diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzoebm" "b/data_all_eng_slimpj/shuffled/split2/finalzzoebm" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzoebm" @@ -0,0 +1,5 @@ +{"text":"\\section{INTRODUCTION}\nLiquid helium ($^4$He) has a reputation for being the first substance in which\none is able to observe many macroscopic quantum phenomena. In particular, it\nwas the first system that could sustain superfluid flow, \\cite{bulk} and as a\nconsequence display a number of amazing properties such as second sound,\nquantized vortices and the fountain effect. Furthermore, thin superfluid helium\nfilms were the first two-dimensional systems experimentally proven to undergo a\nKosterlitz-Thouless\ntransition to the normal state. \\cite{film} More recently it may\nhave been observed \\cite{Mochel} that on weakly-binding substrates these films\nare the first-known spatially ordered superfluids. \\cite{VAV}\n\nMore precisely, measurements of the third-sound resonance\nfrequency (which is proportional to the square root of the superfluid density)\nof submonolayer helium films on hydrogen and deuterium substrates apparently\nindicate two independent Kosterlitz-Thouless transitions: the usual superfluid\nto normal transition at a temperature $T_{KT}$ that obeys the expected\nuniversal jump relation, \\cite{NK} and a second new transition at a temperature\n$T_c$ which is roughly $0.5\\,\\,T_{KT}$ for all coverages. The second transition\nappears as a sharp (but not discontinuous) rise or dip in the superfluid\ndensity depending on the substrate.\n\nIn an attempt to explain these experimental results we have\nrecently proposed that below the second critical temperature the superfluid\nhelium film is in a spatially ordered phase exhibiting both off-diagonal\n(superfluid) and diagonal (hexatic) long-range order in the one-particle\ndensity matrix. \\cite{PRL} The main idea behind this proposal is that the\nhexatic to fluid transition is known to be a Kosterlitz-Thouless transition\ndriven by disclination unbinding. \\cite{NH} (Disclinations are defects in the\norientational order of a crystal created by the insertion or removal of a wedge\nof atoms, as shown in Fig.\\ 1.) Therefore, our physical picture of the\nexperiments is that at sufficiently low temperatures the film is in a\nsuperhexatic phase with only a dilute gas of bound vortices and bound\ndisclinations present due to thermal fluctuations. For entropic reasons the\ndisclinations then unbind at $T_c$, leading to a transition from a superhexatic\nto a superfluid phase since the vortices remain bound at this transition and\nthe presence of free disclinations destroys the hexatic long-range order. At\n$T_{KT}$ the vortices then also unbind and the film is finally forced into the\nnormal liquid phase.\n\nOf course, to make sure that the above picture is qualitatively correct we must\nalso consider the interaction between vortices and disclinations. This is even\nmore pressing if one realizes that in a supersolid phase (where all\ndisclination pairs are themselves bound into pairs or triples) this interaction\nis of long range and depends logarithmically on the distance between the two\nkinds of defects. Fortunately, it turns out that this is no longer true in the\nsuperhexatic phase due to the screening of the interaction by the surrounding\ngas of disclination pairs. A renormalization-group analysis actually shows that\nthe vortex-disclination interaction is irrelevant and that the two separate\nKosterlitz-Thouless transitions indeed survive. Nevertheless, the superfluid\ndensity is influenced in a non-universal way by the unbinding of the\ndisclinations and Monte-Carlo simulations even show that on the basis of our\nhypothesis a rough qualitative agreement with the experiments of Chen and\nMochel can be obtained. \\cite{PRL}\n\nHowever, to definitely identify the phase below $T_c$ more detailed information\nis needed. As a first step towards this goal we here present the\ntwo-dimensional hydrodynamic equations of a superhexatic by describing the\nsuperhexatic as a supersolid with free dislocations (i.e.\\ disclination pairs\n\\cite{NH}). As a result of this approach we will also be able to consider the\nhydrodynamics of the supersolid phase, for which there is at present a renewed\ninterest both in the context of Josephson-junction arrays \\cite{Anne} and solid\n$^4$He. \\cite{LG} Moreover, spatially ordered superfluid states have recently\nbeen proposed to be also relevant for the fractional quantum Hall effect,\n\\cite{B} since this effect can be understood as a condensation of composite\nbosons. \\cite{Z} We therefore believe that the methods developed below might,\nif extended to bosons interacting with a Chern-Simons gauge field, also be used\nto obtain a description of the dynamics of such exotic quantum Hall states.\n\nWe have organized the paper in the following manner. In Sec.\\ \\ref{PD} we first\nconsider the normal solid and hexatic phases by formulating a gauge theory that\ndescribes the phonons, the dislocations and the interaction between them. From\nthis theory we then deduce in Sec.\\ \\ref{SOP} for both phases the dynamics of\nthe appropriate hydrodynamic degrees of freedom. In Sec.\\ \\ref{SP} we\nincorporate the effects of the additional superfluid order parameter\n\\cite{fluid} into the hydrodynamic equations derived in Sec.\\ \\ref{SOP} and\ndiscuss the various long-wavelength modes in the supersolid and superhexatic\nphases obtained in this manner. We conclude in Sec.\\ \\ref{DC} with a discussion\non the possible relevance of our work to future experiments on submonolayer\nhelium films and with a physical interpretation of our results.\n\n\\section{GAUGE THEORY OF PHONONS AND DISLOCATIONS}\n\\label{PD}\nIn this section we will derive the long-wavelength (quantum) dynamics of the\nsolid and hexatic phases. The discussion closely follows work by Kleinert,\n\\cite{K1} save that we will not include higher gradient elasticity. This leads\nto a considerable simplification of the theory but implies that we cannot\nproperly treat the dynamics of the disclinations. Fortunately, for our purposes\nonly the dynamics of the dislocations is of importance and this simplification\nis justified.\n\n\\subsection{Solid}\n\\label{S}\nIn the case of an isotropic crystal, the action for the\ndisplacement field $u_i(\\vec{x},\\tau)$ in the presence of a pair of\ndislocations is given by \\cite{K1}\n\\begin{equation}\n\\label{action}\nS[u_i] = \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n\\left\\{ \\frac{\\rho}{2} (\\partial_{\\tau}u_i - \\beta_i)^2\n + \\mu \\left(u_{ij} - \\frac{\\beta_{ij} + \\beta_{ji}}{2}\n \\right)^2\n + \\frac{\\lambda}{2} (u_{ii} - \\beta_{ii})^2 \\right\\}~,\n\\end{equation}\nwhere $u_{ij} = (\\partial_i u_j + \\partial_j u_i)\/2$ is the strain tensor,\n$\\mu$ and $\\lambda$ are the usual Lam\\'e coefficients \\cite{La} and $\\rho$ is\nthe average mass density. The unphysical (and singular) contributions arising\nfrom the multivaluedness of $u_i(\\vec{x},\\tau)$ are compensated by the\nquantities $\\beta_j$ and $\\beta_{ij}$ (also known as the `plastic distortion').\nTheir relationship to the defects is best explained by the Volterra\nconstruction. \\cite{K2} Let $\\cal C$ be a small loop bounding a section of two\ndimensional crystal that is excised from the whole (cf. Fig.\\ 2). The edges of\nthe loop are drawn together and form a line $\\cal L$. This line may be time\ndependent, and its definition is not unique. However, the topological defects\n(i.e.\\ two dislocations with opposite Burgers' vectors) associated with the\ndistortion of the surface are always located at the endpoints of $\\cal L$. If\n$\\pm \\vec{B}$ are the Burgers' vectors of the dislocations constituting the\npair and if $\\vec{v}$ is their velocity then\n$\\beta_{ij} = \\delta_i({\\cal L})B_j$ and\n$\\beta_j = - v_i \\delta_i({\\cal L})B_j$. The delta function $\\delta_i({\\cal\nL})$ is singular on the time-dependent Volterra cutting line ${\\cal L}$ of the\ndislocations and is directed along the normal vector. If the cutting line\n${\\cal L}$ is parameterized by $\\vec{x}(s,\\tau)$ with $0 \\leq s \\leq 1$, this\nmeans mathematically that\n\\begin{equation}\n\\delta_i({\\cal L}) = - \\epsilon_{ij} \\int_0^1 ds~\n \\frac{\\partial x_j(s,\\tau)}{\\partial s}~\n \\delta(\\vec{x} - \\vec{x}(s,\\tau))~,\n\\end{equation}\nwhere $\\epsilon_{ij}$ is the two-dimensional antisymmetric tensor. Note that\nthe dislocations are assumed to be able to move freely, without any friction,\nthrough the crystal because the equations of motion for the displacement field\nallow for time-dependent solutions that precisely correspond to such evolutions\nof the crystal. \\cite{Na} We will come back to the issue of friction in Sec.\\\n\\ref{SOP} when we consider the effects of dissipation.\n\nWe now first perform a Hubbard-Stratonovich transformation by introducing the\nauxillary variable $\\vec p$ (representing the momentum density) and adding the\nquadratic term\n\\begin{eqnarray}\n\\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\frac{1}{2\\rho} (p_i - i\\rho(\\partial_{\\tau}u_i - \\beta_i))^2\n \\nonumber\n\\end{eqnarray}\nto the action, which may now be rewritten as\n\\begin{eqnarray}\nS[p_i,u_i] = \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n\\left\\{ \\frac{p_i^2}{2\\rho} \\right.\n &+& \\mu \\left(u_{ij} - \\frac{\\beta_{ij} +\n \\beta_{ji}}{2} \\right)^2\n \\nonumber \\\\\n &+& \\left. \\frac{\\lambda}{2} (u_{ii} - \\beta_{ii})^2\n - ip_i(\\partial_{\\tau}u_i - \\beta_i) \\right\\}~.\n\\end{eqnarray}\nIntegrating out $\\vec{p}$ would return the original action up to an unimportant\nconstant. In a similar manner we then also introduce the symmetric stress\ntensor $\\sigma_{ij}$, to decouple the terms quadratic in the strain. This\nresults in\n\\begin{eqnarray}\nS[p_i,\\sigma_{ij},u_i] =\n \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n\\left\\{ \\frac{p_i^2}{2\\rho} \\right.\n &+& \\frac{1}{4\\mu} \\left(\\sigma_{ij}^2 -\n \\frac{\\nu}{1+\\nu} \\sigma_{ii}^2 \\right)\n \\nonumber \\\\\n&-& \\left. ip_i(\\partial_{\\tau}u_i - \\beta_i)\n + i\\sigma_{ij} \\left( u_{ij} -\n \\frac{\\beta_{ij} + \\beta_{ji}}{2}\n \\right) \\right\\}~,\n\\end{eqnarray}\nwith $\\nu = \\lambda\/(2\\mu + \\lambda)$. The partition function is now given by\nthe functional integral\n\\begin{equation}\nZ = \\int d[p_i] \\int d[\\sigma_{ij}] \\int d[u_i]~\n \\exp \\left\\{-\\frac{1}{\\hbar} S[p_i,\\sigma_{ij},u_i] \\right\\}~,\n\\end{equation}\nwhere the integration over $\\sigma_{ij}$ is only over the symmetrical part\nsince we have not included higher gradient elasticity.