]{HT_inc} (it also follows from Lemma \\ref{lem:graphs}\\ref{item:irr}). Put $M_0=\\mathbb{S}^{3}\\setminus \\mathrm{Tub}(K)\\subseteq M$. Put $M_0=\\mathbb{S}^{3}\\setminus \\mathrm{Tub}(K)\\subseteq M$. By \\cite[Proposition 1(1) and the proof of Proposition 2]{HT_inc} there are, up to isotopy, at most two incompressible surfaces $\\Sigma_0\\subseteq M_0$ such that the surfaces $\\Sigma\\subseteq M$ obtained by capping $\\Sigma_0$ are tori (see also proof of \\cite[Lemma 2.2]{BW_surgery}). Such $\\Sigma_0$ are denoted by $S_{1}(0)$ and $S_{1}(1)$ in \\cite{HT_inc}, and they have exactly one boundary component each. By \\cite[Proposition 1.5(a)]{Przytycki_inc} they are unknotted (see Definition 1.3 loc. cit), so by Theorem 1.4 loc.\\ cit, $\\Sigma$ remains incompressible. Moreover, by \\cite[Theorem 1(e)]{HT_inc} $S_1(0)$ is isotopic to $S_1(1)$ if and only if $d_1=1$. Thus our claim will follow once we show that $\\Sigma=\\Sigma_1,\\Sigma_2$ for $\\Sigma_0=S_1(0),S_1(1)$. \n\t\n\tTo this end, we recall the description of $S_{1}(0)$, $S_{0}(1)$ from \\cite[p.\\ 227, Fig.\\ 1]{HT_inc}. Consider Figure \\ref{fig:knot} as a knot in $\\mathbb{R}^{3}$, and let $\\mathscr{p}$ be a vertical plane between the \\enquote{twists}. It meets the knot in four points. Connect them to get a quadrilateral $\\mathscr{q}\\subseteq \\mathscr{p}$: it is a $0$-handle for $S_{1}(0)$. Now attach to it two $1$-handles along the bands with $d_{1}$ and $d_{2}$ twists, respectively: this gives a punctured torus $S_{1}(0)$. The torus $S_{1}(1)$ is constructed the same way, with $\\mathscr{q}$ replaced by its complement in the compactification $\\bar{\\mathscr{p}}=\\mathbb{S}^{2}\\times \\{\\mathrm{pt}\\}\\subseteq (\\mathbb{S}^{2}\\times \\mathbb{R})\\sqcup \\{\\pm \\infty\\}=\\mathbb{S}^{3}$.\n\t\n\tThe loops based at $\\mathscr{q}$ and going around the $1$-handles of $S_1(0)$ are isotopic to the non-dotted circles in Figure \\ref{fig:handles}, which by Remark \\ref{rem:pi_1-Mumford} are fibers $\\gamma_j$ over $L_{j,0}$. Hence $S_{1}(0)$ caps to $\\Sigma_1$. To see the $1$-handles of $S_{1}(1)$, we need to change the base point from $\\mathscr{q}$ to its complement. The gluing described in Remark \\ref{rem:pi_1-Mumford} shows that this amounts to changing, say, $L_{2,0}$ to $L_{2,\\infty}$. This way we get $\\Sigma_2$, as claimed.\n\\end{proof}\n\n\\begin{rem}[A direct argument for minimality of $\\Sigma$]In the above proof, we have used the classification of graph manifolds to show that the collections $\\Sigma_1\\sqcup\\Sigma_2$ in \\ref{item:d_1>1} and $\\Sigma_1$ in \\ref{item:d_1=1} of incompressible tori decomposing $M$ to Seifert fibered spaces is minimal. However, this can be also seen directly, as follows.\n\t\n\nFirst, we note that since $\\Gamma(M|\\Sigma_{i})$ is a tree, one can use van Kampen theorem as in \\cite{Mumford-surface_singularities} to compute\n\\begin{equation}\\label{eq:pi_1-Sigma}\n\\pi_{1}(M|\\Sigma_{1})=\\langle \\delta_{1},\\delta_{2}|[\\gamma_1,\\gamma_2]=1 \\rangle, \\mbox{ where } \\delta_j\\mbox{ and } \\gamma_j=\\delta_{j}^{d_{j}}\\mbox{ are as in Remark \\ref{rem:pi_1-Mumford}.}\n\\end{equation}\n\n\nSuppose that $d_1,d_2>1$, but $M|\\Sigma_i$ is Seifert fibered for some $i\\in \\{1,2\\}$. Then by \\cite[Proposition 1.11]{Hatcher_3M}, the incompressible torus $\\Sigma_{3-i}\\subseteq M|\\Sigma_{i}$ is either horizontal or vertical. In the first case, it covers the base orbifold, which has two boundary components, so comparing Euler characteristics as done in \\cite[p.\\ 22]{Hatcher_3M} shows that $M|\\Sigma_i$ has no singular fibers, so $M|\\Sigma_{i}\\cong (\\mathbb{D}^{1}\\times \\mathbb{S}^{1})\\times \\mathbb{S}^{1}$, a contradiction with \\eqref{eq:pi_1-Sigma}. In the second case, cutting $M|\\Sigma_{i}$ further along $\\Sigma_{3-i}$ preserves the Seifert fibration: therefore by the equality in \\ref{item:d_1>1}, it restricts to the unique one on $M(2,0;\\tfrac{1}{d_{j}})$. Hence after gluing $\\Sigma_{3-i}$ back, the fiber of one fibration becomes a loop around the degenerate fiber of the other. This is a contradiction, because the former is central in $\\pi_{1}$, and the latter is not.\n\nSuppose now that $d_{2}>1$, but $M$ is Seifert fibered. Again, the incompressible torus $\\Sigma_{1}$ is either horizontal or vertical. In the first case, since $M|\\Sigma_{1}$ is connected, \\cite[p.\\ 22]{Hatcher_3M} implies that $M|\\Sigma_1\\cong \\Sigma_1\\times \\mathbb{D}^{1}$, a contradiction with \\eqref{eq:pi_1-Sigma}. In the second case, we obtain that $M|\\Sigma_{1}$ is a Seifert fiber space. We have shown that this is possible only if $d_{1}=1$, in which case the general fiber is represented by $\\gamma_{1}\\in \\pi_{1}(M)$. In particular, $\\langle \\gamma_{1}\\rangle \\subseteq \\pi_{1}(M)$ is a (nontrivial) central subgroup. Hence by \\eqref{eq:pi_1} $\\delta_{2}=[\\gamma_{1},\\lambda]=1$ and $\\delta_{1}=[\\delta_{2}^{d_{2}},\\lambda^{-1}]=1$, so $\\gamma_{1}=1$; a contradiction.\n\\end{rem}\n\n\n\n\\section{$\\mathbb{C}^{*}$-fibrations and complete algebraic vector fields on $S_{p_1,p_{2}}$}\\label{sec:AAut}\n\nIn this section we prove Theorem \\ref{thm:AAut}\\ref{item:AAut_thm}. It follows from \\cite[Theorem 1.6]{KKL_avf} that a flow of any complete algebraic vector field on $S_{p_{1},p_{2}}$ preserves some $\\mathbb{C}^{*}$-fibration of $S_{p_{1},p_{2}}$ (see Proposition \\ref{prop:fibrations}\\ref{item:vf} for details). Therefore, to understand $\\mathrm{AAut}_{\\mathrm{hol}}(S_{p_{1},p_{2}})$ we need to classify all $\\mathbb{C}^{*}$-fibrations of $S_{p_{1},p_{2}}$. \n\nTheir degenerate fibers are affine curves with Euler characteristic at least $e_{\\mathrm{top}}(\\mathbb{C}^{*})$, see \\cite[III.1.8]{Miyan-OpenSurf}, i.e.\\ unions of $\\mathbb{C}^{*}$ and $\\mathbb{C}^{1}$. Therefore, we begin by classifying all curves in $S_{p_{1},p_{2}}$ which are isomorphic to $\\mathbb{C}^{1}$. This is done in Lemma \\ref{lem:C1}. In Lemma \\ref{lem:C*_lines} we show that any $\\mathbb{C}^{*}$-fibration of $S_{p_{1},p_{2}}$ comes from a $\\mathbb{C}^{*}$-fibration of a complement of suitably chosen pair of disjoint $\\mathbb{C}^{1}$'s; which by Lemma \\ref{lem:tori} turns out to be a torus $\\mathbb{C}^{*}\\times \\mathbb{C}^{*}$. The union of such tori is exactly the complement of $Z$ from Theorem \\ref{thm:AAut}\\ref{item:AAut_thm}.\n\n\\begin{notation}\\label{not:S}\n\tAs in Notation \\ref{not:d}, we put $d_j=\\deg p_{j}+1$ and write $S:= S_{p_{1},p_{2}}$. We use notation introduced in Construction \\ref{constr} and denote by $T_{j}$ the maximal $(-2)$-twig of $D$ meeting $A_{j}$. Additionally, we denote by $\\mathscr{c}=\\{x_1+x_2=1\\}\\subseteq \\P^{1}\\times \\P^{1}$ the curve of type $(1,1)$ passing through $(\\infty,\\infty)$ and both base points $(1,0)$, $(0,1)$ of $\\phi^{-1}$, and by $C:=\\phi^{-1}_{*}\\mathscr{c}$ its proper transform on $X$. \n\t\n\tIt would be sometimes convenient to work with explicit equations of $S$. To write them, denote by $\\hat{p}_{j}=t^{d_{j}-1}p_{j}(t^{-1})$, $j\\in \\{1,2\\}$ the polynomial obtained from $p_{j}$ by writing its coefficients in the opposite order. Note that $\\hat{p}_{j}(0)=1$ because $p_{j}$ is monic. Now $S\\subseteq \\mathbb{C}^{4}=\\operatorname{Spec}\\mathbb{C}[x_1,x_2,y_1,y_2]$ is given by\n\t\\begin{equation}\\label{eq:S}\n\ty_1x_1^{d_1}=x_2-\\hat{p}_1(x_1),\\quad y_2x_2^{d_2}=x_1-\\hat{p}_2(x_2).\n\t\\end{equation}\n\tNote that $S$ is in fact a surface in $\\mathbb{C}^{3}$ because, say, $x_1$ can be computed from the second equation.\n\\end{notation}\n\n\n\n\n\\begin{lem}[{$\\mathbb{C}^{1}$'s in $S$, cf.\\ \\cite[Lemma 8.1]{Kal-Ku1_AL-theory}}]\\label{lem:C1}\n\tA curve $\\Gamma\\subseteq X$ satisfies $\\Gamma\\cap S\\cong \\mathbb{C}^{1}$ if and only if one of the following holds:\n\t\\begin{enumerate}\n\t\t\\item\\label{item:A} $\\Gamma=A_{j}$ for some $j\\in \\{1,2\\}$. Then $\\Gamma^{2}=-1$ and $\\Gamma\\cdot D=\\Gamma\\cdot T_j=1$.\n\t\t\\item\\label{item:L} $\\Gamma=L_{j,1}$ and $p_{j}=t^{d_{j}-1}$ for some $j\\in \\{1,2\\}$. Then $\\Gamma^{2}=-d_{j}$, $\\Gamma\\cdot D=\\Gamma\\cdot L_{3-j,\\infty}=1$ and $\\Gamma\\cdot A_{3-j}=1$. \n\t\t\\item\\label{item:C} $\\Gamma=C$ and $p_{j}=1$ or $t^{d_{j}-2}(t-1)$ for both $j\\in \\{1,2\\}$. Then $\\Gamma^{2}=2-d_{1}-d_{2}$, $\\Gamma\\cdot D=2$, $\\Gamma$ meets $D$ at $L_{1,\\infty}\\cap L_{2,\\infty}$ and $\\Gamma\\cdot A_{j}=1$ for $j\\in \\{1,2\\}$.\n\t\\end{enumerate}\n\tMoreover, $A_{1}\\cdot A_{2}=0$, $L_{1,1}\\cdot L_{2,1}=1$ and $A_{j}\\cdot L_{j,1}=C\\cdot L_{j,1}=0$ for $j\\in \\{1,2\\}$.\n\\end{lem}\n\\begin{proof}\n\tClearly, $A_{j}\\cap S\\cong \\mathbb{C}^{1}$ and $A_{j}^{2}=-1$ as in \\ref{item:A}. We will now check when $L_{j,1}\\cap S \\cong \\mathbb{C}^{1}$. Because $\\ll_{j,1}$ meets $\\phi_{*}D$ transversally on $\\ll_{3-j,0}$ and $\\ll_{3-j,\\infty}$, we have $L_{j,1}\\cdot(D+A_{3-j})=2$ and $L_{j,1}$ meets $L_{3-j,\\infty}$, so $L_{j,1}\\cap S\\cong \\mathbb{C}^{1}$ if and only if $L_{j,1}$ meets $A_{3-j}$, that is, all $d_{j}$ blowups over $\\ll_{j,1}\\cap \\ll_{3-j,0}$ touch the image of $L_{j,1}$ (hence $L_{j,1}^{2}=\\ll_{j,1}^{2}-d_{j}=-d_{j}$). The proper transform of $\\ll_{j,1}$ after the first blowup meets the exceptional curve at the point with coordinate $(x_{3-j}-1)\/x_{j}=0$ (see Construction \\ref{constr}), so the next blowup, if occurs, touches it if and only if the coefficient near $t^{d_{j}-2}$ in $p_{j}$ is zero. Because these two curves meet transversally, the infinitely near points on the next proper transforms will have coordinate zero, too. Following Construction \\ref{constr} we conclude that $L_{j,1}\\cap S\\cong \\mathbb{C}^{1}$ if and only if $p_{j}=t^{d_{j}-1}$, as in \\ref{item:L}.\n\n\tTo check when $C\\cap S\\cong \\mathbb{C}^{1}$, we argue in a similar way. We have $C\\cdot(D+A_{1}+A_{2})=4$ and $C$ meets $D$ at $L_{1,\\infty}\\cap L_{2,\\infty}$, so $C\\cap S\\cong \\mathbb{C}^{1}$ if and only if all the blowups in the decomposition of $\\phi$ touch the image of $C$. After the first blowup over $\\ll_{j,1}\\cap \\ll_{3-j,0}$ the proper transform of $\\mathscr{c}$ meets the exceptional curve at a point of coordinate $(x_{3-j}-1)\/x_{j}=1$, so all the remaining blowups touch it if and only if $p_{j}=1$ (i.e.\\ there are no more blowups) or $p_{j}=t^{d_{j}-1}-t^{d_{j}-2}$, since as before the next infinitely near points will have coordinate zero on the exceptional curves. In this case, $C^{2}=\\mathscr{c}^{2}-(d_{1}+d_{2})=2-d_{1}-d_{2}$, as in \\ref{item:C}.\n\t\\smallskip\n\t\n\tIt remains to show that if $\\Gamma\\cap S\\cong \\mathbb{C}^{1}$ for some curve $\\Gamma\\subseteq X$ then $\\Gamma$ equals $A_{j}$, $L_{1,j}$ or $C$. Consider first the case when $\\Gamma$ is vertical for the $\\P^{1}$-fibration induced by $|L_{j,\\infty}|$, so $\\Gamma\\subseteq \\phi^{*}\\ll_{j,t}$ for some $t\\in \\mathbb{C}^{1}$ because $\\Gamma\\not\\subseteq D$. If $t\\not\\in\\{0,1\\}$ then $\\Gamma\\cap S\\cong \\mathbb{C}^{*}$, which is false. If $t=0$ then $\\Gamma=A_{j}$ because $\\phi^{*}\\ll_{j,0}=L_{j,0}+T_{j}+A_{j}$. If $t=1$ then $\\Gamma=L_{1,j}$ or $A_{3-j}$ because $(\\phi^{*}\\ll_{j,1})_{\\mathrm{red}}=L_{j,1}+T_{3-j}+A_{3-j}$.\n\t\n\tConsider now the case when $\\Gamma$ is horizontal for both $\\P^{1}$-fibrations induced by $|L_{j,\\infty}|$, $j\\in \\{1,2\\}$. Then $\\Gamma$ meets both $L_{j,\\infty}$, so the point $\\Gamma\\cap D$ equals $L_{1,\\infty}\\cap L_{2,\\infty}$. Since $\\Gamma$ is disjoint from $L_{1,0}+L_{2,0}=[1,1]$, the linear system of the latter induces a $\\P^{1}$-fibration for which $\\Gamma$ is vertical. The curves $L_{j,\\infty}$ are $1$-sections for this $\\P^{1}$-fibration, so $\\Gamma\\cdot L_{j,\\infty}=1$. It follows that $\\phi(\\Gamma)\\subseteq \\P^{1}\\times \\P^{1}$ is of type $(1,1)$, passes through $(\\infty,\\infty)$, and because $\\Gamma\\cdot L_{j,0}=0$, $\\phi(\\Gamma)$ passes through the base point of $\\phi^{-1}$ on $\\ll_{j,0}$, that is, through $(0,1)$ for $j=1$ and $(1,0)$ for $j=2$. Thus $\\phi(\\Gamma)=\\mathscr{c}$, as claimed.\n\\end{proof}\n\n\\begin{lem} \\label{lem:C*_lines}\n\tFor any $\\mathbb{C}^{*}$-fibration of $S$ one can find on $S$ two disjoint, vertical curves isomorphic to $\\mathbb{C}^{1}$, whose closures $\\Gamma_{1},\\Gamma_{2}\\subseteq X$ satisfy $\\Gamma_{i}^{2}\\geq -1$, with strict inequality if $\\Gamma_{i}\\cdot D\\geq 2$.\n\\end{lem}\n\\begin{proof}\n\tFix a $\\mathbb{C}^{*}$-fibration $f$ of $S$. Let $\\alpha_{1}\\colon X''\\to X$ be the minimal resolution of the base points of $f$ on $X$, and let $\\alpha_{2}\\colon X''\\to X'$ be the contraction of all vertical $(-1)$-curves which are superfluous in the subsequent images of $(\\alpha_{1}^{*}D)_{\\mathrm{red}}$. Put $\\alpha=\\alpha_{2}\\circ\\alpha_{1}^{-1}$ and $D'=((\\alpha^{-1})^{*}D)_{\\mathrm{red}}$.\n\t\n\tDenote the circular subdivisor of $D$ by $R$. Then $\\alpha_{1}^{*}R$ contains a unique circular subdivisor $R''$. Because all fibers are trees, $R''$ contains a horizontal component, say $H$. Suppose $R''-H$ is vertical. Then it is connected and meets $H$ twice, so $H$ is a $2$-section, in fact $H=D''_{\\mathrm{hor}}$; a contradiction with Lemma \\ref{lem:untwisted}. Therefore, $R''$ contains two $1$-sections. In particular, $R''_{\\mathrm{hor}}=D''_{\\mathrm{hor}}$, so $\\alpha_{1}$ is a composition of inner blowups with respect to the images of $R''$, hence it does not touch $\\alpha_{1}^{*}(D-R)$. Because the latter contains no $(-1)$-curves and meets $R''$ in branching components of $D''$, $\\alpha_{2}$ does not touch it, either. In other words, $\\alpha$ is a composition of inner blowups on the images of $R''$ and does not touch the twigs of $D''$.\n\t\n\tIn particular, $D'_{\\mathrm{hor}}$ consists of two $1$-sections, both contained in $R'=(\\alpha_{2})_{*}R''$. \n\t\n\tConsider the case when $D'\\vert$ has a $(-1)$-curve, say $\\Gamma$, and let $F$ be the fiber containing it. Because $D'$ has no superfluous vertical $(-1)$-curves, $\\beta_{D'}(\\Gamma)\\geq 3$, so $\\Gamma\\subseteq R'$ meets a maximal twig $T$ of $D'$. Because $\\alpha^{-1}$ does not touch $T$, we get $\\alpha^{-1}_{*}(T+\\Gamma)=T_{j}+L_{j,0}=[(2)_{d_{j}},1]$ for some $j\\in \\{1,2\\}$. Since $\\beta_{F_{\\mathrm{red}}}(\\Gamma)\\leq 2$, $\\Gamma$ meets $R'_{\\mathrm{hor}}$. The latter consists of $1$-sections, so $\\Gamma$ has multiplicity $1$ in $F$, and thus $\\Gamma$ is a tip of $F$. Because $F$ contracts to a $0$-curve, it follows that $F=[1,(2)_{d_{j}},1]$. The $(-1)$-curve $\\Gamma'=F-T-\\Gamma$ satisfies $\\Gamma'\\cdot D'=1+(F-T-\\Gamma)\\cdot D'_{\\mathrm{hor}}=1$, so $A=\\alpha^{-1}_{*}\\Gamma'$ satisfies $A\\cdot D=A\\cdot T_{j}=1$, thus $A=A_{3-j}$ by Lemma \\ref{lem:C1}. It follows that $F=\\alpha^{*}(A_{j}+T_{j}+L_{3-j,0})$, so our $\\P^{1}$-fibration is in fact induced by a pullback of $|L_{3-j,\\infty}|$. Therefore, $\\alpha=\\mathrm{id}$ and the curves $A_{j}\\cap S$ and $A_{3-j}\\cap S$ are as required.\n\t\n\tConsider the case when $D'\\vert$ contains no $(-1)$-curves. Because $R'$ is circular, $R'_{\\mathrm{hor}}$ has $2-b_{0}(R'\\vert)\\in \\{0,1,2\\}$ nodes. Let $F_r$ be a fiber passing through a node $r\\in R'_{\\mathrm{hor}}$. Then $(F_{r}\\cdot D'_{\\mathrm{hor}})_{r}=2$, so because each connected component of $D'\\vert$ meets $D'_{\\mathrm{hor}}$, $F_{r}\\cap D'=\\{r\\}$. Because $X'\\setminus D'\\cong S$ is affine, $F_{r}\\cap (X'\\setminus D')\\cong \\mathbb{C}^{1}$. The curve $F_{r}$ meets each $1$-section in $D'_{\\mathrm{hor}}$ once, so the above description of $\\alpha$ shows that $\\alpha_{2}$ touches $(\\alpha_{2}^{-1})_{*}F_{r}$ at most once. Therefore, $(\\alpha^{-1}_{*}F_{r})^{2}\\geq F_{r}^{2}-1=-1$, and the equality holds if and only if the proper transform of $F_{r}$ is touched by $\\alpha_{2}$ and not by $\\alpha_1$, in which case $\\alpha^{-1}_{*}F_{r}\\cdot D=1$. Hence $F_{r}\\cap S$ is as in the statement of the lemma. \n\t\n\tSimilarly, if $A\\subseteq X'$ is a vertical $(-1)$-curve such that $A\\cdot D'=1$, then $A\\not\\subseteq D'$ by assumption, and $\\alpha_{2}$ does not touch $(\\alpha_{2}^{-1})_{*}A$, hence $A\\cap S$ is as in the statement, and clearly $A$ is disjoint from any $F_{r}$. Therefore, it remains to find $b_{0}(R'\\vert)$ disjoint vertical $(-1)$-curves $A\\subseteq X'$ such that $A\\cdot D'=1$.