\n\nWe can now perform the integration over the displacement field. Because the\naction is linear in $u_i$ this simply leads to the constraint\n\\begin{equation}\n\\label{con}\n\\partial_{\\tau}p_j = \\partial_i \\sigma_{ij}~.\n\\end{equation}\nThis constraint can be automatically satisfied if we introduce the vector field\n$A_j$ and the tensor field $A_{ij}$ by setting\n$\\sigma_{ij} = \\epsilon_{ik} \\partial_k A_j\n + \\epsilon_{ki} \\partial_\\tau A_{kj}$\nand $p_j = \\epsilon_{ki} \\partial_i A_{kj}$. Substituting these relations into\nthe action we find that the interaction between the gauge fields (i.e.\\ the\nphonons) and the dislocations is\ngiven by\n\\begin{equation}\n\\label{Sint}\nS_{int}[A_{ij},A_j] = \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\{ -iA_i \\alpha_i + iA_{ij} J_{ij} \\}~,\n\\end{equation}\nwhere after several partial integrations the unphysical singularities of\n$\\beta_i$ and $\\beta_{ij}$ have disappeared and only the dislocation density\nand the dislocation current density\nremain. Introducing also the function\n$\\delta({\\cal P}) = \\delta(\\vec{x}(1,\\tau)) -\n \\delta(\\vec{x}(0,\\tau))$,\nwhich denotes the difference between a delta function at one end of the cutting\nline ${\\cal L}$ and a delta function at the other end, these densities and\ncurrents can conveniently be written as\n$\\alpha_j = \\delta({\\cal P})B_j$ and\n$J_{ij} = - v_i \\delta({\\cal P})B_j$, respectively. As a direct consequence of\nthe above definitions they obey the conservation law\n\\begin{equation}\n\\label{Claw}\n\\partial_{\\tau} \\alpha_j = \\partial_i J_{ij}~.\n\\end{equation}\n\nIn addition, the dynamics of the phonons is determined by the remaining\nquadratic terms in the action which expressed in terms of the gauge fields\n$A_j$ and $A_{ij}$ yield\n\\begin{equation}\nS_0[A_{ij},A_j] = \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\left\\{ \\frac{(\\epsilon_{ik} \\partial_k A_{ij})^2}{2\\rho}\n + \\frac{1}{4\\mu} \\left(\\sigma_{ij}^2 -\n \\frac{\\nu}{1+\\nu} \\sigma_{ii}^2 \\right) \\right\\}~,\n\\end{equation}\nwith $\\sigma_{ij}$ equal to\n$\\epsilon_{ik}(\\partial_k A_j - \\partial_\\tau A_{kj})$.\nComparing this result with Eq.\\ (\\ref{action}) we observe that the stress and\nthe physical part of the strain\n$u^{Phys}_{ij} \\equiv u_{ij} - (\\beta_{ij} + \\beta_{ji})\/2$\nare related by\n$\\sigma_{ij} = 2\\mu u^{Phys}_{ij} + \\lambda \\delta_{ij}\n u^{Phys}_{kk}$\nand therefore by\n\\begin{equation}\n\\label{uphys}\nu^{Phys}_{ij} = \\frac{1}{2\\mu} \\left( \\sigma_{ij}\n - \\frac{\\nu}{1+\\nu} \\delta_{ij} \\sigma_{kk} \\right)~.\n\\end{equation}\nWe will have need of the latter relation in Sec.\\ \\ref{SOP}, when we discuss\nhydrodynamics. A more formal way to justify it is to add to the action\n$S[u_i]$ a source term\n\\begin{eqnarray}\n\\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n K_{ij} \\left(u_{ij} - \\frac{\\beta_{ij} + \\beta_{ji}}{2} \\right)\n = \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n K_{ij} u^{Phys}_{ij} \\nonumber\n\\end{eqnarray}\nand perform the same manipulations as before. We then find that the source\n$K_{ij}$ indeed couples linearly to the right-hand side of Eq.\\ (\\ref{uphys}).\n\nFollowing Kleinert, we now notice that the above theory has\na gauge symmetry as a result of the fact that the gauge fields\n$A_i$ and $A_{ij}$ are not uniquely determined if the stresses $\\sigma_{ij}$\nand momenta $p_i$ are known. Indeed,\n$\\sigma_{ij}$ and $p_i$ are invariant under the gauge transformation $A_i\n\\rightarrow A_i + \\partial_{\\tau} \\Lambda_i$ and $A_{ij} \\rightarrow A_{ij} +\n\\partial_i \\Lambda_j$. Hence $S_0[A_{ij},A_j]$ is also invariant. Moreover, due\nto the conservation law in Eq.\\ (\\ref{Claw}), the interaction\n$S_{int}[A_{ij},A_j]$ is invariant too.\n\nTo calculate the partition function we therefore need some gauge-fixing\nprocedure. The symmetry of $\\sigma_{ij}$ requires that\n\\begin{equation}\n\\epsilon_{ij} \\sigma_{ij} = \\partial_j A_j\n - \\partial_{\\tau}(A_{jj}) = 0~.\n\\end{equation}\nWe would now like to write the gauge fields as the appropriate derivatives of\nunconstrained fields. Using the above gauge symmetry we can always take\n$A_i = \\epsilon_{ij} \\partial_j \\chi$ and $A_{ii}=0$.\nThis, however, does not completely fix the gauge because these conditions are\nstill invariant under the smaller group of transformations\n$\\chi \\rightarrow \\chi + \\partial_{\\tau} \\Lambda$ and\n$A_{ij} \\rightarrow A_{ij} + \\partial_i (\\epsilon_{jk}\n \\partial_k \\Lambda)$.\nTo see more clearly the consequences of this residual symmetry we\nexpand $A_{ij}$ into its longitudinal and transverse components (with respect\nto both indices), i.e.\n\\begin{equation}\nA_{ij} = \\partial_i (\\partial_j A^{LL})\n + \\partial_i (\\epsilon_{jk} \\partial_k A^{LT})\n + \\epsilon_{ik} \\partial_k (\\partial_j A^{TL})\n + \\epsilon_{ik} \\partial_k\n (\\epsilon_{jl} \\partial_l A^{TT})~,\n\\end{equation}\nwhere we have introduced four new fields. The tracelessness of $A_{ij}$ can\nthen be fulfilled by taking $A^{LL} = -A^{TT}$. In addition, the residual gauge\nsymmetry can now be written as\n$\\chi \\rightarrow \\chi + \\partial_{\\tau} \\Lambda$ and\n$A^{LT} \\rightarrow A^{LT} + \\Lambda$. This shows that instead of the fields\n$\\chi$ and $A^{LT}$ we must use the gauge-invariant field $\\chi' \\equiv \\chi\n- \\partial_{\\tau} A^{LT}$ together with $\\Lambda$ as integration variables. The\nassociated change of\nmeasure can be incorporated in the normalization and the same is true for the\n`volume' $\\int d[\\Lambda]$ of the residual gauge group because the action is\ngauge invariant and therefore cannot depend on $\\Lambda$. After this\ngauge-fixing procedure the partition function thus becomes\n\\begin{equation}\nZ = \\int d[A^{TT}] \\int d[A^{TL}] \\int d[\\chi']~\n \\exp \\left\\{-\\frac{1}{\\hbar} \\left(S_0[A^{TT},A^{TL},\\chi'] +\n S_{int}[A^{TT},A^{TL},\\chi'] \\right) \\right\\}~.\n\\end{equation}\nNote that we are left with three physical degrees of freedom, which is the\ncorrect number in two dimensions since $\\sigma_{ij}$ and $p_i$ contain in\nprinciple a total of five degrees of freedom but we have two constraints in\nEq.\\ (\\ref{con}). Note also that the transformation from $\\sigma_{ij}$ and\n$p_i$ to $A^{TT}$, $A^{TL}$ and $\\chi'$ is a linear one so that the Jacobian\ninvolved in the calculation of the partition function is simply an unimportant\nconstant. In particular, the stress is given by\n\\begin{equation}\n\\label{stress}\n\\sigma_{ij}= \\epsilon_{ik}\\epsilon_{j\\ell}\\partial_k \\partial_\\ell \\chi'\n + \\partial_\\tau\\left(\n \\partial_i \\partial_j A^{TL}\n + \\epsilon_{ik} \\partial_k \\partial_j A^{TT}\n + \\epsilon_{jk} \\partial_k \\partial_i A^{TT}\n \\right)~,\n\\end{equation}\nwhich is manifestly symmetric in $i$ and $j$.\n\nA straightforward calculation now shows that the free part of the action is\n\\begin{eqnarray}\n\\label{s0}\nS_0[A^{TT},A^{TL},\\chi'] &=&\n \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\left\\{ \\frac{1}{2\\mu} (\\partial_{\\tau} \\partial^2 A^{TT})^2\n + \\frac{1}{2\\rho} (\\partial_i \\partial^2 A^{TT})^2\n \\right. \\nonumber \\\\\n &+& \\left.\n \\frac{1}{4\\mu(1+\\nu)} (\\partial_{\\tau} \\partial^2 A^{TL})^2\n + \\frac{1}{2\\rho} (\\partial_i \\partial^2 A^{TL})^2\n + \\frac{1}{4\\mu(1+\\nu)} (\\partial^2 \\chi')^2\n \\right. \\nonumber \\\\\n &-& \\left.\n \\frac{1}{2\\mu} \\frac{\\nu}{1+\\nu}\n (\\partial_{\\tau} \\partial^2 A^{TL})(\\partial^2 \\chi')\n \\right\\}~.\n\\end{eqnarray}\nIt contains four modes: The part involving $A^{TT}$ has a pair of modes\n(corresponding to $\\pm \\vec{k}$) with\n$\\omega^2 = \\mu \\vec{k}^2\/\\rho$. These modes therefore represent the transverse\nphonons with a speed of sound $\\sqrt{\\mu\/\\rho}$. The part involving $A^{TL}$\nand $\\chi'$ has another pair of modes with a dispersion obeying\n$\\omega^2 = (2\\mu + \\lambda) \\vec{k}^2\/\\rho$.\nThese represent the longitudinal phonons with a speed of sound $\\sqrt{(2\\mu +\n\\lambda)\/\\rho}$. Interestingly, these results can be understood much more\neasily if we introduce the field\n\\begin{equation}\n\\label{chidp}\n\\chi'' \\equiv \\chi' - \\nu \\partial_{\\tau} A^{TL}~,\n\\end{equation}\nsince then the above action becomes\n\\begin{eqnarray}\n\\label{uncoupled}\nS_0[A^{TT},A^{TL},\\chi''] &=&\n \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\left\\{ \\frac{1}{2\\mu} (\\partial_{\\tau} \\partial^2 A^{TT})^2\n + \\frac{1}{2\\rho} (\\partial_i \\partial^2 A^{TT})^2\n \\right. \\nonumber \\\\\n &+& \\left.\n \\frac{1-\\nu}{4\\mu} (\\partial_{\\tau} \\partial^2 A^{TL})^2\n + \\frac{1}{2\\rho} (\\partial_i \\partial^2 A^{TL})^2\n + \\frac{1}{4\\mu(1+\\nu)} (\\partial^2 \\chi'')^2\n \\right\\}~,\n\\end{eqnarray}\nso that the fields are completely uncoupled. Notice that the $\\chi''$ field has\nno kinetic term, which explains why we obtained above only four modes instead\nof six, as might have been expected in first instance.\n\nFurthermore, if we introduce the usual Burgers' field $\\vec{b}(\\vec{x},\\tau)$\nfor the total dislocation density, which is nothing more than the sum of the\ndensity $\\alpha_i$ over all dislocation pairs, then the interaction with the\ndislocations acquires the form\n\\begin{equation}\nS_{int}[A^{TT},A^{TL},\\chi'] =\n \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n i \\left\\{ \\chi' \\epsilon_{ij} \\partial_j b_i\n - A^{TT} \\partial_{\\tau} \\partial_i b_i\n \\right\\}~,\n\\end{equation}\nwhere we have made use of Eq.\\ (\\ref{Claw}) to express the longitudinal part\nof the current density $J_{ij}$ in terms of the time derivative of $b_i$.\nDecomposing $\\vec{b}$ into its transverse and longitudinal parts, i.e.\\\n$b_i = \\partial_i b^L + \\epsilon_{ij} \\partial_j b^T$, the interaction finally\nbecomes\n\\begin{equation}\n\\label{int}\nS_{int}[A^{TT},A^{TL},\\chi'] =\n \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n i \\left\\{ \\chi' \\partial^2 b^T\n - A^{TT} \\partial_{\\tau} \\partial^2 b^L\n \\right\\}~.