\n\t\n\tAssume that every connected component $V$ of $R'\\vert$ is contained in a fiber $F_{V}\\neq V$. Let $A_V\\subseteq F_V$ be a $(-1)$-curve. Then $A_V\\not\\subseteq D'$, $A_V\\cdot D'_{\\mathrm{hor}} \\leq (F_V-V)\\cdot D'_{\\mathrm{hor}}=0$ and $A_V\\cdot D'\\vert=1$ because $F_V$ is a tree, so $A_V\\cdot D'=1$. This ends the proof because $A_V\\cap A_{V'}=\\emptyset$ for $V\\neq V'$.\n\t\n\tThus we can assume that $R'\\vert$ contains a fiber. We have $3\\leq \\#R'\\leq \\#D'=\\rho(X')$ by Lemma \\ref{lem:S0}. Hence $X'$ contains a degenerate fiber $F$. Let $A$ be a $(-1)$-curve in $F$, so $A\\not\\subseteq D'$. \n\t\n\tSuppose that $F\\subseteq D'\\vert+A$. Then $A$ is a unique $(-1)$-curve in $F$, so it has multiplicity $\\mu\\geq 2$ in $F$. Denote our $\\mathbb{C}^{*}$-fibration of $S$ by $f$. Because $D'$ contains a fiber, $f(S)\\subseteq\\mathbb{C}^{1}$, and we can assume that $f(A\\cap S)=\\{0\\}$. Because $\\mathbb{C}[S]$ is a UFD, $A|_{S}$ is a divisor of a regular function, say $g$. Then $f\/g^{\\mu}\\in \\mathbb{C}[S]^{*}=\\mathbb{C}^{*}$, so $f=\\lambda g^{\\mu}$ for some $\\lambda\\in \\mathbb{C}^{*}$. A general fiber of such $f$ is reducible, a contradiction.\n\t\n\tSuppose $L\\cdot D'\\geq 2$ for every $(-1)$-curve $L\\subseteq F$. Then $2\\leq L\\cdot D'\\leq b_0(F\\cap D')$ because $F$ is a tree, and $b_0(F\\cap D')\\leq F\\cdot D'_{\\mathrm{hor}}=2$ because $D'$ is connected. It follows that $A$ is a unique $(-1)$-curve in $F$, and meets two connected components of $F\\cap D'$.\u00a0Because $F$ meets $D'_{\\mathrm{hor}}$ in components of multiplicity $1$, it follows that $F$ is a chain meeting $D'$ in tips contained in $D'\\vert$. Hence $F\\subseteq D'\\vert+A$; a contradiction.\n\t\n\tThus we can assume $A\\cdot D'=1$. If $b_0(R'\\vert)<2$ then the lemma follows, so we can assume $b_0(R'\\vert)=2$. Lemma \\ref{lem:S0}\\ref{item:units} implies that $D'$ contains exactly one fiber, so some connected component $V$ of $R'\\vert$ is contained in a degenerate fiber $F$ as above. Recall that $A\\subseteq F$ is a $(-1)$-curve such that $A\\not\\subseteq D'$, $A\\cdot D'=1$ and $F\\not\\subseteq D'\\vert+A$.\n\t\n\tFor every component $L$ of $F-D'\\vert$ we have $L\\cdot D'_{\\mathrm{hor}}\\leq (F'-V)\\cdot D'_{\\mathrm{hor}}=0$, so since $F$ is a tree, $L\\cdot D'=1$ and all such $L$'s are disjoint. If two of them are $(-1)$-curves then the lemma follows. Hence we can assume that $A$ is a unique $(-1)$-curve in $F$. Put $T=0$ if $A$ meets $V$, otherwise denote by $T$ the maximal twig of $D'$ meeting $A$. If $T\\neq 0$ then $\\alpha^{-1}$ does not touch $A+T$, hence by Lemma \\ref{lem:C1}, $A+T=\\alpha_{*}(A_{j}+T_{j})=[1,(2)_{d_{j}}]$ for some $j\\in \\{1,2\\}$. Moreover, if $T$ meets a curve $L\\not\\subseteq D'$ with $L\\cdot D'=1$ then Lemma \\ref{lem:C1} gives $L=A$. Hence $T\\cdot(F_{\\mathrm{red}}-A-V)=0$. In any case, let $\\tau$ be the contraction of $A+T$. Then $\\tau$ does not touch $F_{\\mathrm{red}}-V$, and the fiber $\\tau_{*}F$ has a unique $(-1)$-curve, contained in $\\tau_{*}V$. Perform further inner contractions with respect to the images of the chain $\\tau_{*}V$ until one of the components of the image of $F$ meeting a $1$-section in the image of $D'_{\\mathrm{hor}}$ becomes a $(-1)$-curve. Then the latter has multiplicity $1$ in the image of $F$, so it is not its unique $(-1)$-curve. It follows that the image of $F$ is of type $[1,1]$ and both its components are in the image of $V$. Hence $F\\subseteq D'\\vert+A$; a contradiction. \t\n\\end{proof}\n\nIn the following, we fix coordinates on $S=S_{p_{1},p_{2}}$, treating it as a closed subset of $\\mathbb{C}^{4}$ given by \\eqref{eq:S}.\n\n\\begin{lem}[Open tori in $S$]\\label{lem:tori}\n\tFix an open subset $U\\subseteq S$ and a ring homomorphism $\\sigma\\colon \\mathbb{C}[x_1,x_2,y_1,y_2]\\to \\mathbb{C}[v_{1}^{\\pm 1},v_{2}^{\\pm 1}]$ as in \\ref{item:AA}--\\ref{item:LC} below. Then $\\sigma$ induces an isomorphism $U\\cong \\mathbb{C}^{*}\\times \\mathbb{C}^{*}$.\n\t\\begin{enumerate}\n\t\t\\item\\label{item:AA} $U=S\\setminus (A_1+A_2)$ and \n\t\t$\\sigma(x_{j})=v_{j}$, $\\sigma(y_{j})= (v_{3-j}-\\hat{p}_{j}(v_{j}))v_j^{-d_{j}}$ for $j\\in \\{1,2\\}$;\n\t\t\\item\\label{item:AL} In case $p_{3-j}=1$ for some $j\\in \\{1,2\\}$:\n\t\t $U=S\\setminus (A_{j}+L_{j,1})$ and\n\t\t \\begin{equation*}\n\t\t \\sigma(x_{j})=v_{j},\\quad\n\t\t \\sigma(x_{3-j})=(v_{j}-1)v_{3-j}^{-1},\\quad\n\t\t \\sigma(y_{j})=(v_{j}-1-\\hat{p}_{j}(v_{j})v_{3-j})\\cdot (v_{j}^{d_j}v_{3-j})^{-1},\\quad\n\t\t \\sigma(y_{3-j})=v_{3-j};\n\t\t \\end{equation*}\n\t\t\\item\\label{item:LC} In case $p_{1}=p_{2}=1$:\n\t\t $U=S\\setminus (C+L_{j,1})$ for some $j\\in \\{1,2\\}$ and\n\t\t \\begin{equation*}\n\t\t \\sigma(x_{j})=-(v_{2}+1)v_{1}^{-1},\\quad\n\t\t \\sigma(x_{3-j})=-(v_{1}+v_{2}+1)\\cdot (v_{1}v_{2})^{-1},\\quad\n\t\t \\sigma(y_{j})=(v_{1}+1)v_{2}^{-1},\\quad\n\t\t \\sigma(y_{3-j})=v_{2}.\n\t\t \\end{equation*}\n\t\\end{enumerate}\nThe volume form $\\tfrac{dv_{1}}{v_{1}}\\wedge \\tfrac{dv_{2}}{v_{2}}$ on $U$ extends, up to a sign, to a volume form $\\tfrac{dx_{1}}{x_{1}}\\wedge \\tfrac{dx_{2}}{x_{2}}$.\n\\end{lem}\n\\begin{proof}\n\tWe check that by a direct computation that $\\sigma$ is zero on the ideal of $S$, so it gives a morphism $\\mathbb{C}^{*}\\times \\mathbb{C}^{*}\\to S$. It has a rational inverse, given by\n\t\\begin{equation*}\n\t\\mbox{\\ref{item:AA} }v_{j}\\mapsto x_{j},\\ j\\in \\{1,2\\};\\quad \n\t\\mbox{\\ref{item:AL} }v_{j}\\mapsto x_{j}, v_{3-j}\\mapsto y_{3-j};\\quad\n\t\\mbox{\\ref{item:LC} }v_{1}\\mapsto y_{1}y_{2}-1, v_{2}\\mapsto y_{3-j}.\n\t\\end{equation*}\n\tThis inverse is regular on the complement of:\n\t\\begin{equation*}\n\t\\mbox{\\ref{item:AA} }\\{x_{1}x_{2}=0\\}=(A_{1}+A_{2})|_{S};\\quad \n\t\\mbox{\\ref{item:AL} }\\{x_{j}y_{3-j}=0\\}=(A_{j}+L_{j,1})|_{S};\\quad\n\t\\mbox{\\ref{item:LC} }\\{y_{3-j}(1-y_{1}y_{2})=0\\}=(L_{j,1}+C)|_{S};\n\t\\end{equation*}\t\n\tand sends $\\tfrac{dv_{1}}{v_{1}}\\wedge \\tfrac{dv_{2}}{v_{2}}$ to $\\pm \\tfrac{dx_{1}}{x_{1}}\\wedge \\tfrac{dx_{2}}{x_{2}}$. This proves the lemma. \n\\end{proof}\n\\begin{rem*}\n\tThe formulas $\\{y_{3-j}=0\\}=L_{j,1}|_{S}$, $\\{y_{1}y_{2}=1\\}=C|_{S}$ used in the above proof do \\emph{not} hold for general $p_{1},p_{2}$, but they do hold under the assumptions of \\ref{item:AL} and \\ref{item:LC}, respectively.\n\\end{rem*}\t\n\n\n\\begin{prop}[{Complete algebraic vector fields on $S_{p_{1},p_{2}}$, cf.\\ \\cite[Theorem 19]{Kal-Ku1_AL-theory}}]\\label{prop:fibrations}\\ \n\t\\begin{enumerate}\n\t\t\\item\\label{item:fibr} Every $\\mathbb{C}^{*}$-fibration of $S$ restricts to a $\\mathbb{C}^{*}$-fibration of one of the open tori $U\\subseteq S$ from Lemma \\ref{lem:tori}\n\t\t\\item\\label{item:extend} For any $U\\subseteq S$ from Lemma \\ref{lem:tori} and any $a_{j}\\geq d_{j}$, $j\\in \\{1,2\\}$, the complete vector field $\\nu_{a_{1},a_{2}}=v_{1}^{a_{1}}v_{2}^{a_{2}}(a_{2}v_{1}\\tfrac{\\d}{\\d v_{1}}-a_{1}v_{2}\\tfrac{\\d}{\\d v_{2}})$ on $U$ extends to a complete vector field on $S$.\n\t\t\\item\\label{item:vf} Every complete algebraic vector field on $S$ restricts to a complete vector field on some torus $U\\subseteq S$ from Lemma \\ref{lem:tori}. In particular, it has divergence zero with respect to the volume form $\\tfrac{dx_1}{x_1}\\wedge \\tfrac{dx_2}{x_2}$.\n\t\t\\item\\label{item:AAut} For each open torus $U\\subseteq S$ from Lemma \\ref{lem:tori}, the group $\\mathrm{AAut}_{\\mathrm{hol}}(S)\\cap \\mathrm{AAut}_{\\mathrm{hol}}(U)$ acts on $U$ $m$-transitively for every $m$.\n\t\\end{enumerate}\n\\end{prop}\n\\begin{proof}\n\t\\ref{item:fibr} Fix a $\\mathbb{C}^{*}$-fibration $p$ of $S$ and let $\\Gamma_{1},\\Gamma_{2}\\subseteq X$ be as in Lemma \\ref{lem:C*_lines}. Then $p$ restricts to a $\\mathbb{C}^{*}$-fibration of $U=S\\setminus (\\Gamma_{1}+\\Gamma_{2})$. We claim that $U$ is as in Lemma \\ref{lem:tori}. Because for $i\\in \\{1,2\\}$ $\\Gamma_{i}\\cap S\\cong \\mathbb{C}^{1}$, $\\Gamma_{i}$ is one of the curves from Lemma \\ref{lem:C1}. If $\\{\\Gamma_{1},\\Gamma_{2}\\}=\\{A_{1},A_{2}\\}$ then $U$ is as in Lemma \\ref{lem:tori}\\ref{item:AA}. Assume that $\\Gamma_{1}=C$. Because $C\\cdot D=2$, by Lemma \\ref{lem:C*_lines} $0\\leq \\Gamma_{1}^{2}=2-d_1-d_2$, so $d_1=d_2=1$, i.e.\\ $p_1=p_2=1$. In this case, $C\\cap S$ meets $A_j \\cap S$ for both $j\\in \\{1,2\\}$, so $\\Gamma_{2}=L_{1,j}$ for some $j\\in \\{1,2\\}$ and therefore $U$ is as in Lemma \\ref{lem:tori}\\ref{item:LC}. We are left with the case when $\\Gamma_{1}=L_{1,j}$ for some $j\\in \\{1,2\\}$ and $\\Gamma_{2}\\neq C$. As before, the condition $\\Gamma_{1}^{2}\\geq -1$ from Lemma \\ref{lem:C*_lines} implies by Lemma \\ref{lem:C1}\\ref{item:L} that $d_{j}=1$, i.\\ e.\\ $p_{j}=1$. In this case, $L_{1,j}\\cap S$ meets $A_{3-j}\\cap S$ and $L_{3-j,1}\\cap S$, so $\\Gamma_{2}=A_{j}$ or $C$. The second case is excluded by assumption, so the first case holds and $U$ is as in Lemma \\ref{lem:tori}\\ref{item:AL}.\n\t\n\t\\ref{item:extend} follows from a direct computation, see \\cite[Corollary 2]{Andersen-tori}.\n\t\n\t\\ref{item:vf} Part \\ref{item:extend} implies that complete algebraic vector fields on $S$ do not share a rational first integral. Hence by \\cite[Theorem 1.6]{KKL_avf}, for any complete algebraic vector field $\\xi$ on $S$ there is a $\\mathbb{C}^{1}$- or $\\mathbb{C}^{*}$-fibration of $S$ preserved by the flow of $\\xi$. Because $\\kappa(S)=0$, $S$ admits no $\\mathbb{C}^{1}$-fibrations, hence $\\xi$ preserves one of the $\\mathbb{C}^{*}$-fibrations from \\ref{item:fibr}. In each case, the complement of $U$ is a union of vertical curves isomorphic to $\\mathbb{C}^{1}$, so the flow of $\\xi$ cannot send them to non-degenerate fibers in $U$. It follows that $\\xi$ restricts to a complete vector field on $U$. In particular, by \\cite[Corollary 1]{Andersen-tori} it preserves the volume form $\\tfrac{dv_{1}}{v_{1}}\\wedge \\tfrac{dv_{2}}{v_{2}}$, where $v_{1},v_{2}$ are coordinates on $U\\cong \\mathbb{C}^{*}\\times \\mathbb{C}^{*}$. Lemma \\ref{lem:tori} implies that $\\xi$ preserves $\\tfrac{dx_{1}}{x_{1}}\\wedge \\tfrac{dx_{2}}{x_{2}}$, too.\n\t\n\t\\ref{item:AAut} Choose integers $a_{ij}$, $i,j\\in \\{1,2\\}$ such that $a_{ij}\\geq d_{j}$ and $\\det [a_{ij}]=1$, and put $\\sigma_{i}=\\nu_{a_{i1},a_{i2}}$, $f_{i}=v_{1}^{a_{i1}}v_{2}^{a_{i2}}\\in\\ker \\sigma_{i}\\setminus \\ker \\sigma_{3-i}$. Now the result follows from \\cite[Propostion 8.9]{Kal-Ku1_AL-theory}. Indeed, although that result is formulated in loc.\\ cit.\\ for $G=\\mathrm{AAut}_{\\mathrm{hol}}(U)$, the proof actually shows $m$-transitivity for the group generated by the elements of the flows of $q(f_{i})\\cdot \\sigma_{i}$ for $q\\in \\mathbb{C}[t]$.\n\\end{proof}\n\n\\begin{proof}[Proof of Theorem \\ref{thm:AAut}]Part \\ref{item:Aut_thm} was shown in Corollary \\ref{cor:st}\\ref{item:Aut}. For \\ref{item:AAut_thm}, note that the open subset $S\\setminus Z$ of $S$ is precisely the union of all $U\\subseteq S$ from Lemma \\ref{lem:tori}. By Proposition \\ref{prop:fibrations}\\ref{item:vf}, $S\\setminus Z$ is fixed by $\\mathrm{AAut}_{\\mathrm{hol}}(S)$. On the other hand, by Proposition \\ref{prop:fibrations}\\ref{item:AAut} each $U\\subseteq S$ from Lemma \\ref{lem:tori} is contained in a single orbit of $\\mathrm{AAut}_{\\mathrm{hol}}(S)$. Because every two such $U$, being open and dense in $S$, have nonempty intersection, they are all contained in the same orbit of $\\mathrm{AAut}_{\\mathrm{hol}}(S)$, which therefore equals $S\\setminus Z$. Proposition \\ref{prop:fibrations}\\ref{item:AAut} implies that $\\mathrm{AAut}_{\\mathrm{hol}}(S)$ acts on this orbit $m$-transitively for every $m$, as claimed.\n\\end{proof}\n\n\\begin{rem}[Density property for $S_{p_{1},p_{2}}$]\n\tIn view of Theorem \\ref{thm:AAut}\\ref{item:AAut_thm}, it is natural to ask for an even stronger property than $m$-transitivity of $\\mathrm{AAut}_{\\mathrm{hol}}(S)$, namely for the \\emph{algebraic density property} (ADP), see \\cite{Kal-Ku1_AL-theory}. Recall that a smooth affine variety $X$ has ADP if the Lie algebra $\\operatorname{VF_{alg}}(X)$ of algebraic vector fields on $X$ coincides with the Lie algebra $\\operatorname{Lie_{alg}}(X)$ generated by complete ones. Since in our case all vector fields preserve $Z$, one should rather ask if the \\emph{relative} ADP holds, see \\cite{KLM_relative-density}, namely, if there exists $l\\geq 0$ such that $I^{l}\\operatorname{VF_{alg}}(S)\\subseteq \\operatorname{Lie_{alg}}(S)$, where $I$ is the ideal of $Z$. This, however, is not true since by Lemma \\ref{lem:tori}\\ref{item:vf} all elements of $\\operatorname{Lie_{alg}}(S)$ preserve the volume form $\\omega=\\tfrac{dx_1}{x_1}\\wedge \\tfrac{dx_2}{x_2}$. Nonetheless, one could ask for the relative \\emph{volume} density property, defined as follows (see \\cite{KK_avdp}). Let $X$ be a normal variety with a volume form $\\omega$, and let $Y\\subseteq X$ be a closed subset containing $\\operatorname{Sing} X$. Denote by $\\operatorname{VF_{alg}^{\\omega}}(X,Y)$ the Lie algebra of those algebraic vector fields on $X$ which vanish on $Y$ and preserve the volume form $\\omega$; and by $\\operatorname{Lie_{alg}^{\\omega}}(X,Y)$ the Lie algebra generated by complete vector fields in $\\operatorname{VF_{alg}^{\\omega}}(X,Y)$. We say that $X$ has an algebraic volume density property (AVDP) relative to $Y$ if there exists $l\\geq 0$ such that $I^{l}\\operatorname{VF_{alg}^{\\omega}}(X,Y)\\subseteq \\operatorname{Lie_{alg}^{\\omega}}(X,Y)$, where $I$ is the ideal of $Y$. In this setting, one could ask the following:\n\\begin{question}\n\tLet $Z\\subseteq S_{p_{1},p_{2}}$ be as in Theorem \\ref{thm:AAut}\\ref{item:AAut_thm}. Does $S_{p_{1},p_{2}}$ have AVDP relative to $Z$?\n\\end{question}\nFor $p_{1}=p_{2}=1$ (so $Z=\\emptyset$), the answer to this question is positive \\cite[Theorem 6]{KK_avdp}\n\\end{rem}\n\n\\bibliographystyle{amsalpha}\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\nIn $1882$, M. Noether claimed the following statement which was later proven by Lefschetz:\nFor $d \\ge 4$, a very general smooth degree $d$ surface $X$ in $\\mathbb{P}^3$ has Picard number $\\rho(X)=1$.\nThis motivates the definition of the \\emph{Noether-Lefschetz locus}, denoted by $\\NL_d$ parametrizing the \nspace of smooth degree $d$ surfaces $X$ in $\\mathbb{P}^3$ with $\\rho(X)>1$.\nOne of the interesting problems is to understand the geometry of the Noether-Lefschetz locus. By the Lefschetz $(1,1)$-theorem, we can look at an irreducible\ncomponent of the Noether-Lefschetz locus locally as a Hodge locus (see \\cite[\\S $5$]{v5} for more details).\nIn particular, denote by $U_d \\subseteq \\mathbb{P}(H^0(\\mathbb{P}^3, \\op))$ \nthe open subscheme parametrizing smooth projective hypersurfaces in $\\mathbb{P}^3$ of degree $d$.\nLet $\\mathcal{X} \\xrightarrow{\\pi} U_d$ be the corresponding universal family. For a given $F \\in U_d$, denote by $X_F$ the surface $X_F:=\\pi^{-1}(F)$. \nLet $X \\in U_d$ and $U \\subseteq U_d$ be a simply connected neighbourhood of $X$ in $U_d$ (under the analytic topology).