\n\\end{equation}\nThe total action $S = S_0 + S_{int}$ reduces for time-independent\nconfigurations to the one we previously used for a discussion of the critical\nproperties of the superhexatic. \\cite{PRL} Integrating out the fields $A^{TT}$\nand $\\chi'$ we can now find the time-dependent interaction among the\ndislocations. Physically, these interactions are thus associated with phonon\nexchange and the time dependence arises due to the finite speeds of sound. This\npicture also explains why the effective action for $\\chi'$ contains just\none pair of modes: The self-interaction of the\ntransverse dislocation density can only be mediated by\nlongitudinal phonons.\n\n\\subsection{Hexatic}\n\\label{Hex}\nUp to this point the dislocation density has not been an independent\ndynamical variable, since we have specified the positions of the dislocations\nat all times and thus neglected the influence of the phonon dynamics on\ntheir motion. However, to describe the hexatic phase we want to\nintegrate out the dislocations in the\nplasma (or continuous) approximation. \\cite{NH} For that we need the free\naction of the field $\\vec{b}$. Here we can again make use of the results\nobtained by Kleinert, who showed that the energy associated with the nonlinear\nstresses at the heart of the defect can be lumped into a `core contribution' to\nthe action. \\cite{K1,K3} In our notation this contribution becomes\n\\begin{eqnarray}\n\\label{dislo}\nS_0[b_i] &=& \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\frac{E_c}{2} b_i\n \\left( \\frac{\\rho}{\\mu}\n \\frac{\\partial_{\\tau}^2}{\\partial^2} + 1 \\right) b_i\n \\nonumber \\\\\n &=& \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\frac{E_c}{2} \\left\\{\n b^T \\left( \\frac{\\rho}{\\mu} \\partial_{\\tau}^2\n + \\partial^2 \\right) b^T\n + b^L \\left( \\frac{\\rho}{\\mu} \\partial_{\\tau}^2\n + \\partial^2 \\right) b^L\n \\right\\} ~.\n\\end{eqnarray}\nThis action represents free propagation of the dislocation density fluctuations\nwhich, as mentioned previously, are permitted by the classical equations of\nmotion \\cite{Na} and neglects dissipative coupling of the dislocation cores to\nthe phonons.\n\nIntegrating out the Burgers' field using Eqs.\\ (\\ref{int}) and (\\ref{dislo}),\nwe obtain the following results. The effective action for $A^{TT}$ becomes\n\\begin{eqnarray}\nS^{eff}[A^{TT}] =\n \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\left\\{ \\frac{1}{2\\mu} (\\partial_{\\tau} \\partial^2 A^{TT})^2\n \\right. + \\frac{1}{2\\rho} (\\partial_i \\partial^2 A^{TT})^2\n \\hspace*{1.0in} \\nonumber \\\\\n + \\left.\n \\frac{1}{2E_c} (\\partial_{\\tau} \\partial_i A^{TT})\n \\left(\n \\frac{\\rho}{\\mu}\\frac{\\partial_{\\tau}^2}{\\partial^2} + 1\n \\right)^{-1}\n (\\partial_{\\tau} \\partial_i A^{TT})\n \\right\\}~.\n\\end{eqnarray}\nAs shown in Fig.\\ 3a, it contains two pairs of modes which for\n$\\vec{k}^2 \\gg \\mu\/2E_c$ all have a dispersion obeying\n$\\omega^2 \\simeq \\mu \\vec{k}^2\/\\rho$. However, for\n$\\vec{k}^2 \\ll \\mu\/2E_c$, one pair of modes has a dispersion\n$\\omega^2 \\simeq \\mu^2\/E_c\\rho + 2\\mu\\vec{k}^2\/\\rho$ with a gap whereas\nthe other pair of modes is gapless with\n$\\omega^2 \\simeq 2E_c \\vec{k}^4\/\\rho$. This is consistent with our expectation\nthat in the hexatic phase there should only be one pair of transverse\ngapless modes with a softer dispersion than that of the transverse phonon\nmodes in a true solid.\n\nThe effective action for $\\chi'$ and $A^{TL}$ in the hexatic phase is\n\\begin{eqnarray}\nS^{eff}[A^{TL},\\chi'] &=&\n \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\left\\{\n \\frac{1}{4\\mu(1+\\nu)} (\\partial_{\\tau} \\partial^2 A^{TL})^2\n \\right.\n + \\frac{1}{2\\rho} (\\partial_i \\partial^2 A^{TL})^2\n \\nonumber \\\\\n &+& \\frac{1}{4\\mu(1+\\nu)} (\\partial^2 \\chi')^2\n - \\frac{1}{2\\mu} \\frac{\\nu}{1+\\nu}\n (\\partial_{\\tau} \\partial^2 A^{TL})(\\partial^2 \\chi')\n \\nonumber \\\\\n &+& \\left.\n \\frac{1}{2E_c} (\\partial_i \\chi')\n \\left(\n \\frac{\\rho}{\\mu}\\frac{\\partial_{\\tau}^2}{\\partial^2} + 1\n \\right)^{-1} (\\partial_i \\chi')\n \\right\\}~.\n\\end{eqnarray}\nIntegrating out also $A^{TL}$ we finally arrive at the effective action for\n$\\chi'$. It reads\n\\begin{eqnarray}\nS^{eff}[\\chi'] = \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\left\\{ \\frac{1}{4\\mu(1+\\nu)} (\\partial^2 \\chi')^2\n + \\frac{1}{2E_c} (\\partial_i \\chi')\n \\left(\n \\frac{\\rho}{\\mu}\\frac{\\partial_{\\tau}^2}{\\partial^2}\n + 1\n \\right)^{-1} (\\partial_i \\chi')\n \\right. \\hspace*{0.5in} \\nonumber \\\\\n \\left.\n + \\frac{1}{2} \\left(\\frac{1}{2\\mu}\n \\frac{\\nu}{1+\\nu} \\right)^2\n (\\partial_{\\tau} \\partial^2 \\chi')\n \\left( \\frac{\\partial_{\\tau}^2}{2\\mu(1+\\nu)}\n + \\frac{\\partial^2}{\\rho} \\right)^{-1}\n (\\partial_{\\tau} \\partial^2 \\chi')\n \\right\\}\n\\end{eqnarray}\nand also contains two pairs of modes (cf. Fig.\\ 3b).\nFor $\\vec{k}^2 \\gg \\mu\/2E_c$ we\nrecover of course the ordinary sound dispersions\n$\\omega^2 \\simeq 2\\mu\\vec{k^2}\/(\\rho(1-\\nu))\n = (2\\mu +\\lambda)\\vec{k}^2\/\\rho$\nand $\\omega^2 \\simeq \\mu\\vec{k}^2\/\\rho$. However, in the limit\n$\\vec{k}^2 \\ll \\mu\/2E_c$ these evolve into a pair of gapped modes with\n$\\omega^2 \\simeq 2\\mu^2\/(E_c\\rho(1-\\nu))$ and a pair of propagating modes\nwith $\\omega^2 \\simeq 2\\mu(1+\\nu)\\vec{k}^2\/\\rho$, respectively.\nClearly, the same mode structure is\nalso present in the effective action for $A^{TL}$ (obtained by integrating out\n$\\chi'$ instead of $A^{TL}$) which indicates that in the\nhexatic phase the longitudinal velocity is renormalized downwards to\n$\\sqrt{2\\mu(1+\\nu)\/\\rho}$.\n\n\\section{HYDRODYNAMICS OF SPATIALLY ORDERED PHASES}\n\\label{SOP}\nWe now turn to the linear hydrodynamics of the solid and hexatic phases that\nfollows from the theory presented above. For the sake of clarity, and because\nit will turn out to be less important for our purposes, we will not discuss\ntemperature fluctuations in the following. However, having derived the\nrelevant energy\ndensities in Secs. \\ref{S} and \\ref{Hex} it is in principle straightforward to\ninclude temperature fluctuations in our theory and, in particular, to arrive at\nthe extension of the hydrodynamic equations presented below that is required if\none wants to consider also the hydrodynamic mode due to energy conservation.\nAfter the equations of motion for the hydrodynamic variables are determined,\nwe can find the propagating and diffusive modes. This is done as before,\nby Fourier transforming the equations of motion and determining the dispersion\n$\\omega(k)$.\nPropagating modes appear as complex roots of a characteristic equation and\nwill always occur in pairs. Each physically distinct propagating excitation\nsuch as longitudinal or transverse sound corresponds therefore to two\nroots or modes.\n\nWe start by considering the mass-density fluctuation $\\delta\\rho$ above the\naverage mass density $\\rho$ and initially neglect the possible presence of\nvacancies and interstitials. To lowest order in the strain, the density\nfluctuation equals $-\\rho u^{Phys}_{ii}$ so up to that order we obtain\n\\begin{equation}\n\\partial_{\\tau} \\delta\\rho = - \\rho \\partial_{\\tau} u^{Phys}_{ii}\n = - \\frac{\\rho}{2(\\mu + \\lambda)} \\partial_{\\tau}\n \\sigma_{ii}~,\n\\end{equation}\nif we make use of Eq.\\ (\\ref{uphys}) to relate the stress and the strain. Using\nalso the decomposition\n$\\sigma_{ii} = \\partial^2 \\chi' + \\partial_{\\tau}\n \\partial^2 A^{TL}$ from\nEq.\\ (\\ref{stress})\nwe may write this as a pair of continuity equations\n\\begin{mathletters}\n\\label{hydro}\n\\begin{equation}\n\\partial_{\\tau} \\delta\\rho = \\partial_i g_i^L~,\n\\end{equation}\n\\begin{equation}\n\\partial_{\\tau} g_j^L = \\partial_i \\pi_{ij}^D~,\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nwhere the longitudinal momentum density is given by\n\\begin{equation}\ng_i^L = - \\frac{\\rho}{2(\\mu + \\lambda)} \\partial_{\\tau}\n \\partial_i (\\chi' + \\partial_{\\tau} A^{TL})\n\\end{equation}\nand the diagonal part of the stress tensor by\n\\begin{equation}\n\\pi_{ij}^D = - \\frac{\\rho \\delta_{ij}}{2(\\mu + \\lambda)}\n \\partial_{\\tau}^2 (\\chi' + \\partial_{\\tau} A^{TL})~.\n\\end{equation}\n\nIn the absence of defects, the hydrodynamic quantities $g_i^L$ and\n$\\pi_{ij}^D$ are precisely equal to the longitudinal part of $p_i$ and the\ndiagonal part of $\\sigma_{ij}$, respectively. This can be seen in the\nfollowing manner. In an ideal solid we have no defects, and variation\nof the action in Eq.\\ (\\ref{uncoupled}) gives $\\chi''=0$ or\n$\\chi'=\\nu \\partial_\\tau A^{TL}$, which we may use to eliminate $\\chi'$.\nThe equation of motion for $A^{TL}$ generated in this way is\n\\begin{equation}\n\\label{ATLmotion}\n\\partial_{\\tau}^2 A^{TL} = - \\frac{2\\mu + \\lambda}{\\rho}\n \\partial^2 A^{TL}.\n\\end{equation}\nIf we substitute this back into the definitions of $g_i^L$ and $\\pi_{ij}^D$\nwe obtain the longitudinal part of $p_i$ and the\ndiagonal part of $\\sigma_{ij}$ as given in section \\ref{PD}.\nIn the presence of defects with their own dynamics this is no longer true,\nsince the dislocation density couples to the gauge fields and alters\nthe equations of motion. To avoid confusion about this point we have,\ntherefore, introduced a new notation for the hydrodynamic\nmomentum density and stress tensor which we will use for the rest of\nthe paper.\n\nWe also note that the above equations\nare not Galilean invariant and are therefore only valid in a specific reference\nframe. This is a result of the fact that the gauge theory of Sec.\\ \\ref{PD} has\nimplicitly used the existence of an ideal lattice with respect to which the\ndisplacements $\\vec{u}(\\vec{x},\\tau)$ are defined. \\cite{K2} Hence, the\nprefered reference frame corresponds to that frame in which this ideal lattice\nis at rest. This is the case for all the hydrodynamic equations presented\nbelow.\n\nIn this ideal solid without interstitials or vacancies the pressure\nfluctuation (following from\n$\\pi_{ij}^D = - \\delta_{ij} \\delta p$) equals\n\\begin{equation}\n\\label{press}\n\\delta p = \\frac{\\rho}{2(\\mu + \\lambda)}\n (1 + \\nu) \\partial_{\\tau}^3 A^{TL}\n = \\frac{\\rho}{2\\mu + \\lambda} \\partial_{\\tau}^3\n A^{TL}\n\\end{equation}\nand the mass-density fluctuation becomes\n\\begin{equation}\n\\delta\\rho = - \\frac{\\rho}{2(\\mu + \\lambda)}\n (1 + \\nu) \\partial_{\\tau} \\partial^2 A^{TL}\n = - \\frac{\\rho}{2\\mu + \\lambda}\n \\partial_{\\tau} \\partial^2 A^{TL}~.