\nThen, $\\pi|_{\\pi^{-1}(U)}$ induces a variation of Hodge structure $(\\mathcal{H}, \\nabla)$ on $U$, where $\\mathcal{H}:=R^2\\pi_*\\mathbb{Z} \\otimes \n\\mathcal{O}_U$ and $\\nabla$ is the Gauss-Manin connection. Note that $\\mathcal{H}$ defines a local system on $U$ whose fiber over \na point $F \\in U$ is $H^2(X_F,\\mathbb{Z})$. Consider a non-zero element $\\gamma_0 \\in H^2(X_F,\\mathbb{Z}) \\bigcap H^{1,1}(X_F,\\mathbb{C})$\nsuch that $\\gamma_0 \\not= c_1(\\mathcal{O}_{X_F}(k))$ for $k \\in \\mathbb{Z}_{>0}$. \nThis defines a section $\\gamma \\in (\\mathcal{H} \\otimes \\mathbb{C})(U)$. Let $\\overline{\\gamma}$ be the image\nof $\\gamma$ in $\\mathcal{H}\/F^2(\\mathcal{H} \\otimes \\mathbb{C})$. The Hodge loci corresponding to $\\gamma$, denoted $\\NL(\\gamma)$ is then defined as\n\\[ \\NL(\\gamma):=\\{G \\in U | \\overline{\\gamma}_G=0\\},\\]\nwhere $\\overline{\\gamma}_G$ denotes the value at $G$ of the section $\\overline{\\gamma}$. For an irreducible component $L \\subset \\NL_d$ and $X \\in L$, general, we can find $\\gamma \\in H^{1,1}(X,\\mb{Z}):= \nH^2(X,\\mb{Z}) \\bigcap H^{1,1}(X,\\mb{C})$ such that $\\overline{\\NL(\\gamma)}=L$ (the closure taken in $U_d$ under Zariski topology).\n\nOne of the first results in this direction is due to Green, Griffiths, Voisin and others (\\cite{M3, GH, v2}) which states that \nfor an irreducible component $L$ of the Noether-Lefschetz locus, that for $d \\ge 4$,\n\\[d-3 \\le \\codim (L,U_d) \\le \\binom{d-1}{3}.\\] \nThe upper bound follows easily from the fact that $\\dim H^{2,0}(X)=\\binom{d-1}{3}$ for any $X \\in U_d$ (see \\cite[\\S $6$]{v5}).\nWe say that $L$ is a \\emph{general} component if $\\codim L=\\binom{d-1}{3}$ and \\emph{special} otherwise. \nIt was proven by Ciliberto, Harris and Miranda \\cite{ca1} that for $d \\ge 4$, \nthe Noether-Lefschetz locus has infinitely many general components and the union of these components is Zariski dense in $U_d$. \nThe guiding principle of much work in the area has been the expectation that special components should be due to the presence of low degree curves.\nVoisin \\cite{v3} and Green \\cite{M3} independently \nproved that for $d \\ge 5$, $\\codim L=d-3$ if and only if $L$ parametrizes surfaces of degree $d$ containing a line. \nIf $d-3<\\codim L \\le 2d-7$ then $\\codim L=2d-7$ and $L$ parametrizes the surfaces containing a conic.\nOtwinowska \\cite{A1} proved that for an integer $b>0$ and $d \\gg b$ if $\\codim L \\le bd$ then $L$ parametrizes surfaces containing a curve of \ndegree at most $b$.\n\n\n\nFor $r \\ge 3$, we define the level $r$-\\emph{Noether-Lefschetz locus}, denoted $\\NL_{r,d}$ to be the space parametrizing surfaces with \nPicard number greater than or equal to $r$. It has been conjectured by Griffiths and Harris in \\cite{GH} that for $r < d$, an irreducible component of $\\NL_{r,d}$ is \nof codimension greater than or equal to $(r-1)(d-3)-\\binom{r-3}{2}$. Furthermore, the component of $\\NL_{r,d}$ parametrizing\nsurfaces containing $r-1$ coplanar lines is of this codimension. In this article we prove (in Theorem \\ref{gh6}) that:\n\\begin{thm}\\label{ite1}\n Let $r \\ge 3$ and $d \\gg r$. Let $L$ be an irreducible component of $\\NL_{r,d}$. Then, $\\codim L \\ge (r-1)(d-3)-\\binom{r-3}{2}$.\nFurthermore, there exists a component $L$ of $\\NL_{r,d}$ of this codimension parametrizing surfaces containing $r-1$ coplanar lines.\n\\end{thm}\n\nThe techniques used to prove this result is a combination of deformation theory and Hodge theory. Instead of looking at the Hodge locus corresponding to a Hodge class, we study the Hodge locus corresponding to\na $\\mb{Z}$-module of Hodge classes. We then use a result due to Otwinowska, \\cite[Theorem $1$]{ot}, to show that if the codimension of an irreducible component $L$ of $\\NL_{r,d}$\nis less than or equal to $(r-1)(d-3)-\\binom{r-3}{2}$, then for a general $X \\in L$ there exists a lattice $\\Lambda \\subset H^{1,1}(X,\\mb{Z})$ generated by classes of curves of degree less than or\nequal to $r-1$ such that $L$ is locally of the form $\\NL(\\Lambda)$ (see Proposition \\ref{gh14}), where $\\NL(\\Lambda)$ is the intersection of $\\NL(\\gamma)$ for all $\\gamma \\in \\Lambda$,\n$\\bigcap_{\\gamma \\in \\Lambda} \\NL(\\gamma)$.\n\n\n\nWe now use the theory of semi-regularity as introduced in \\cite{b1} to reduce the problem to a question in flag Hilbert schemes. \nFirst to fix some notations, for a Hilbert polynomial $P$ for some curve $C$ in $\\p3$, we denote by $H_P$ the corresponding Hilbert scheme, parametrizing curves (schemes with pure dimension $1$) with \nHilbert polynomial $P$.\nThroughout this article we denote by $Q_d$ the Hilbert polynomial of a degree $d$\nsurface in $\\p3$. We denote by $H_{P,Q_d}$ the corresponding flag Hilbert scheme parametrizing pairs $(C,X)$ such that $C \\in H_P, X \\in H_{Q_d}$ and $C \\subset X$.\nA curve $C$ on a smooth surface in $\\p3$ is said to be \\emph{semi-regular} if $H^1(\\mo_X(C))=0$. \nWe prove that\n\\begin{thm}\n Let $X$ be a smooth surface in $\\p3$ of degree $d$ and $C \\subset X$, a semi-regular curve with Hilbert polynomial, say $P$. For any irreducible component, $L'$ of $\\ov{\\NL([C])}$\n (the closure is taken in $U_d$ under Zariski topology)\n there exists an irreducible component $H'$ of $H_{P,Q_d}$ containing the pair $(C,X)$ such that $\\pr_2(H')_{\\red}$ \n coincides with $L'_{\\red}$, where $\\pr_2$ is \n the second projection map from $H_{P,Q_d}$ to $H_{Q_d}$. In particular, if $C$ is reduced, connected and $d \\ge \\deg(C)+4$ then this holds true and the irreducible component $H'$ is uniquely determined by $L'$.\n\\end{thm}\nSee Theorem \\ref{dim} and Lemma \\ref{hf11} for the precise statements and its proof. Using this proof we further show that under the above \nbound on the codimension of $L$, the lattice will infact be generated by classes of lines (see Lemma \\ref{gh4}). \nFinally, we do a computation in Proposition \\ref{gh13}, to determine the ``arrangement'' of these lines which would \nhelp us determine the component with the correct codimension.\n\n\n\n\n\n{\\bf{Acknowledgement:}} I would like to thank Prof. R. Kloosterman for reading the preliminary version of this article and several helpful discussions. \n\n\\section{Introduction to Noether-Lefschetz locus}\n\n\\begin{para}\n In this section we recall the basic definitions of Noether-Lefschetz locus. See \\cite[\\S $9, 10$]{v4} and \\cite[\\S $5, 6$]{v5} for a detailed presentation of the subject.\n\\end{para}\n\n\n\\begin{note}\\label{n1}\nBy a \\emph{component} of $\\NL_d$, we mean an irreducible component. By a \\emph{surface} we always mean a smooth surface in $\\mathbb{P}^3$.\nDenote by $Q_d$ the Hilbert polynomial of degree $d$ surfaces in $\\p3$. Given, a Hilbert polynomial $P$, denote by $H_P$ the corresponding Hilbert scheme and by $H_{P,Q_d}$ the corresponding flag Hilbert scheme.\nAlso, for a point $u \\in U_d$, denote by $X_u$ the fiber $\\pi^{-1}(u)$.\n\\end{note}\n\n\\begin{note}\\label{nl1}\n Let $X \\in U_d$ and $\\mo_X(1)$, the very ample line bundle on $X$ determined by the closed immersion $X \\hookrightarrow \\p3$ arising (as in \\cite[II.Ex.$2.14$(b)]{R1}) from the graded homomorphism\nof graded rings $S \\to S\/(F_X)$, where $S=\\Gamma_*(\\mo_{\\p3})$ and $F_X$ is the defining equations of $X$.\nDenote by $H_X$ the very ample line bundle $\\mo_X(1)$. \nNote that a very ample line bundle on $X_u$ for any $u \\in U$ remains very ample in the family $\\mc{X}$, hence the corresponding cohomology class remains of type $(1,1)$ \nin $\\mc{X}$.\n\\end{note}\n\n \n \\begin{para}\n Let $X$ be a surface.\n The Lefschetz hyperplane section theorem implies, \\[H^{2}(X,\\mb{C})\\cong H^{2}(X,\\mb{C})_{\\prim} \\oplus \\mb{C}H_X,\\]\n where $H_X$ is the very ample line bundle on $X$ and $H^2(X,\\mb{C})_{\\prim}$ is the primitive cohomology. \n This gives us a natural projection map from $H^{2}(X,\\mb{C})$ to $H^{2}(X,\\mb{C})_{\\prim}$. For $\\gamma \\in H^{2}(X,\\mb{C})$, denote by $\\gamma_{\\prim}$ the image of $\\gamma$ under this morphism.\n Since the very ample line bundle $H_X$ remains of type $(1,1)$ in the family $\\mc{X}$, we can therefore conclude that\n $\\gamma \\in H^{1,1}(X)$ remains of type $(1,1)$ if and only if $\\gamma_{\\prim}$ remains of type $(1,1)$. In particular,\n $\\NL(\\gamma)=\\NL(\\gamma_{\\prim})$.\n \\end{para}\n \n\\begin{defi} \nWe now discuss the tangent space to the Hodge locus, $\\NL(\\gamma)$.\nWe know that the tangent space to $U$ at $X$, $T_XU$ is isomorphic to $H^0(\\N_{X|\\p3})$.\nThis is because $U$ is an open subscheme of the Hilbert scheme $H_{Q_d}$, the tangent space of which at the point $X$ is simply\n$H^0(\\N_{X|\\p3})$. \nGiven the variation of Hodge structure above, we have (by Griffith's transversality) the differential map:\n\\[\\overline{\\nabla}:H^{1,1}(X) \\to \\mathrm{Hom}(T_XU,H^2(X,\\mathcal{O}_X))\\]induced by the Gauss-Manin connection.\nGiven $\\gamma \\in H^{1,1}(X)$ this induces a morphism, denoted $\\overline{\\nabla}(\\gamma)$ from $T_XU$ to $H^2(\\mo_X)$.\nThe tangent space at $X$ to $\\NL(\\gamma)$ is then defined to be $\\ker(\\overline{\\nabla}(\\gamma))$.\n\\end{defi}\n\n\\begin{para}\\label{gr0}\n The boundary map \\[\\rho:H^0(\\N_{X|\\p3}) \\to H^1(\\T_X)\\] arising from the long exact sequence associated to the \n short exact sequence:\n \\[0 \\to \\T_X \\to \\T_{\\p3}|_X \\to \\N_{X|\\p3} \\to 0\\]\n is called the \\emph{Kodaira-Spencer} map. The morphism $\\overline{\\nabla}(\\gamma)$ is related to the Kodaira-Spencer map as we will see below.\n\\end{para}\n\n\\begin{para}\\label{c9}\n Note that there exists a natural cup product morphism, \n \\[H^1(X,\\T_X) \\otimes H^{1}(X,\\Omega^1_X) \\xrightarrow{\\bigcup} H^2(X,\\mathcal{O}_X).\\]\n For $\\gamma \\in H^1(\\Omega_X^1)$ this induces a morphism, denoted $\\bigcup \\gamma$, from $H^1(\\T_X)$ to $H^2(\\mo_X)$.\n We then have the following result in Hodge theory (see \\cite[Theorem $10.21$]{v4}):\n\\end{para}\n\n\\begin{lem}\\label{a3}\nThe differential map $\\overline{\\nabla}(\\gamma)$ conincides with the following:\n\\[T_XU\\cong H^0(\\N_{X|\\p3}) \\xrightarrow{\\rho} H^1(\\T_X) \\xrightarrow{\\bigcup \\gamma} H^2(\\mo_X).\\]\n\\end{lem}\n\n\\section{Hodge locus and Hilbert flag schemes}\n\n\\begin{para}\n In this section we define what is a semi-regular map. We then briefly study Hodge locus for a family of smooth projective\n surfaces in $\\p3$ and show how it is related to certain Hilbert flag schemes. More specifically, we shall study \n the Hodge locus corresponding to certain effective algebraic cycles which will be semi-regular. For such classes we will see that\n the Hodge locus ``coincides'' with a component of a flag Hilbert scheme. We elaborate on the details in this section.\n\\end{para}\n\n\n\\subsection{Semi-regularity map and tangent space to Hodge locus}\n\n\\begin{para}\nWe start with the definition of a semi-regular curve.\n Let $X$ be a surface and $C \\subset X$, a curve in $X$. Since $X$ is smooth, $C$ is local complete intersection in $X$.\n This gives rise to the short exact sequence:\\[0 \\to \\mo_X(-C) \\to \\mo_X \\to i_*\\mo_C \\to 0,\\]where $i$ is the natural inclusion\n morphism from $C$ into $X$. Note that, $\\mo_X(C)$ is locally free $\\mo_X$-module, hence flat. Therefore, tensoring this short exact sequence by $\\mo_X(C)$\n we get\n \\begin{equation}\\label{sh1}\n 0 \\to \\mo_X \\to \\mo_X(C) \\to \\N_{C|X} \\to 0 \n \\end{equation} \n is exact,\n where $\\N_{C|X}$ is the normal sheaf\n $\\Hc_X(\\mo_X(-C),i_*\\mo_C)$ which is isomorphic to the sheaf $i_*\\mo_C \\otimes_{\\mo_X} \\mo_X(C)$ (see \\cite[Ex. II.$5.1$(b)]{R1}). The \\emph{semi-regularity map}\n is the morphism \\[\\pi:H^1(\\N_{C|X}) \\to H^2(\\mo_X)\\]which arises from the long exact sequence associated to the short\n exact sequence (\\ref{sh1}). We say that $C$ is \\emph{semi-regular} if $\\pi$ is injective. \n\n\\end{para}\n\n\n\n\\begin{para}\n The Lefschetz hyperplane section theorem implies that $H^1(\\mo_X)=0$. Then, the long \n exact sequence associated to (\\ref{sh1}) contains the following segment:\n \\[0 \\to H^1(\\mo_X(C)) \\to H^1(\\N_{C|X}) \\xrightarrow{\\pi} H^2(\\mo_X).\\]\n So, $H^1(\\mo_X(C))=0$ is equivalent to $\\pi$ being injective, hence $C$ being semi-regular. We now prove a result that would help\n us determine when a curve is semi-regular.\n\\end{para}\n\n\\begin{lem}\\label{hf11}\nLet $C$ be a connected reduced curve and $d \\ge \\deg(C)+4$ then $h^1(\\mathcal{O}_X(C))=0$. In particular, $C$ is semi-regular.\n\\end{lem}\n\n\\begin{proof}\nSince $X$ is a hypersurface in $\\p3$ of degree $d$, $\\I_X \\cong \\mo_{\\p3}(-d)$. Consider the short exact sequence:\n\\[0 \\to \\I_X \\to \\I_C \\to \\mo_X(-C) \\to 0.\\]\nTensoring this by $\\mo_{\\p3}(k)$, we get the following terms in the associated long exact sequence:\n\\[... \\to H^1(\\I_C(k)) \\to H^1(\\mo_X(-C)(k)) \\to H^2(\\I_X(k)) \\to ...\\]\nNow, $H^2(\\mo_{\\p3}(k-d))=0$ (see \\cite[Theorem $5.1$]{R1}) and $\\I_{C}$ is $\\deg(C)$-regular (see\n\\cite[Main Theorem]{gi1}). \nSo, $H^1(\\I_C(k))=0$ for $k \\ge \\deg(C)$. This implies $H^1(\\mo_X(-C)(k))=0$ for $k \\ge \\deg(C)$. \nBy Serre duality, $0=H^1(\\mo_X(-C)(d-4)) \\cong H^1(\\mo_X(C))$. So, $C$ is semi-regular.\n\\end{proof}\n\n\n\n\n\n\n\n\n\\begin{para}\nLet $X$ be a surface and $C \\subset X$ be a curve.\n We now do a computation to show that for $d \\ge \\deg(C)+4$, $\\dim |C|=0$, where $|C|$ is the linear system of $C$ in $X$. \n\\end{para}\n\n\\begin{lem}\\label{a4e}\nLet $d \\ge 5$ and $C$ be an effective divisor on a smooth degree $d$ surface $X$ of the form $\\sum_i a_iC_i$, where $C_i$ are integral curves with $\\deg(C)+4 \\le d$.\nThen, $h^0(\\N_{C|X})=0.$ In particular, $\\dim |C|=0$, where $|C|$ is the linear system associated to $C$.\n\\end{lem}\n\n\\begin{proof}\nLet $C=\\sum_ia_iC_i$ with $C_i$ integral. Then, \\[\\deg((\\mathcal{O}_X(C)|_C \\otimes \\mo_C)|_{C_i})=a_iC_i^2+\\sum_{j \\not= i} a_jC_i.C_j.\\]\nDenote by $e_i:=\\deg(C_i)$. \nUsing the adjunction formula and the fact that $K_X \\cong \\mathcal{O}_X(d-4)$, we have that \n\\begin{eqnarray*}\n\\deg((\\mathcal{O}_X(C)|_C \\otimes \\mo_C)|_{C_i})&=&2a_i\\rho_a(C_i)-2a_i-(d-4)a_ie_i+\\sum_{j \\not= i}a_jC_i.C_j\\\\\n&\\le&a_i(e_i^2-(d-1)e_i)+\\sum_{j \\not=i}a_jC_iC_j\\\\\n&\\le&a_i(e_i^2-3e_i-e_i\\sum_ja_je_j)+\\sum_{j \\not=i}a_je_ie_j.\n\\end{eqnarray*}\nThe first inequality follows from the bound on the genus of a curve in $\\mathbb{P}^3$ in terms of its degree (see \\cite[Example $6.4.2$]{R1}).\nThe second inequality follows from the facts that $d \\ge \\deg(C)+4$ and $C_i.C_j \\le e_ie_j$. It then follows directly that $\\deg((\\mathcal{O}_X(C)|_C \\otimes \\mo_C)|_{C_i})<0$.\nThis implies that $h^0(C_i,(\\mathcal{O}_X(C)|_C \\otimes \\mc{O}_C)|_{C_i})=0$ for all $i$. So, \n$h^0(\\N_{C|X})=h^0(C,\\mathcal{O}_X(C)|_C \\otimes \\mo_C)=0.$\n\nSince $h^1(\\mathcal{O}_X)=0$ (by Lefschetz hyperplane section Theorem) and $h^0(\\mathcal{O}_X)=1$, using \nthe long exact sequence associated to the short exact sequence \n\\begin{equation}\\label{eq2}\n0 \\to \\mathcal{O}_X \\to \\mathcal{O}_X(C) \\to \\mathcal{O}_X(C)|_C \\otimes \\mo_C \\to 0\n\\end{equation}\nwe get that $h^0(\\mathcal{O}_X(C))=1$. Since $|C|=\\mb{P}(H^0(\\mo_X(C)))$, the lemma follows.\n\\end{proof}\n\n\n\n\n\n\n\n\n\\subsection{Flag Hilbert scheme and Hodge locus}\n\n\\begin{para}\n In this section we introduce the basic definitions of flag Hilbert schemes. See \\cite[\\S $4$]{S1} for further details. We then prove the main result of this section which relates Hodge locus to Hilbert schemes.\n\\end{para}\n\n\n\\begin{para}\nGiven an $m$-tuple of polynomials $\\mathcal{P}(t)=(P_1(t),P_2(t),...,P_m(t))$, we define the contravariant functor, called the \\emph{Hilbert flag functor} \nrelative to $\\mathcal{P}(t)$,\\[FH_{\\mathcal{P}(t)}:(\\mbox{schemes}) \\to \\mbox{sets}\\]\n\\[S \\mapsto \\{(X_1,X_2,...,X_m)|X_1 \\subset X_2 \\subset ... \\subset \\mathbb{P}^3_S\\}\\]such that the Hilbert polynomial of $X_i$ is $P_i(t)$ and $X_i$ is an\n$S$-closed subscheme of $X_{i+1}$. We call such an $m$-\\emph{tuple a flag relative to} $\\mathcal{P}(t)$.