\n\\end{equation}\nSubstituting the equation of motion Eq.\\ (\\ref{ATLmotion})\nfor $A^{TL}$ into Eq.\\ (\\ref{press}), we obtain the desired constitutive\nequation\n\\begin{equation}\n\\delta p = \\frac{2\\mu + \\lambda}{\\rho} \\delta\\rho~.\n\\end{equation}\nTogether with the hydrodynamic equations (\\ref{hydro}) this correctly leads to\nthe sound equation\n\\begin{equation}\n\\label{sound}\n\\partial_{\\tau}^2 \\delta\\rho\n = - \\frac{2\\mu + \\lambda}{\\rho} \\partial^2 \\delta\\rho\n = - c_{||}^2 \\partial^2 \\delta\\rho~,\n\\end{equation}\nwith $c_{||}$ the longitudinal sound velocity.\n\nHowever, as stressed first by Martin, Parodi, and Pershan \\cite{M} and again by\nZippelius, Halperin, and Nelson \\cite{ZHN} we are not in general allowed to\nassume that the crystal is ideal, without vacancies or interstitials. We must\ninclude the effects of (long-wavelength) fluctuations in the net defect density\n$n_{\\Delta}$, which is defined as the density\nof vacancies minus the density of interstitials. To do so we can make use of\nthe fact that in a hexagonal system these defects can be regarded as a `bound\nstate' of three dislocations with radial Burgers' vectors pointing\nsymmetrically outward (interstitial) or inward (vacancy). \\cite{N} This is\nillustrated for an interstitial in Fig.\\ 4. As a result the interaction of the\nnet defect density with the gauge fields is given by\n\\begin{equation}\nS_{int}[A^{TT},A^{TL},\\chi'] =\n \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n i \\frac{\\gamma_{\\Delta}}{\\mu} n_{\\Delta} \\partial^2\n \\chi'~,\n\\end{equation}\nwhere $V_0$ denotes the area deficit induced by a defect in an otherwise\nperfect crystal and $\\gamma_\\Delta =\\mu {c_{||}^2} {V_0}\/{2}{c_{\\perp}^2}$.\nWe can verify this result by noting that in the static case (and $n_{\\Delta}\n\\rightarrow in_{\\Delta}$ because of our conventions in the imaginary time\nformalism of Sec.\\ \\ref{PD}) the Euler-Lagrange equation for the Airy stress\nfunction, following from the action in Eq.\\ (\\ref{s0}) together with the above\ninteraction, becomes\n$\\partial^2 \\chi = 2(\\mu + \\lambda) V_0 n_{\\Delta}$\nwhich correctly leads to\n$\\int d\\vec{x}~u^{Phys}_{ii} = V_0 \\int d\\vec{x}~n_{\\Delta}$.\nFurthermore, the free action of the defects becomes\n(cf. Eq.\\ (\\ref{dislo}))\n\\begin{equation}\n\\label{free}\nS_0[n_{\\Delta}] = \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\frac{E_{\\Delta}}{2} n_{\\Delta}\n \\left( \\frac{\\rho}{\\mu} \\frac{\\partial_{\\tau}^2}{\\partial^2}\n + 1 \\right) n_{\\Delta}~,\n\\end{equation}\nwith $E_{\\Delta}$ of order $E_c V_0$.\n\nRedoing our calculations with $n_{\\Delta}$ non-zero, we find that\n$n_{\\Delta}$ displaces the $\\chi'$ field. Therefore $\\chi''$ in Eq.\\\n(\\ref{chidp}) is also non-zero. Moreover, we now obtain instead of Eq.\\\n(\\ref{sound}) the coupled set of equations\n\\begin{mathletters}\n\\label{sounds}\n\\begin{equation}\n\\partial_{\\tau}^2 \\delta\\rho =\n - c_{||}^2 \\left(\n 1 + \\frac{\\nu \\gamma_{\\Delta}^2}\n {E_{\\Delta} \\mu}\n \\right) \\partial^2 \\delta\\rho\n + i \\gamma_{\\Delta}\n \\left(\n 1 - \\frac{2 \\gamma_{\\Delta}^2}\n {E_{\\Delta} \\mu}\n \\right) \\partial^2 n_{\\Delta}~,\n\\end{equation}\n\\begin{equation}\n\\partial_{\\tau}^2 n_{\\Delta} =\n - c_{\\perp}^2 \\left(\n 1 + \\frac{2 \\gamma_{\\Delta}^2}\n {E_{\\Delta} \\mu}\n \\right) \\partial^2 n_{\\Delta}\n + i \\frac{\\gamma_{\\Delta}\\lambda}\n {E_{\\Delta}\\rho^2} \\partial^2 \\delta\\rho~,\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nfor the longitudinal degrees of freedom. Note that the density fluctuation\n$\\delta\\rho$ receives a contribution from both the lattice vibrations as well\nas\nfrom the net defect density, since\n\\begin{equation}\n\\delta\\rho =\n - \\frac{1}{c_{||}^2} \\partial_{\\tau} \\partial^2 A^{TL}\n + \\frac{2i \\gamma_{\\Delta}}{c_{||}^2} n_{\\Delta}~.\n\\end{equation}\nAs a result the longitudinal momentum density has also two contributions\n\\begin{equation}\ng^L_i =\n - \\frac{1}{c_{||}^2} \\partial_i \\partial_{\\tau}^2 A^{TL}\n - \\frac{2i \\gamma_{\\Delta}}{c_{||}^2} J^L_i~,\n\\end{equation}\nwhere $\\vec{J}^L$ is the longitudinal part of the net defect current density\nobeying the continuity equation\n$\\partial_{\\tau} n_{\\Delta} = - \\partial_i J^L_i$.\n\nThis almost completes our discussion of the hydrodynamical description (without\ndissipation) of the solid phase. However, we have not yet obtained the\ntransverse modes. From our expressions for the strain tensor $u_{ij}$ one can\neasily show that in the solid phase the transverse part of the displacement\nfield is given by\n\\begin{equation}\n\\label{disp}\nu^T_i = \\frac{1}{\\mu} \\epsilon_{ij}\n \\partial_j (\\partial_{\\tau} A^{TT})\n = \\frac{1}{\\rho c_{\\perp}^2} \\epsilon_{ij}\n \\partial_j (\\partial_{\\tau} A^{TT})~,\n\\end{equation}\nwhere $c_{\\perp}$ is the transverse speed of sound. Hence, the transverse\ndynamics of the lattice is solely determined by the transverse phonons and we\nhave the additional hydrodynamic equation\n\\begin{equation}\n\\partial_{\\tau}^2 A^{TT} =\n - c_{\\perp}^2 \\partial^2 A^{TT}~,\n\\end{equation}\nwhich is completely uncoupled from the previous ones and in particular does not\ndepend on the net defect density $n_{\\Delta}$. Moreover, if we introduce the\nstandard hexatic order parameter $\\vartheta_6$, which is equal to the local\nbond angle and may therefore be written as\n\\begin{equation}\n\\vartheta_6 \\equiv \\frac{1}{2} \\epsilon_{ij} \\partial_i u_j\n = - \\frac{1}{2\\rho c_{\\perp}^2}\n (\\partial_{\\tau} \\partial^2 A^{TT})~,\n\\end{equation}\nthe equation for $A^{TT}$ is equivalent to\n\\begin{equation}\n\\partial_{\\tau}^2 \\vartheta_6 =\n - c_{\\perp}^2 \\partial^2 \\vartheta_6~,\n\\end{equation}\nso that $\\vartheta_6$ can also be used to describe the transverse phonons.\n\n{}From Eq.\\ (\\ref{disp}) we also find that the transverse part of the momentum\ndensity is given by\n\\begin{equation}\ng_i^T = \\frac{1}{c_{\\perp}^2} \\epsilon_{ij}\n \\partial_j (\\partial_{\\tau}^2 A^{TT})~.\n\\end{equation}\nTherefore the stress tensor has the nondiagonal contribution\n\\begin{equation}\n\\pi_{ij}^{ND} = - \\frac{1}{c_{\\perp}^2} \\epsilon_{ij}\n (\\partial_{\\tau}^3 A^{TT})\n = 2\\rho c_{\\perp}^2 \\epsilon_{ij} \\vartheta_6~,\n\\end{equation}\nand both the longitudinal as well as the transverse hydrodynamic equations in\nthe solid phase can be summarized by\n\\begin{mathletters}\n\\begin{equation}\n\\frac{\\partial \\delta\\rho}{\\partial t} = -\\nabla \\cdot \\vec{g}~,\n\\end{equation}\n\\begin{equation}\n\\label{momS}\n\\frac{\\partial \\vec{g}}{\\partial t} =\n - c^2 \\nabla \\delta\\rho\n - \\gamma_{\\Delta} \\nabla n_{\\Delta}\n + 2\\rho c_{\\perp}^2 \\nabla \\times \\vartheta_6~,\n\\end{equation}\n\\begin{equation}\n\\label{transS}\n\\frac{\\partial \\vartheta_6}{\\partial t} =\n \\frac{1}{2\\rho} \\nabla \\times \\vec{g}~,\n\\end{equation}\n\\begin{equation}\n\\frac{\\partial n_{\\Delta}}{\\partial t} =\n -\\nabla \\cdot \\vec{J}~,\n\\end{equation}\n\\begin{equation}\n\\label{longS}\n\\frac{\\partial \\vec{J}}{\\partial t} =\n - c_{\\Delta}^2 \\nabla n_{\\Delta}\n + \\gamma \\nabla \\delta\\rho~,\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nafter a transformation to real time, which in this case not only means that\n$\\tau \\rightarrow it$ but also\n$\\vec{g} \\rightarrow i\\vec{g}$ and\n$n_{\\Delta} \\rightarrow in_{\\Delta}$. Moreover, note that the constants\n$c$, $\\gamma_{\\Delta}$, $c_{\\Delta}$, and $\\gamma$ should here be\ninterpreted as renormalized quantities which are determined in terms of the\nmicroscopic parameters of our gauge theory by a comparison with\nEq.\\ (\\ref{sounds}).\n\nNow we are ready to discuss dissipation. In principle dissipation has already\nbeen included because there is a coupling between the net defect density\n$n_{\\Delta}$ and the phonons. Hence if for example an interstitial were, in a\ndiscrete picture, to tunnel from one location to another there would be a\n`shake up' of the phonon field. However, if we treat $n_{\\Delta}$ as a smooth\ncontinuously varying field, the action in Eq.\\ (\\ref{free}) is quadratic and\nthe bilinear coupling\n$n_{\\Delta} \\partial^2 \\chi'$ produces only mixing of the collective modes\nbut no real dissipation. Therefore, we choose to include effective dissipation\nin the same manner as explained in detail by Zippelius, Halperin and Nelson.\n\\cite{ZHN} Using their notation we first of all add to the right-hand side of\nEq.\\ (\\ref{momS}) the terms\n$(\\eta \\nabla^2 \\vec{g}\n + \\zeta \\nabla (\\nabla \\cdot \\vec{g}) )\/\\rho$\nassociated with the dissipative part of the stress tensor $\\pi_{ij}$ and\nrepresenting viscous diffusion of the momentum density.\n\nNext the question arises how we need to modify\nEq.\\ (\\ref{longS}). This equation is a result of the fact that we have allowed\nthe dislocations, and hence the interstitials and vacancies, to move freely\nthrough the lattice and used Eq.\\ (\\ref{free}) for the free action of the\ndefects. If the defects effectively experience friction (for example due to\ntheir interaction with the phonons), then it is more appropriate to add a\nLeggett friction term \\cite{TL} and use\n\\begin{equation}\n\\label{diss}\nS_0[n_{\\Delta}] = \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\frac{E_{\\Delta}}{2} n_{\\Delta}\n \\left( \\frac{\\rho}{\\mu}\n \\frac{\\partial_{\\tau}^2}{\\partial^2}\n + i \\frac{\\rho}{\\mu} \\xi \\partial_{\\tau} + 1\n \\right) n_{\\Delta}\n\\end{equation}\ninstead. The dispersions then indeed obey\n$\\omega^{\\pm} \\simeq \\pm c_{\\perp} k - i \\xi k^2\/2$ at long wavelengths, and we\nmust add the term\n$\\xi \\nabla (\\nabla \\cdot \\vec{J})$ to the right-hand side of Eq.\\\n(\\ref{longS}). If we further assume that the transverse part of the defect\ncurrent density behaves as in a gas and simply diffuses to zero with a\ndiffusion constant $\\kappa$, we obtain in total\n\\begin{mathletters}\n\\label{HS}\n\\begin{equation}\n\\frac{\\partial \\delta\\rho}{\\partial t} = -\\nabla \\cdot \\vec{g}~,\n\\end{equation}\n\\begin{equation}\n\\frac{\\partial \\vec{g}}{\\partial t} =\n - \\frac{B}{\\rho} \\nabla \\delta\\rho\n - \\gamma_{\\Delta} \\nabla n_{\\Delta}\n + 2\\rho c_{\\perp}^2 \\nabla \\times \\vartheta_6\n + \\frac{\\eta}{\\rho} \\nabla^2 \\vec{g}\n + \\frac{\\zeta}{\\rho} \\nabla (\\nabla \\cdot \\vec{g})~,\n\\end{equation}\n\\begin{equation}\n\\frac{\\partial \\vartheta_6}{\\partial t} =\n \\frac{1}{2\\rho} \\nabla \\times \\vec{g}~,\n\\end{equation}\n\\begin{equation}\n\\frac{\\partial n_{\\Delta}}{\\partial t} =\n -\\nabla \\cdot \\vec{J}~,\n\\end{equation}\n\\begin{equation}\n\\label{diff}\n\\frac{\\partial \\vec{J}}{\\partial t} =\n - c_{\\Delta}^2 \\nabla n_{\\Delta}\n + \\gamma \\nabla \\delta\\rho\n + \\kappa \\nabla^2 \\vec{J}\n + \\xi \\nabla (\\nabla \\cdot \\vec{J})~,\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nwith $B = \\rho \\partial p\/\\partial\\rho|_{n_{\\Delta},T} =\n \\rho c^2$\nthe appropriate isothermal bulk modulus in view of the fact that the pressure\nis a function of both the particle density as well as the net defect density.\n{}From thermodynamics we therefore also conclude that\n$\\gamma_{\\Delta} = \\partial p\/\\partial n_{\\Delta}|_{\\rho,T}$.\n\nIt is interesting to point out that these hydrodynamic equations differ from\nthe results obtained by Zippelius, Halperin, and Nelson. In particular, their\nEq.\\ (3.32) differs from our\nEq.\\ (\\ref{diff}) and reads\n\\begin{equation}\n\\vec{J} =\n - \\Gamma_{\\Delta} \\nabla\n \\left( \\frac{n_{\\Delta}}{\\chi_{\\Delta}} -\n \\gamma_{\\Delta} \\delta\\rho\n \\right)~.\n\\end{equation}\nThe difference can easily be traced back to the fact that Zippelius, Halperin,\nand Nelson assume on phenomenological grounds that the dynamics of the net\ndefect density is purely diffusive. Indeed, we exactly reproduce their results\nif we use in our calculation a free action of the form\n\\begin{equation}\nS_0[n_{\\Delta}] = \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\frac{E_{\\Delta}}{2} n_{\\Delta}\n \\left(\n \\frac{i\\partial_{\\tau}}{D_{\\Delta} \\partial^2} + 1\n \\right) n_{\\Delta}\n\\end{equation}\ninstead of Eq.\\ (\\ref{free}). We can therefore consider the hydrodynamic\nequations of Zippelius, Halperin and Nelson as the overdamped (or classical)\nlimit of our Eq.\\ (\\ref{HS}). Clearly, Kleinert's more microscopic approach\ndoes not lead to purely diffusive but in first instance to propagating behavior\nof\nthe defects, which is appropriate for the quantum crystals of interest in Sec.\\\n\\ref{SP}. We now turn to the modification of the above results in the hexatic\nphase.\n\nIn the hexatic phase there are free dislocations present and $\\chi''$ replaces\n$n_{\\Delta}$ as the appropriate dynamical degree of freedom. To see most\nclearly how this comes about we will work perturbatively in $1\/E_c$. Up to\nfirst order in $1\/E_c$\nthe effective action for $\\chi''$ is\n\\begin{eqnarray}\n\\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\left\\{ \\frac{1}{4\\mu(1+\\nu)} (\\partial^2 \\chi'')^2\n + \\frac{1}{2E_c} (\\partial_i \\chi'')\n \\left(\n \\frac{\\rho}{\\mu}\n \\frac{\\partial_{\\tau}^2}{\\partial^2} + 1\n \\right)^{-1} (\\partial_i \\chi'')\n \\right\\}~, \\nonumber\n\\end{eqnarray}\nwhich upon Fourier transformation displays two modes with\n$\\omega^2 = c_{\\perp}^2 \\vec{k}^2 + 2\\mu^2(1+\\nu)\/(E_c\\rho)$. So in the\nlimit $E_c \\rightarrow \\infty$ (which physically means that we are looking at\nthe nonhydrodynamic regime\n$\\vec{k}^2 \\gg \\mu\/2E_c$) we approximately have\n\\begin{equation}\n\\partial_{\\tau}^2 \\chi'' = - c_{\\perp}^2 \\partial^2 \\chi''~,\n\\end{equation}\nwhereas the equations of motion for $A^{TL}$ and $A^{TT}$ are\n\\begin{equation}\n\\partial_{\\tau}^2 A^{TL} =\n - c_{||}^2 \\nabla^2 A^{TL}\n + \\frac{\\nu}{1-\\nu^2} \\partial_{\\tau} \\chi''\n\\end{equation}\nand\n\\begin{equation}\n\\partial_{\\tau}^2 A^{TT} =\n - c_{\\perp}^2 \\nabla^2 A^{TT}~,\n\\end{equation}\nrespectively. For the mass-density fluctuation we now find\n\\begin{equation}\n\\delta\\rho = - \\frac{1}{c_{||}^2}\n \\partial_{\\tau} \\partial^2 A^{TL}\n - \\frac{\\rho}{2(\\mu + \\lambda)} \\partial^2 \\chi''\n\\end{equation}\nand for the stress tensor\n\\begin{equation}\n\\pi_{ij}^D = - \\delta_{ij}\n \\left\\{\n \\frac{1}{c_{||}^2} \\partial_{\\tau}^3 A^{TL}\n + \\frac{\\rho}{2(\\mu + \\lambda)} \\partial_{\\tau}^2 \\chi''\n \\right\\}\n = - \\delta_{ij}\n \\left\\{ c_{||}^2 \\delta\\rho\n + \\frac{1}{2(1+\\nu)} \\partial^2 \\chi''\n \\right\\}~.\n\\end{equation}\n\nPutting all this together we obtain in first instance the following set of\nhydrodynamic equations for the hexatic phase\n\\begin{mathletters}\n\\begin{equation}\n\\partial_{\\tau} \\delta\\rho = \\nabla \\cdot \\vec{g}^L~,\n\\end{equation}\n\\begin{equation}\n\\partial_{\\tau} \\vec{g}^L = - \\nabla \\delta p\n = - c_{||}^2 \\nabla \\delta\\rho\n - \\frac{1}{2(1+\\nu)} \\nabla (\\nabla^2 \\chi'')~,\n\\end{equation}\n\\begin{equation}\n\\partial_{\\tau}^2 \\chi'' = - c_{\\perp}^2 \\nabla^2 \\chi''~,\n\\end{equation}\n\\begin{equation}\n\\partial_{\\tau}^2 A^{TT} =\n - c_{\\perp}^2 \\nabla^2 A^{TT}~.\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nCombining the longitudinal and transverse parts as before, this equals\n\\begin{mathletters}\n\\begin{equation}\n\\partial_{\\tau} \\delta\\rho = \\nabla \\cdot \\vec{g}~,\n\\end{equation}\n\\begin{equation}\n\\label{mom}\n\\partial_{\\tau} \\vec{g} =\n - c_{||}^2 \\nabla \\delta\\rho\n - \\frac{1}{2(1+\\nu)} \\nabla (\\nabla^2 \\chi'')\n + 2\\rho c_{\\perp}^2 \\nabla \\times \\vartheta_6~,\n\\end{equation}\n\\begin{equation}\n\\partial_{\\tau}^2 \\chi'' = - c_{\\perp}^2 \\nabla^2 \\chi''~,\n\\end{equation}\n\\begin{equation}\n\\partial_{\\tau} \\vartheta_6 =\n - \\frac{1}{2\\rho} \\nabla \\times \\vec{g}~,\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nand clearly reduces to the hydrodynamic equations for the ideal crystal if we\nput $\\chi''=0$.\n\nWe now have to consider how the above picture changes for a finite value of\n$E_c$. Here we can use the results of\nSec.\\ \\ref{Hex}. In the hydrodynamic regime\n$\\vec{k}^2 \\ll \\mu\/2E_c$ we saw that the transverse speed of sound was\nrenormalized to zero, because we found the quadratic (particle-like) dispersion\n$\\omega^2 = 2E_c \\vec{k}^4\/\\rho$. As a result we have for the transverse\npart of the hydrodynamic equations\n\\begin{equation}\n\\partial_{\\tau}^2 A^{TT} =\n \\frac{2E_c}{\\rho} \\partial^4 A^{TT}~,\n\\end{equation}\nwhich implies that in the right-hand side of Eq.\\ (\\ref{mom}) we must replace\n$2\\rho c_{\\perp}^2 \\nabla \\times \\vartheta_6$ by\n$- 4 E_c \\vec{e}_z \\times \\nabla (\\nabla^2 \\vartheta_6)$. This gives\n\\begin{mathletters}\n\\begin{equation}\n\\partial_{\\tau} \\vec{g}^T =\n - 4 E_c \\vec{e}_z \\times \\nabla (\\nabla^2 \\vartheta_6)~,\n\\end{equation}\n\\begin{equation}\n\\partial_{\\tau} \\vartheta_6 =\n - \\frac{1}{2\\rho} \\nabla \\times \\vec{g}^T~,\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nwhich is in complete agreement with Zippelius, Halperin, and Nelson if we\nidentify the Frank constant $K_A$ with $8E_c$.\n\nFor the longitudinal part we need to analyze the dynamics of $\\chi''$ and\n$A^{TL}$. A straightforward calculation shows that the effective action for\nthese fields contains precisely the four modes already found in Sec.\\\n\\ref{Hex}. The propagating modes with $\\omega^2 = 2\\mu(1+\\nu)\n\\vec{k}^2\/\\rho$ obey\n$\\chi' = \\chi'' + \\nu \\partial_{\\tau} A^{TL} =0$ and are therefore indeed\nassociated with density fluctuations proportional to $\\partial^2 \\chi''$. We\nthus need to use a renormalized longitudinal speed of sound equal to\n\\begin{equation}\nc = \\sqrt{\\frac{2\\mu (1+\\nu)}{\\rho}}\n = \\sqrt{\\frac{2\\mu}{\\rho}\n \\frac{2\\mu + 2\\lambda}{2\\mu + \\lambda}}\n\\end{equation}\nthat is always smaller than the longitudinal speed of sound in the solid phase.\nIn fact, this actually exhausts the longitudinal hydrodynamic modes since the\nother modes in the effective action for $\\chi''$ and $A^{TL}$ are gapped. As a\nresult we now obtain in real time the following set of hydrodynamic equations\nfor the hexatic phase\n\\begin{mathletters}\n\\begin{equation}\n\\frac{\\partial \\delta\\rho}{\\partial t}\n = - \\nabla \\cdot \\vec{g}~,\n\\end{equation}\n\\begin{equation}\n\\label{momH}\n\\frac{\\partial \\vec{g}}{\\partial t} =\n - c^2 \\nabla \\delta\\rho\n - \\frac{K_A}{2}\n \\,\\,\\vec{e}_z \\times \\nabla (\\nabla^2 \\vartheta_6)~,\n\\end{equation}\n\\begin{equation}\n\\label{transH}\n\\frac{\\partial \\vartheta_6}{\\partial t} =\n \\frac{1}{2\\rho} \\nabla \\times \\vec{g}~,\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nnot including dissipation.\n\nTo include dissipation we again follow Zippelius, Halperin and Nelson and add\nto the right-hand side of Eq.\\ (\\ref{momH}) the terms $(\\eta \\nabla^2 \\vec{g}\n + \\zeta \\nabla (\\nabla \\cdot \\vec{g}) )\/\\rho$.\nHowever, we do not add the term $\\kappa \\nabla^2 \\vartheta_6$ to the right-hand\nside of Eq.\\ (\\ref{transH}) because, just as in the solid phase, the\ndissipation of the transverse modes is already accounted for in the term $\\eta\n\\nabla^2 \\vec{g}$ that is added to the momentum equation. Put differently, a\nterm of the form $\\kappa \\nabla^2 \\vartheta_6$ can be absorbed by an\nappropriate redefinition of $K_A$ and $\\eta$. Again introducing the isothermal\nbulk modulus\n$B = \\rho dp\/d\\rho|_T = \\rho c^2$ we then find\n\\begin{mathletters}\n\\label{HH}\n\\begin{equation}\n\\frac{\\partial \\delta\\rho}{\\partial t}\n = - \\nabla \\cdot \\vec{g}~,\n\\end{equation}\n\\begin{equation}\n\\frac{\\partial \\vec{g}}{\\partial t} =\n - \\frac{B}{\\rho} \\nabla \\delta\\rho\n - \\frac{K_A}{2}\n \\vec{e}_z \\times \\nabla (\\nabla^2 \\vartheta_6)\n + \\frac{\\eta}{\\rho} \\nabla^2 \\vec{g}\n + \\frac{\\zeta}{\\rho} \\nabla (\\nabla \\cdot \\vec{g}) ~,\n\\end{equation}\n\\begin{equation}\n\\label{theta}\n\\frac{\\partial \\vartheta_6}{\\partial t} =\n \\frac{1}{2\\rho} \\nabla \\times \\vec{g}\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nas our final result for the hexatic phase. Apart from the absence of a\ndissipative term in Eq.