\n\\end{para}\n\n\\begin{para}\n The functor $FH_{\\mathcal{P}(t)}$ is representable by a projective scheme, $H_{\\mathcal{P}(t)}$ which parametrizes all such flags relative to $\\mathcal{P}(t)$. We call this the \\emph{Hilbert flag scheme}.\n\\end{para}\n\n\n\n\n\\begin{thm}\\label{dim}\n Let $X$ be a surface, $C$ be a semi-regular curve in $X$ and $\\gamma \\in H^{1,1}(X,\\mb{Z})$ be the class of the curve $C$. For any irreducible component $L'$ of $\\overline{\\NL(\\gamma)}$\n (the closure is taken in the Zariski topology on $U_d$)\n there exists an irreducible component $H'$ of $H_{P,Q_d}$ containing the pair $(C,X)$ such that the associated reduced scheme $\\pr_2(H')_{\\red}$ \n coincides with $L'_{\\red}$, where $\\pr_2$ is \n the second projection map from $H_{P,Q_d}$ to $H_{Q_d}$. Furthermore, if $d \\ge \\deg(C)+4$ then such $H'$ is uniquely determined by $L'$.\n\\end{thm}\n\n\\begin{proof}\nThe first part of the theorem follows directly from \\cite[Theorem $7.1$]{b1}. \n\nFurthermore, Lemma \\ref{a4e} implies that $\\dim |C|=0$. So, given an irreducible component, say $L'$ of $\\overline{\\NL(\\gamma)}$ such that $X$ is a general\n element, there exists an unique irreducible component $H'$ of $H_{P,Q_d}$ containing the pair $(C,X)$ such that $\\pr_2(H')_{\\red}$ coincides with $L'_{\\red}$. \n This proves the rest of the theorem.\n \n \n \n \\end{proof}\n \n \n\n\n\n\n\n\n\n\n\n \n \n\n\n\n\n\n\n\\section{Variation of lattices}\n\\begin{para}\n In this section we give a formula to compute the dimension of an irreducible component of a Hodge locus (see Proposition \\ref{gh12}). \n This result will be particularly useful to prove the asymptotic case of a Griffiths-Harris conjecture, which we see in the next section.\n\\end{para}\n\n\n\\begin{para}\\label{pam6}\nLet $X$ be a surface of degree $d$. \nAn \\emph{augmented lattice } $\\Lambda_X$ \\emph{ on } $X$ \\emph{ of rank } $r$, is a rank $r$ $\\mb{Z}$-submodule $\\Lambda_X \\subset H^2(X,\\mb{Z})$ generated by the class of the very ample line bundle $H_X$ \n(as in \\ref{nl1})\nand cohomology classes of $r-1$ reduced curves, say $C_1,...,C_{r-1}$ such that\n$\\Lambda_X$ is saturated in the sense that for all $\\lambda \\in \\Lambda_X, c \\in \\mb{Q}$ if $c\\lambda \\in H^2(X,\\mb{Z})$ then $c\\lambda \\in \\Lambda_X$. For such $\\Lambda_X$, \nwe say that $C_i$ for $i=1,...,r-1$ \\emph{generate} $\\Lambda_X$.\nWe say that $\\Lambda_X$ is \\emph{prime} if $C_1,...,C_{r-1}$ are integral.\n\\end{para}\n\n\n\n\\begin{para}\\label{pam3}\nLet $\\Lambda_X$ be as in \\ref{pam6}. \nWe can define \\[ \\NL(\\Lambda_X):=\\{G \\in U | \\overline{\\gamma}_G=0, \\mbox{ for all } \\gamma \\in \\Lambda_X\\}.\\]\nFor a surface $X$ and reduced curve $C \\subset X$ such that the cohomology class $[C]$ of $C$ is not a $\\mb{Q}$-multiple of $c_1(H_X)$, denote by $\\Lambda_X^0$ the rank $2$ $\\mb{Z}$-module generated by $[C]$ and \n$c_1(H_X)$, where $c_1$ is the first Chern class map.\nSince a very ample line bundle remains of type $(1,1)$ is the family $\\mc{X}$, $\\NL(\\Lambda_X^0)$ coincides with $\\NL([C])$. More generally, for a rank $r$ augmented lattice $\\Lambda_X$ generated by $C_1,...,C_{r-1}$ we have, $\\NL(\\Lambda_X)$ is isomorphic to the fiber product $\\NL([C_1]) \\times_{H_{Q_d}} ... \\times_{H_{Q_d}} \\NL([C_{r-1}])$.\n\\end{para}\n\n\\begin{para}\\label{pam2}\nLet $\\Lambda_X$ be as before of rank $2$, generated by a reduced curve, say $C$. \nLet $P$ be the Hilbert polynomial of $C$. Assume $d \\ge \\deg(C)+4$. Using Theorem \\ref{dim} we can conclude that for general $X' \\in \\NL(\\Lambda_X)$ there exists a curve $C' \\subset X'$ such that \n$\\overline{\\NL([C'])}$ is an irreducible component of $\\overline{\\NL(\\Lambda_X)}$ and $C'$ deforms to $C$, i.e., $C'$ has the same Hilbert polynomial $P$. Denote by $\\Lambda_{X'}$ the augmented lattice on $X'$ of \nrank $2$ generated by $C'$.\nTheorem \\ref{dim} again implies that there exists an unique irreducible component, denoted $H_{\\Lambda_{X'}}$, of $H_{P,Q_d}$ containing the pair $(C',X')$ such that $\\pr_2(H_{\\Lambda_{X'}})$ is isomorphic to $\\overline{\\NL(\\Lambda_{X'})}$.\nDenote by $L_{\\Lambda_{X'}}:=\\pr_1(H_{\\Lambda_{X'}})$. From now on, we will always assume that $\\overline{\\NL(\\Lambda_X)}$ is irreducible, which is equivalent to $X$ \nbeing general in $\\NL(\\Lambda_X)$, in particular,\naway from the points of intersection of any two irreducible components of $\\overline{\\NL(\\Lambda_X)}$. \n\\end{para}\n\n\\begin{para}\nSuppose now that $\\Lambda_X$ is of rank $r$ generated by $C_1,...,C_{r-1}$. Let $P_1,...,P_{r-1}$ be the Hilbert polynomials of $C_1,...,C_{r-1}$, respectively. Consider the natural morphism \n \\[p:H_{P_1,Q_d} \\times_{H_{Q_d}} ... \\times_{H_{Q_d}} H_{P_{r-1},Q_d} \\to H_{Q_d}.\\] Assume $d \\ge \\sum_{i=1}^{r-1}\\deg(C_i)+4$. Using Theorem \\ref{dim}, we can conclude that for every irreducible component $L'$\n of $\\overline{\\NL(\\Lambda_X)}$ there exists an unique irreducible component, say $H$ of $H_{P_1,Q_d} \\times_{H_{Q_d}} ... \\times_{H_{Q_d}} H_{P_{r-1},Q_d}$ containing $(C_1,X) \\times ... \\times (C_{r-1},X)$ \n such that $p(H)$ coincides with $L'$. Similarly as in \\ref{pam2}, by taking $X$ general in $\\NL(\\Lambda_X)$ we can ensure that $\\overline{\\NL(\\Lambda_X)}$ is irreducible. \n Denote by $H_{\\Lambda_X}$ the irreducible component of $H_{P_1,Q_d} \\times_{H_{Q_d}} ... \\times_{H_{Q_d}} H_{P_r,Q_d}$ such that $p(H_{\\Lambda_X})$ coincides with $\\overline{\\NL(\\Lambda_X)}$. \n Denote by $L_{\\Lambda_X}:=\\pr(H_{\\Lambda_X})$, where $\\pr$ is the natural projection map from $H_{P_1,Q_d} \\times_{H_{Q_d}} ... \\times_{H_{Q_d}} H_{P_{r-1},Q_d}$ to $H_{P_1} \\times ... \\times H_{P_{r-1}}$.\n\\end{para}\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\begin{prop}\\label{gh12}\nLet $r \\ge 3$, $X$ be a surface of degree $d$ and $\\Lambda_X$ be an augmented lattice of rank $r+1$ generated by $r$ reduced curves $C_1,...,C_r$. Assume that $\\sum_{i=1}^r \\deg(C_i)+4 \\le d$.\nThen, the dimension of $\\ov{\\NL(\\Lambda_X)}$ is given by the following formula: \\[\\codim \\ov{\\NL(\\Lambda_X)}=\\codim I_d(C)-\\dim L_{\\Lambda_X}\\] where $C=C_1' \\bigcup ... \\bigcup C_r'$ for a general $r$-tuple $(C_1',...,C_r')$ in $L_{\\Lambda_X}$\nand $I_d(C)$ is the degree $d$ graded piece in the ideal $I(C)$ of $C$.\n\\end{prop}\n\n\\begin{proof}\nConsider the diagram,\n\\[\\begin{diagram}\nH_{\\Lambda_X}&\\rTo^{\\pr_1}&L_{\\Lambda_X}\\\\\n\\dTo^{\\pr_2}& \\\\\n\\ov{\\NL(\\Lambda_X)}\\\\\n\\end{diagram}\\]\nDenote by $P_i$ the Hilbert polynomial of curves $C_i$, respectively. Recall, $L_{\\Lambda_X}$ is contained in $H_{P_1} \\times ... \\times H_{P_r}$. \nFor an $r$-tuple $(C_1,...,C_r) \\in L_{\\Lambda_X}$, the fiber of $\\pr_1$ parametrizes the space of smooth degree $d$ surfaces containing $C=C_1 \\bigcup ... \\bigcup C_r$, which is an open subscheme in $\\mb{P}(I_d(C))$.\nSince $I_d(C)$ is irreducible, the dimension of the generic fiber of $\\pr_1$ is equal to $\\dim I_d(C)-1$, where $(C_1,...,C_r) \\in L_{\\Lambda_X}$ is a general element.\nThe fiber of $\\pr_2$ over $\\pr_2((C_1,...,C_r,X))$ is isomorphic to $|C_1|\\times ... \\times |C_r|$. But, Lemma \\ref{a4e} implies $\\dim |C_i|=0$ for $i=1,...,r$.\nSo, the dimension of the generic fiber of $\\pr_2$ is zero.\nThen, \\[\\dim H_{\\Lambda_X}=\\dim L_{\\Lambda_X}+\\dim I_d(C)-1=\\dim \\NL(\\Lambda_X).\\]\n\\[\\mbox{So, } \\codim \\overline{\\NL(\\Lambda_X)}=\\dim \\mb{P}(H^0(\\mo_{\\p3}(d)))-\\dim \\overline{\\NL(\\Lambda_X)}=h^0(\\mo_{\\p3}(d))-\\dim I_d(C)-\\dim L_{\\Lambda_X}.\\]\nThis finishes the proof of the proposition.