\\ (\\ref{theta}) it agrees with the findings of\nZippelius, Halperin and Nelson and therefore contains the same mode structure\nas derived in that paper. For completeness sake, we mention however that the\nequations of motion for the hexatic order parameter $\\vartheta_6$ can be\nderived from an effective action\n\\begin{equation}\nS^{eff}[\\vartheta_6] = \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\frac{1}{2} \\vartheta_6\n \\left( 4\\rho \\frac{\\partial_{\\tau}^2}{\\partial^2}\n + 4i \\eta \\partial_{\\tau} - K_A \\partial^2\n \\right) \\vartheta_6~,\n\\end{equation}\nthat can easily be understood physically: The first term on the right-hand side\ncorresponds to the kinetic energy\n$\\int d\\vec{x} \\rho (\\partial_{\\tau} \\vec{u})^2\/2$\nof the displacement field. The second term is a Leggett friction term and the\nlast term corresponds to the usual Frank energy, which is responsible for the\nfact that the hexatic to liquid transition is of the Kosterlitz-Thouless type.\n\n\\section{HYDRODYNAMICS OF SUPERFLUID PHASES}\n\\label{SP}\nHaving arrived at the hydrodynamic equations for the solid and hexatic phases,\nour next objective is to incorporate the effects of the additional hydrodynamic\ndegree of freedom associated with the phase of the superfluid order parameter.\nFortunately, from the microscopic theories developed for superfluid liquids\n\\cite{HM} and gases \\cite{KD} it is well known how we should proceed to obtain\nthe hydrodynamic (two-fluid) equations for the superfluid phases starting from\nthe equations for the normal phase. The procedure consists in principle of four\nsteps. First, the total (average) density $\\rho$ of the system is split up into\na normal density $\\rho_n$ and a superfluid density $\\rho_s$. In general these\ndensities are tensors of second rank, but for systems with hexagonal symmetry\nwhich are of interest here they are proportional to the identity $\\delta_{ij}$\nand can be considered as scalars. Second, the total momentum density $\\vec{g}$\nis similarly split up into a normal component\n$\\rho_n \\vec{v}_n$ and a superfluid component $\\rho_s \\vec{v}_s$ with a\nsuperfluid velocity that is purely longitudinal\n($\\nabla \\times \\vec{v}_s = 0$). Third, for an effectively isotropic system the\ndissipative terms in the momentum equation must be generalized to\n\\begin{eqnarray}\n\\eta \\nabla^2 \\vec{v}_n\n + \\zeta_1 \\frac{\\rho_s}{\\rho}\n \\nabla (\\nabla \\cdot (\\vec{v}_s - \\vec{v}_n))\n + \\zeta_2 \\nabla (\\nabla \\cdot \\vec{v}_n)~. \\nonumber\n\\end{eqnarray}\nFinally, we must add the dynamics of the superfluid velocity, which is\nbasically determined from the Josephson relation and reads\n\\begin{equation}\n\\frac{\\partial \\vec{v}_s}{\\partial t} =\n - \\frac{B}{\\rho^2} \\nabla \\delta\\rho\n + \\zeta_3 \\frac{\\rho_s}{\\rho}\n \\nabla (\\nabla \\cdot (\\vec{v}_s - \\vec{v}_n))\n + \\zeta_4 \\nabla (\\nabla \\cdot \\vec{v}_n)~,\n\\end{equation}\nwhere $B=\\rho^2 d\\mu\/d\\rho|_T$ is again the isothermal bulk modulus and\n$\\mu$ is the chemical potential per unit mass. We again leave out temperature\nfluctuations since we are primarily interested in third-sound modes, for which\nthese fluctuations are (at least qualitatively) unimportant.\n\n\\subsection{Supersolid}\n\\label{HSS}\nTo apply the above procedure to Eq.\\ (\\ref{HS}) we must realize that we are\nhere in fact already dealing with a two-fluid hydrodynamics. We must therefore\nnot only split up the total momentum density $\\vec{g}$ into a normal and a\nsuperfluid component but also the net defect current, i.e.\\\n$\\vec{J} = \\vec{J}_n + \\vec{J}_s$. Moreover, we have to account for the fact\nthat the chemical potential, just like the pressure, is a function of the\nparticle density and the net defect density. In this manner we arrive at the\nfollowing hydrodynamic equations\n\\begin{mathletters}\n\\begin{equation}\n\\frac{\\partial \\delta\\rho}{\\partial t} = -\\nabla \\cdot \\vec{g}~,\n\\end{equation}\n\\begin{eqnarray}\n\\label{momrho}\n\\frac{\\partial \\vec{g}}{\\partial t} =\n - \\frac{B}{\\rho} \\nabla \\delta\\rho\n &-& \\gamma_{\\Delta} \\nabla n_{\\Delta}\n + 2\\rho c_{\\perp}^2 \\nabla \\times \\vartheta_6\n \\nonumber \\\\\n &+& \\eta \\nabla^2 \\vec{v}_n\n + \\zeta_1 \\frac{\\rho_s}{\\rho}\n \\nabla (\\nabla \\cdot (\\vec{v}_s - \\vec{v}_n))\n + \\zeta_2 \\nabla (\\nabla \\cdot \\vec{v}_n)~,\n\\end{eqnarray}\n\\begin{equation}\n\\frac{\\partial \\vartheta_6}{\\partial t} =\n \\frac{1}{2\\rho} \\nabla \\times \\vec{g}~,\n\\end{equation}\n\\begin{equation}\n\\frac{\\partial \\vec{v}_s}{\\partial t} =\n - \\frac{B}{\\rho^2} \\nabla \\delta\\rho\n + \\beta_{\\Delta} \\nabla n_{\\Delta}\n + \\zeta_3 \\frac{\\rho_s}{\\rho}\n \\nabla (\\nabla \\cdot (\\vec{v}_s - \\vec{v}_n))\n + \\zeta_4 \\nabla (\\nabla \\cdot \\vec{v}_n)~,\n\\end{equation}\n\\begin{equation}\n\\frac{\\partial n_{\\Delta}}{\\partial t} =\n -\\nabla \\cdot \\vec{J}~,\n\\end{equation}\n\\begin{equation}\n\\label{momdel}\n\\frac{\\partial \\vec{J}}{\\partial t} =\n - c_{\\Delta}^2 \\nabla n_{\\Delta}\n + \\gamma \\nabla \\delta\\rho\n + \\kappa \\nabla^2 \\vec{J}_n\n + \\xi_1 \\nabla (\\nabla \\cdot \\vec{J}_s)\n + \\xi_2 \\nabla (\\nabla \\cdot \\vec{J}_n)~,\n\\end{equation}\n\\begin{equation}\n\\frac{\\partial \\vec{J}_s}{\\partial t} =\n - \\frac{B_{\\Delta} \\rho_s}{\\rho^2} \\nabla n_{\\Delta}\n + \\beta \\rho_s \\nabla \\delta\\rho\n + \\xi_3 \\frac{\\rho_s}{\\rho}\n \\nabla (\\nabla \\cdot \\vec{J}_s)\n + \\xi_4 \\nabla (\\nabla \\cdot \\vec{J}_n)~,\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nwith\n$\\beta_{\\Delta} = - \\partial \\mu\/\\partial\n n_{\\Delta}|_{\\rho,T}$.\nThese represent nine equations for the nine unknown functions $\\delta\\rho$,\n$\\vec{v}_n$, $\\vec{v}_s$, $\\vartheta_6$, $n_{\\Delta}$, $\\vec{J}_n$ and\n$\\vec{J}_s$.\n\nAlthough a complete analysis of the various hydrodynamic modes is now possible,\nwe will consider here only the situation which is most relevant to experiments,\nnamely that the normal part of the two-dimensional system is\nclamped to an underlying substrate. As a result we have\n$\\vec{v}_n = \\vec{J}_n = \\vec{0}$ and\nEqs.\\ (\\ref{momrho}) and (\\ref{momdel}) determining the normal properties of\nthe supersolid are no longer valid. The hydrodynamic equations therefore reduce\nto\n\\begin{mathletters}\n\\label{clamped}\n\\begin{equation}\n\\frac{\\partial^2 \\delta\\rho}{\\partial t^2} =\n \\frac{B \\rho_s}{\\rho^2} \\nabla^2 \\delta\\rho\n - \\beta_{\\Delta} \\rho_s \\nabla^2 n_{\\Delta}\n + \\zeta_3 \\frac{\\rho_s}{\\rho}\n \\frac{\\partial}{\\partial t}(\\nabla^2 \\delta\\rho)~,\n\\end{equation}\n\\begin{equation}\n\\frac{\\partial^2 n_{\\Delta}}{\\partial t^2} =\n \\frac{B_{\\Delta} \\rho_s}{\\rho^2} \\nabla^2 n_{\\Delta}\n - \\beta \\rho_s \\nabla^2 \\delta\\rho\n + \\xi_3 \\frac{\\rho_s}{\\rho}\n \\frac{\\partial}{\\partial t} (\\nabla^2 n_{\\Delta})~.\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nThey contain two pairs of\npropagating modes, which in the limit of a small coupling\nconstant\n$\\beta \\ll BB_{\\Delta}\/\\beta_{\\Delta} \\rho^4$ essentially correspond to\na pair of\nthird-sound modes with $\\delta\\rho$ unequal to zero but a constant net defect\ndensity and a pair of modes with an oscillating net defect density.\n\nOne might have expected that the coupling of a superfluid\ndensity to a propagating defect density would have resulted\nin one pair of gapped excitations and one pair of gapless excitations\ninstead. Consider, for example, two identical superfluid layers.\nIf the layers are uncoupled the dynamics of the\nphases $\\vartheta_1$ and $\\vartheta_2$ of the layers is determined by the\naction\n\\begin{equation}\nS_{layers}[\\vartheta_1,\\vartheta_2] =\n \\int_0^{\\hbar \\beta} d\\tau \\int d\\vec{x}~\n \\left\\{ \\frac{\\rho^2}{2B} (\\partial_{\\tau} \\vartheta_1)^2 +\n \\frac{\\rho_s}{2} (\\nabla \\vartheta_1)^2\n + \\frac{\\rho^2}{2B} (\\partial_{\\tau} \\vartheta_2)^2 +\n \\frac{\\rho_s}{2} (\\nabla \\vartheta_2)^2\n \\right\\}~,\n\\end{equation}\nwhich clearly has two pairs of\ngapless (third-sound) modes, one pair for each superfluid.\nIf we couple the order parameters by allowing the particles\nto tunnel with an amplitude $-J\/\\rho$\nfrom one layer to the other we must add\na Josephson coupling\n\\begin{equation}\nS_{tunnel}[\\vartheta_1,\\vartheta_2] = - \\int d\\tau \\int d\\vec{x}~\n J \\cos(\\vartheta_1 - \\vartheta_2)\n\\end{equation}\nto this action. The hydrodynamics modes couple to form two in-phase and\ntwo out-of-phase excitations. The\nmodes with $\\vartheta_1$ and $\\vartheta_2$ oscillating out of phase get\ngapped (i.e.\\\n$\\omega^2 \\simeq BJ\/\\rho^2$ for $\\vec{k}^2 \\ll J\/\\rho_s$) and only the\nmodes with $\\vartheta_1$ and $\\vartheta_2$ oscillating in phase remain gapless.\nYet in Eq.\\ (\\ref{clamped}) we find only gapless modes.\n\nThis paradox can be resolved by noting that\nwe have made the standard assumption\\cite{ZHN} that both the\ntotal number of particles and the net number of defects is conserved. Hence,\nafter an atom has tunneled from a lattice site to the position of a vacancy, a\nnew vacancy is created near the original site of the atom.\nThe analogous process for the\ntwo coupled superfluid layers in not simply tunneling of\nindividual atoms from one layer to another, but rather the\nexchange of a pair of atoms in different layers, returning the\nsystem to its original state. Such a process is not a\nJosephson coupling and therefore the modes remain gapless.\nThe existence of separate conservation laws for the particle and defect\ndensity thus allows in principle two separate broken symmetries.\n\nWe also note in passing\nthat the third-sound modes in Eq.\\ (\\ref{clamped}) are not present\nin the hydrodynamic equations proposed by Andreev and Lifshitz \\cite{A} and\nconsidered in more detail by Liu. \\cite{L} This is a result of the fact that\nthese authors use a somewhat different physical picture for the supersolid\nphase: They assume that the superfluid current density is carried by (Bose\ncondensed) defects and that the normal current density is solely due to lattice\nvibrations. Hence if we take\n$\\vec{v}_n = \\vec{0}$, which in their context means that $\\partial\n\\vec{u}\/\\partial t = \\vec{0}$, only transport of defects is possible and only\nthe latter two modes survive. However, as a consequence of their picture the\nhydrodynamic equations in the (normal) solid phase describe only longitudinal\nand transverse sound modes in an ideal lattice and do not include the effect of\nvacancies or interstitials. As explained above this is incorrect in principle\nand one should at least also allow for a normal current density due to the\nmotion of defects. In addition, we have seen in Sec.\\ \\ref{SOP} that even in\nthe presence of defects the density fluctuations are equal to $-\\rho\nu^{Phys}_{ii}$. It is therefore perfectly reasonable that if there is\nsuperfluid mass transport possible in the solid, it can be caused both by the\nmotion of defects and by lattice vibrations. Indeed, as an existence proof of\nthis latter possibility we can for instance consider superfluid $^4$He in a\nweak periodic and commensurate potential, which is clearly a supersolid without\ndefects.\n\nWhile it is generically possible to have both density and\ndefect superfluid modes, we might expect however, for realistic films on\nrealistic substrates, that in a supersolid it may be harder for particles to\nperform ring exchanges\\cite{K2,ring} than for vacancies\nto exchange positions. Thus, {\\it a priori}, we might expect the effective\nsuperfluid stiffness for the density fluctuations to be smaller than that\nof the vacancies, perhaps to the point where the former is entirely absent.\n\n\\subsection{Superhexatic}\nWe next turn to the superhexatic phase. In a similar manner as in Sec.\\\n\\ref{HSS} we obtain from Eq.\\ (\\ref{HH}) the full set of hydrodynamic equations\n\\begin{mathletters}\n\\begin{equation}\n\\frac{\\partial \\delta\\rho}{\\partial t}\n = - \\nabla \\cdot \\vec{g}~,\n\\end{equation}\n\\begin{eqnarray}\n\\frac{\\partial \\vec{g}}{\\partial t} =\n - \\frac{B}{\\rho} \\nabla \\delta\\rho\n &-& \\frac{K_A}{2}\n \\vec{e}_z \\times \\nabla (\\nabla^2 \\vartheta_6)\n \\nonumber \\\\\n &+& \\eta \\nabla^2 \\vec{v}_n\n + \\zeta_1 \\frac{\\rho_s}{\\rho}\n \\nabla (\\nabla \\cdot (\\vec{v}_s - \\vec{v}_n))\n + \\zeta_2 \\nabla (\\nabla \\cdot \\vec{v}_n)~,\n\\end{eqnarray}\n\\begin{equation}\n\\label{theta6}\n\\frac{\\partial \\vartheta_6}{\\partial t} =\n \\frac{1}{2\\rho} \\nabla \\times \\vec{g}~,\n\\end{equation}\n\\begin{equation}\n\\frac{\\partial \\vec{v}_s}{\\partial t} =\n - \\frac{B}{\\rho^2} \\nabla \\delta\\rho\n + \\zeta_3 \\frac{\\rho_s}{\\rho}\n \\nabla (\\nabla \\cdot (\\vec{v}_s - \\vec{v}_n))\n + \\zeta_4 \\nabla (\\nabla \\cdot \\vec{v}_n)~,\n\\end{equation}\n\\end{mathletters}\n\n\\noindent\nthat leads to the usual two-fluid hydrodynamics of a superfluid if we omit Eq.\\\n(\\ref{theta6}) and put $\\vartheta_6 = 0$. Therefore these equations allow for\nfirst and second sound, \\cite{second} and for a pair of transverse modes\ninvolving $\\vec{v}_n^T$ and $\\vartheta_6$ which are either dispersive or\npropagating depending on the sign of\n$\\Delta = K_A\/4\\rho - (\\eta\/\\rho_n)^2$: If $\\Delta \\leq 0$ we have two\npurely dispersive modes with\n$\\omega^{\\pm} = -i(\\eta\/\\rho_n \\pm \\sqrt{-\\Delta}) \\vec{k}^2\/2$,\nwheras if $\\Delta > 0$ we have two propagating modes and the particle-like\ndispersion\n$\\omega^{\\pm} = \\pm \\sqrt{\\Delta}~ \\vec{k}^2\/2\n - i (\\eta\/\\rho_n) \\vec{k}^2\/2$.\nHowever, considering again the case $\\vec{v}_n = \\vec{0}$ the hydrodynamic\nequations now simply reduce to\n\\begin{equation}\n\\frac{\\partial^2 \\delta\\rho}{\\partial t^2} =\n \\frac{B \\rho_s}{\\rho^2} \\nabla^2 \\delta\\rho\n + \\zeta_3 \\frac{\\rho_s}{\\rho}\n \\frac{\\partial}{\\partial t}(\\nabla^2 \\delta\\rho)~,\n\\end{equation}\nwhich contains only a pair of third-sound modes with the velocity\n$c_3 = \\sqrt{B\\rho_s\/\\rho^2}$ and the diffusion constant\n$D_3 = \\zeta_3 \\rho_s\/\\rho$.\n\n\\section{CONCLUSIONS AND DISCUSSION}\n\\label{DC}\nIn this paper we have derived the hydrodynamic equations for the supersolid\nand superhexatic phases of a neutral two-dimensional Bose fluid. For the\nsupersolid these equations are rather complex, since they incorporate the\neffects of defect motion and lattice vibrations on both the normal and\nsuperfluid parts of the momentum density. Our physical picture for the\ninfluence on the superfluid part is roughly speaking that in a mean-field\ntheory the condensate wavefunction $\\Psi(\\vec{x},t)$ obeys the Schr\\\"odinger\nequation\n\\begin{equation}\ni\\hbar \\frac{\\partial \\Psi(\\vec{x},t)}{\\partial t} =\n \\left\\{\n - \\frac{\\hbar^2 \\nabla^2}{2m}\n + \\int d\\vec{x}'~ V(\\vec{x}-\\vec{x}') n(\\vec{x}',t)\n \\right\\} \\Psi(\\vec{x},t)~,\n\\end{equation}\nwhere $m$ is the mass of the Bose particles and $V(\\vec{x}-\\vec{x}')$ is their\ninteraction. In addition, $n(\\vec{x},t)$ is the particle density which will be\ndetermined by an additional mean-field theory that, for a supersolid, shows the\ninstability associated with the formation of a density wave. Hence the\n(thermal) average\n$\\langle n(\\vec{x},t) \\rangle$ is periodic in space and independent of time. As\na result the condensate wavefunction is, if we neglect density fluctuations,\nalso periodic and we have indeed both diagonal as well as off-diagonal\nlong-range order. Fluctuations in the density, however, induce variations in\nthe phase of the wavefunction and therefore in the superfluid velocity. Because\nthese density fluctuations can be caused by both lattice vibrations and\noscillations in the net defect density we conclude that both mechanisms can\nlead to superfluid motion. Together with the existence of a conservation law\nfor the net number of defects, this explains from a more microscopic view\nwhy we found two third-sound modes and two modes with an oscillatory net defect\ndensity in the case of a supersolid adsorbed onto a substrate.\n\nFor the superhexatic phase we have shown that the hexatic long-range order\nleads to an additional (as compared to the superfluid) hydrodynamic degree of\nfreedom that affects only the transverse modes and is therefore at long\nwavelengths decoupled from the superfluid momentum density. This can also be\nunderstood from the above picture, since variations in the orientational order\nparameter $\\vartheta_6$ do not lead to density fluctuations in first instance.\nAs a result we find on a substrate only two third-sound modes and thus at the\nhydrodynamic level of description nothing to distinguish the superhexatic from\nthe superfluid. Although this is in agreement with the experiments of Chen and\nMochel, who indeed only observe one third-sound branch below the second\ncritical temperature $T_c$, it is unfortunate for the purpose of suggesting a\npossible identification of the superhexatic phase. On the basis of our results\nwe can, however, conclude that a more microscopic probe is needed if one wants\nto detect the orientational order present in a superhexatic helium film. In our\nopinion this appears to be an important, but also difficult experimental\nchallenge.\n\nFinally, we would also like to point out the possible relevance of our results\nto the recent experiments with bulk solid $^4$He. \\cite{LG} In these\nexperiments Lengua and Goodkind observe at sufficiently high frequencies an\nadditional (resonant) attenuation and velocity change of sound. Moreover, they\nnotice that their data can be explained by a simple model of two coupled wave\nequations which turns out to be identical to the longitudonal part of our\nsolid hydrodynamics derived in\nSec.\\ \\ref{SOP}. Because our two-dimensional hydrodynamics should be able to\ndescribe the propagation of sound perpendicular to the c-axis of hcp $^4$He,\nthis confirms the conjecture of Lengua and Goodkind that the collective mode\nobserved is associated with the motion of defects. For a more detailed\ndiscussion of the coupling between sound and the defects one should of course\nconsider the fully three-dimensional situation and include the anisotropy of\nthe hcp crystal. Work in this direction is in progress.\n\n\\section*{ACKNOWLEDGMENTS}\nThis research was supported by Grant No.\\ DMR-9502555 and DMR-9416906\nfrom the National\nScience Foundation, the ESF Network on Quantum Fluids and Solids,\nthe Swedish Natural Science Research Council and the\nStichting voor Fundamenteel Onderzoek der Materie (FOM) which is\nfinancially supported by the Nederlandse Organisatie voor\nWetenschappelijk Onderzoek (NWO).\nWe thank Huug van Beelen, Henk van Beijeren, Michel Bijlsma,\nReyer Jochemsen, and Anne van Otterlo for stimulating and helpful discussions.\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\nThe (Borel-Moore) homology of a smooth manifold possesses a canonical ring structure in which the product is represented by intersection of cycle classes.\nThe Chow groups of a smooth algebraic variety over the complex numbers also admit a ring structure, and the intersection products are compatible via the cycle class map \\cite[Cor.~19.2]{Ful}. In both contexts, a well-behaved intersection theory for cycles fails to extend to spaces with singularities.\n\nIn topology, this motivated the definition of cohomology and its cup product operation, which in some sense isolates a subset of the space of cycles (with a different equivalence relation) which can be intersected even if the ambient space has singularities. There are several analogues in algebraic geometry.\nThe Friedlander-Lawson theory of algebraic cocycles \\cite{FLcocycle} is a geometric approach built from finite correspondences to projective spaces, and similar constructions underlie the motivic cohomology of Friedlander-Voevodsky \\cite {FV}. \nThe operational Chow cohomology of Fulton \\cite{Ful}, on the other hand, adopts a more formal approach. \n\n\nThe intersection homology groups of Goresky and MacPherson \\cite{GM1} provide an\ninterpolation between cohomology and homology: at least for a normal space $X$ there is a sequence of groups:\n$$H^{\\dim X -*}(X) = IH^{\\overline 0}_* (X) \\to \\cdots \\to IH^{\\overline p}_*(X) \\to \\cdots \\to IH^{\\overline t}_*(X) = H_*(X)$$\nfactoring the cap product map. The decoration $\\overline p$, called the perversity, is a sequence of integers which prescribes how cycles may meet the strata in a suitable stratification of the (possibly singular) space $X$; in the display above it increases from left to right. Each intersection homology group $IH^{\\overline p}_r(X)$ arises as the homology of a complex of chains (either simplicial chains with respect to a triangulation \\cite{GM1}, or singular chains \\cite{King}) of perversity $\\overline p$.