\n\\end{proof}\n\n\n\n\n\\section{A Griffiths-Harris conjecture}\n\n\n\n\n\\begin{para}\nWe now come to the final section of the article, where we prove an asymptotic case of a Griffiths-Harris conjecture. Recall,\na \\emph{Griffiths-Harris conjecture} states in \\cite{GH} that: \n\\begin{center}\nFor $3 \\le r \\le d$, the codimension of an irreducible component of $\\NL_{r,d}$ is at least equal to \\[(r-1)(d-3)-\\binom{r-3}{2}.\\]\nFurthermore, there exists a component of $\\NL_{r,d}$ of this codimension parametrizing the space of surfaces \ncontaining $r-1$ lines on the same plane. \n\\end{center}\n\\end{para}\n\n\\begin{note}\nWe will denote by $\\N_d(r)$ the number, \\[(d-3)(r-1)-\\binom{r-3}{2}.\\]We now recall a result in Noether-Lefschetz locus due to Otwinowska which will help us characterize the irreducible components of $\\NL_{r,d}$ with codimension\nless than or equal to $\\N_d(r)$.\n\\end{note}\n\n\n\\begin{thm}[{\\cite[Theorem $1$]{ot}}]\\label{gh22}\n Let $\\gamma$ be an augmented lattice of rank $2$ on a degree $d$ surface. There exists $C \\in \\mathbb{R}_+^*$ depending only on $r$ such that for $d \\ge C(r-1)^8$ if $\\codim \\NL(\\gamma) \\le (r-1)d$ \n then $\\gamma_{\\prim}=\\sum_{i=1}^ta_i[C_i]_{\\prim}$, where $a_i \\in \\mathbb{Q}^*$, $C_i$ are reduced curves and $\\deg(C_i) \\le (r-1)$ for\n $i=1,...,t$ for some positive integer $t$.\n\\end{thm}\n\n\\begin{para}\nThroughout this section we denote by $r$ an integer greater than or equal to $3$ and for a fixed $r$, denote by $d$, an integer as mentioned in Theorem \\ref{gh22}. We will \\emph{assume} \nthat $d$ is at least $r^3$ which will be used only in a computation in Lemma \\ref{gh4}. The other results do not have any restriction on $d$ in terms of $r$.\n\\end{para}\n\n\\begin{prop}\\label{c3}\nLet $L$ be an irreducible component of $\\NL_{r,d}$. Then $L$ is locally homeomorphic to $\\NL(\\Lambda)$ for some prime augmented lattice $\\Lambda$ of \nrank at least $r$ on a surface $X \\in L$, general.\n\\end{prop}\n\n\\begin{proof}\nLet $L \\subset \\NL_{r,d}$ be an irreducible component. Let $X$ be a general element in $L$. This implies that for the Picard lattice $\\Lambda:=\\NS(X)$, $\\NL(\\Lambda)_{\\red}$ is an open subscheme of $L_{\\red}$.\nWe can assume that $\\Lambda$ is a prime lattice. Since $X$ is an element in $\\NL_{r,d}$, the rank of $\\Lambda$ is greater than or equal to $r$.\n\\end{proof}\n\n\n\n\\begin{prop}\\label{gh14}\nIf $\\Lambda$ is an augmented prime lattice of rank $t$ on some degree $d$ surface and $\\codim \\overline{\\NL(\\Lambda)} \\le (r-1)d$. \nThen there exists a prime lattice $\\Lambda'$ of rank greater than or equal to $t$ generated by classes of curves of degree less than or equal to $r-1$,\nsuch that $C_i$ deforms along $\\NL(\\Lambda')$ and\n$\\overline{\\NL(\\Lambda)}_{\\red}=\\overline{\\NL(\\Lambda')}_{\\red}$.\n\\end{prop}\n\n\\begin{proof}\n\n\n\n\n \nLet $X \\in \\NL(\\Lambda)$.\nThere exists a maximal lattice $\\Lambda' \\subset H^2(X,\\mb{Z})$ such that $\\Lambda'$ remains of type $(1,1)$ in $\\NL(\\Lambda)$ i.e., $\\NL(\\Lambda)_{\\red}=\\NL(\\Lambda')_{\\red}$. \nNow, there exists \na surface $X' \\in \\NL(\\Lambda')$ such that the N\\'{e}ron-Severi group $\\mr{NS}(X')$ is the translate (under deformation from $X$ to $X'$) of \n$\\Lambda'$ in $H^2(X',\\mb{Z})$ which we again denote by $\\Lambda'$ for convinience of notation. Then Theorem \\ref{gh22} implies that any $\\gamma \\in \\Lambda'$ is of the form $\\sum_i a_i[C_i]+bH_X$ with $\\deg(C_i) \\le r-1$.\nSo, $\\Lambda'$ can be generated by classes of curves of degree at most $r-1$ and the class of the very ample line bundle $H_X$. \nNow, the class of $[C_i]$ remains of type $(1,1)$ along $\\NL(\\Lambda')$. From Lemma \\ref{hf11} it follows that $C_i$ is semi-regular. Then, \\cite[Theorem $7.1$]{b1}\nimplies that the class of $[C_i]$ remains effective along $\\NL(\\Lambda')$.\nThis proves the proposition.\n\\end{proof}\n\n\\begin{para}\n We now recall a result due to Eisenbud and Harris which we use in the next lemma. Let $P$ be a Hilbert polynomial of a curve in $\\p3$ of degree $e$ and $L$ be an irreducible component of $H_P$.\n The corollary after \\cite[Theorem $1$]{eis} tells us that,\n \\begin{thm}[\\cite{eis}]\\label{ei1}\n For $e>1$, the dimension of $L$ is less than or equal to $3+e(e+3)\/2$.\n \\end{thm}\n\\end{para}\n\n\n\\begin{lem}\\label{gh4}\nLet $\\Lambda$ be a prime augmented lattice of rank $t+1$ on a degree $d$ surface, generated by irreducible curves $C_i$ for $i=1,...,t$ for some positive integer $t$\nand $\\deg(C_i) \\le r-1$. Suppose $\\codim \\NL(\\Lambda)\\le (r-1)d$. Then, $\\sum_{i=1}^t \\deg(C_i) \\le (r-1)$.\n\\end{lem}\n\n\\begin{proof}\nWe prove this by induction on $t$. This is trivially true for $t=1$.\nSuppose this is true for all $t \\le m$. \n\nAssume this is not true for $t=m+1$. In other words, there exists a prime lattice $\\Lambda$ minimally generated by $m+1$ curves such that \n$\\sum_i \\deg(C_i)>r-1$. \nThis implies (after rearranging the indices if necessary) there exists an integer $0 (r-1)$ and $\\sum_{i=1}^{t'-1} \\deg(C_i) \\le (r-1)$.\nThen, $\\sum_{i=1}^{t'} \\deg(C_i) \\le 2(r-1)$.\nDenote by $P$ the Hilbert polynomial of the curve $C_1 \\bigcup ... \\bigcup C_{t'}$.\nWe replace $e$ by $2(r-1)$ in Theorem \\ref{ei1} and conclude that the dimension of the Hilbert scheme $H_P$ \nis less than or equal to $3+(r-1)(2r+1)$. Using Proposition \\ref{gh12}, the codimension \nof $\\NL([C_1+...+C_{t'}])$ is greater than or equal to $\\codim I_d(C_1 \\bigcup ... \\bigcup C_{t'})-\\dim H_P$. Since $r-1<\\deg(C_1+...+C_{t'})\\le 2(r-1)$ and $d \\ge r^3$, we get the following inequality using the upper bound on \nthe arithmetic genus of a curve of degree less than or equal to $2(r-1)$:\n\\begin{eqnarray*}\n \\codim \\NL([C_1+...+C_{t'}]) & \\ge &\\codim I_d(C_1+...C_{t'})-\\dim H_P\\\\\n &\\ge& (rd-(2r-3)(2r-4)\/2+1)-(3+(r-1)(2r+1))\\\\\n &=&(r-1)d+(d-(2r-3)(2r-4)\/2-3-(r-1)(2r+1)+1)\\\\\n &>&(r-1)d\n\\end{eqnarray*}\ncontradicting the assumption.\n\\end{proof}\n\n\n\n\n\\begin{prop}\\label{gh13}\nLet $\\Lambda$ be an augmented lattice of rank $r$ contained in a degree $d$ surface, generated by $l_i$ for $i=1,...,r-1$, where $l_i$ are lines for \nall $i$. Suppose $r \\ge 3$. Then, $\\codim \\NL(\\Lambda) \\ge \\N_d(r)$. Furthermore, if $l_i$ are on the same plane we have\nan equality.\n\\end{prop}\n\n\\begin{proof}\nWe prove this by induction.\n Using Proposition \\ref{gh12}, \\[\\codim \\NL(\\Lambda)=\\codim I_d(\\bigcup_{i=1}^{r-1} l_i)-\\dim L_{\\Lambda}.\\]\n If $r=3$, $\\codim \\NL(\\Lambda)=2d-6=\\N_d(3)$.\n Assume the result holds true for all $r \\le m$ for some integer $m \\ge 4$. We now prove for $r=m+1$.\nDenote by $\\Lambda'$ the lattice generated by $l_i$ for $i=1,...,r-2$ and by $C$ the curve $\\bigcup_{i=1}^{r-2} l_i$. \nLet $t:=\\dim L_{\\Lambda}-\\dim L_{\\Lambda'}$. Note that, $t \\le 4$. \n\nNote that $l_{r-1}.C \\le r-2$.\nDenote by $\\epsilon:=r-2-l_{r-1}.C$. Comparing with the list of values of $t$, we see \n\\begin{enumerate}\n \\item If $\\epsilon=0$ then $l_{r-1}.C=r-2$ and $t \\le 2$. In particular, for a fixed curve $C$ there is a $1$-$1$ correspondence between the \n set of choices of $l_{r-1}$ intersecting $C$ in $r-2$ points and the set of planes $P$ interesting $C$ at $r-2$ distinct collinear points. If for a generic choice of $P$, $P \\bigcap C$ are $r-2$ distinct collinear points\n then all lines in $C$ should lie on the same plane. In that case, $l_{r-1}$ intersect $C$ at $r-2$ points if and only if $l_{r-1}$ is on the same plane as $C$, hence $t=2$ (dimension of the space of lines in $\\mb{P}^2$). \n If this is not the case i.e., \n a generic choice of $P$ does not intersect $C$ in $r-2$ distinct collinear points, then $t \\le 2$.\n\\item $\\epsilon=r-2$ and $t=4$ if $l_{r-1}$ does not intersect $C$.\n\\item $0 < \\epsilon