\n\nOne of the most interesting features of this theory is the existence of intersection pairings\n$$IH^{\\overline p}_r(X) \\otimes IH^{\\overline q}_s(X) \\to IH^{\\overline p + \\overline q}_{r+s - \\dim(X)}(X)$$\n(provided $\\overline p + \\overline q \\leq \\overline t$) generalizing the cap product pairing between cohomology and homology, and providing a generalization of Poincar\\'e duality to singular spaces.\n\nThe author and Eric Friedlander \\cite{intsing} have defined an algebraic cycle counterpart to the geometric approach of Goresky-MacPherson. In particular we have defined perverse Borel-Moore motivic homology groups $H^{\\overline p}_{m}(X, \\mathbb{Z}(r))$ (for a stratified variety $X$ and a perversity $\\overline p$) with a cycle class map to the Goresky-MacPherson theory in Chow degree: $H^{\\overline p}_{2r}(X, \\mathbb{Z}(r)) \\to IH^{\\overline p}_{2r}(X, \\mathbb{Z})$ \\cite[Defn.~2.3, Prop.~2.5]{intsing}. We have constructed also a perverse variation of motivic cohomology $H^{i,s, \\overline p}(X)$ and pairings \\cite[Defn.~5.3, Prop.~6.13]{intsing}:\n$$\\cap : H^{i,s, \\overline p}(X) \\otimes H^{\\overline q}_m (X, \\mathbb{Z}(r) ) \\to H^{\\overline p + \\overline q}_{m-i}(X, \\mathbb{Z}(r-s)) .$$\nThere is a canonical morphism $H^{i, s, \\overline p}(X) \\to H^{\\overline p}_{2 d -i}(X, \\mathbb{Z}(d -s))$ \\cite[Cor.~6.14]{intsing} with $d = \\dim (X)$. To define an intersection product\n$$H^{\\overline p}_{m}(X, \\mathbb{Z}(r)) \\otimes H^{\\overline q}_{n}(X, \\mathbb{Z}(s)) \\to H^{\\overline p + \\overline q}_{m+n - 2 d}(X, \\mathbb{Z}(r+s - d))$$\n(i.e., extend $\\cap$), one needs to involve the stratification in a more substantial way.\n\nIn this paper we approach the problem of defining such a product in Chow degree ($m=2r, n = 2s$) by considering the following (somewhat vague) question: if a singular variety $X$ can be resolved by a smooth variety $\\widetilde X$, to what extent does the intersection product on the Chow groups of $\\widetilde X$ provide a sensible intersection theory for algebraic cycles on $X$? As one can already see from the case of a proper birational morphism $\\pi : \\widetilde X \\to X$ between smooth varieties, the push-forward of the intersection formed on $\\widetilde X$ is in general different from the intersection formed on $X$.\nOn the other hand, one also sees that for a blowup $\\widetilde X \\to X$ of a smooth variety $X$ along a smooth subvariety, the push-forward of the intersection formed on $\\widetilde X$ agrees with the intersection formed on $X$ if the cycles have controlled incidence with $Y$; see Proposition \\ref{smooth case} for a precise statement.\n\nOur main results establish cases in which intersections formed on a resolution provide well-defined products on modifications of Chow groups: instead of considering cycles modulo rational equivalence, certain strata are singled out by the resolution, and both the cycles and the equivalences among them are required to have controlled incidence with the strata. If $k$ is a field of characteristic zero, then any $k$-variety $X$ may be resolved by a sequence of blowups along smooth centers $\\pi : \\widetilde X \\to X$. From such a resolution we define a stratification of $X$ and construct an intersection pairing on perverse Chow groups (i.e., a map $A_{r, \\overline p} (X) \\otimes A_{s, \\overline q}(X) \\to A_{r+s-d}(X)$, possibly with $\\mathbb{Q}$ coefficients) in the following settings:\n\\begin{itemize}\n\\item $s=\\dim(X)-1$, i.e., one of the cycles is a divisor (Theorem \\ref{divisor pairing});\n\\item $r=s=2$ and $\\dim(X)=4$, i.e., there is an intersection pairing on 2-cycles on a fourfold (Theorem \\ref{4fold}); and\n\\item $\\dim(X_{sing})=1$ and $r,s$ arbitrary (Theorem \\ref{one dim sing}).\n\\end{itemize}\nThe pairing is obtained by pushing forward the intersection product of the proper transforms, i.e., is given by $\\alpha, \\beta \\mapsto \\pi_* (\\widetilde \\alpha \\cdot \\widetilde \\beta)$. The stratification is expressed in terms of the geometry of the resolution, and our arguments are accordingly geometric in nature. In all three cases, the basic idea is to understand how (certain) rational equivalences behave under proper transform. For concreteness we explain the idea for 2-cycles $\\alpha, \\beta$ of complementary perversities $\\overline{p}, \\overline{q}$ on a fourfold $X$. If one has $\\alpha \\sim_{\\overline p} \\alpha'$ on $X$, then $e_\\alpha := \\widetilde \\alpha - \\widetilde {\\alpha'}$ is a cycle on $\\widetilde X$ supported over the singular locus of $X$, with the support controlled by the perversity function $\\overline p$.\nThe perversity $\\overline{p}$ provides enough control over the error term $e_\\alpha$ that we can find a cycle which is both rationally equivalent to some multiple of $e_\\alpha$ and, using the complementarity of the perversity condition $\\overline q$, disjoint from $\\widetilde{\\beta}$ (or $e_\\beta$).\nThen, at least with rational coefficients, we have the vanishings $e_\\alpha \\cdot \\widetilde \\beta = \\widetilde \\alpha \\cdot e_\\beta = e_\\alpha \\cdot e_\\beta = 0$ in $A_0(\\widetilde X)$, so that $\\pi_* (\\widetilde \\alpha \\cdot \\widetilde \\beta )$ and $\\pi_*(\\widetilde \\alpha' \\cdot \\widetilde \\beta')$ coincide as classes in $A_0(X)_\\mathbb{Q}$.\nFor divisors the arguments are more conceptual and less technical, but the main point is to move error terms arising from (certain) rational equivalences away from cycles of complementary perversity. When the singular locus of $X$ is one-dimensional, we also employ intersection theory on the exceptional components and subvarieties therein.\n\n\nFor certain pairs of cycles, our product agrees with the one defined by Goresky-MacPherson via the cycle class map, but we do not know if this holds in general. Another basic question is the dependence on the resolution. See \\S \\ref{sec:further} for further discussion of these questions.\n\nIn \\S \\ref{sec:cycles} we establish background and notation. In particular, we discuss various procedures for constructing stratifications from resolutions; later we point out which constructions are necessary in the different contexts. The next three sections are devoted to the three situations mentioned above: in \\S \\ref{sec:divisors} we show the resolution provides a sensible intersection theory for divisors, in \\S \\ref{sec:onedimsing} we handle the case $\\dim(X_{sing}) =1$, and in \\S \\ref{sec:4fold} we study 2-cycles on a fourfold. In \\S \\ref{need integral fibers} we show, through an explicit example, that an ``obvious\" coarsening of our stratification, namely by the fiber dimension of the resolution, is insufficiently fine for the resolution to provide a decent intersection product.\n\n\\medskip\n\\textbf{Conventions.} Throughout we work with schemes separated and of finite type over a field $k$. A variety is an integral $k$-scheme.\n\n\\medskip\n\\textbf{Acknowledgments.} \nThe idea that resolutions might provide pairings on perverse cycle groups emerged in conversations with Eric Friedlander.\nThe author wishes to thank him for his encouragement and interest in this project. The author was partially supported by National Science Foundation Award DMS-0966589.\n\n\n\\section{Cycles, equivalence relations, and stratifications from resolutions} \\label{sec:cycles}\n\nIn this section we adapt some of the basic features of Goresky-MacPherson intersection homology to our algebraic context. This reproduces some material from \\cite[\\S 2]{intsing}. Then we discuss stratifications obtained from resolutions. Finally we demonstrate that our proposal for intersections on a singular variety recovers the usual intersection theory for the blowup of a smooth variety along a smooth subvariety.\n\nA \\textit{stratified variety} is a variety $X$ (say of dimension $d$) equipped with a filtration by closed subsets $X^d \\hookrightarrow X^{d-1} \\hookrightarrow \\cdots \\hookrightarrow X^2 \\hookrightarrow X^1 \\hookrightarrow X$ such that $X^i$ has codimension at least $i$ in $X$. This is usually called a filtered space, which determines a stratified space; since we do not deal with more general stratifications, we ignore the difference.\nA \\textit{perversity} is a non-decreasing sequence of integers $p_1, p_2, \\ldots, p_d$ such that $p_1=0$ and, for all $i$, $p_{i+1}$ equals either $p_i$ or $p_i+1$. Perversities are denoted $\\overline{p}, \\overline{q}$, etc. The perversities we consider range from the zero perversity $\\overline{0}$ with $p_i=0$ for all $i$, to the top perversity $\\overline{t}$, with $p_i=i-1$ for all $i$.\n\nLet $Z_r(X)$ denote the group of $r$-dimensional algebraic cycles on $X$. If $X$ is stratified and $\\overline p$ is a perversity, we say an $r$-cycle $\\alpha$ is \\textit{of perversity $\\overline{p}$} (or satisfies the perversity condition $\\overline{p}$) if for all $i$, the dimension of the intersection $|\\alpha| \\cap X^i$ is no larger than $r-i +p_i$. If the codimension of $X^i$ in $X$ is exactly $i$, then the perversity of a cycle measures its failure to meet properly the closed sets occurring in the stratification of $X$. Let $Z_{r, \\overline{p}}(X) \\subset Z_r(X)$ denote the group of $r$-dimensional cycles of perversity $\\overline{p}$ on the stratified variety $X$. Typically $X^1$ is the singular locus of $X$, and then the condition $p_1=0$ means that no component of the cycle is contained in the singular locus.\n\nIn \\cite[Prop.~2.4]{intsing} we defined and characterized a notion of $\\overline p$-rational equivalence of algebraic $r$-cycles: we identify two elements of $Z_{r, \\overline p}(X)$ if they can be connected by an $\\mathbb{A}^1$-family of $r$-cycles of perversity $\\overline p$. Here we introduce a modification.\nWe say the cycles $\\alpha, \\alpha' \\in Z_{r, \\overline{p}}(X)$ are \\textit{weakly rationally equivalent as perversity $\\overline{p}$ cycles} (or are weakly $\\overline{p}$-rationally equivalent), written $\\alpha \\sim_{w, \\overline{p}} \\alpha'$, if there is an equation of rational equivalence all of whose terms satisfy the perversity condition $\\overline{p}$. Explicitly this means there exist subvarieties $W_1, \\ldots, W_a$ of dimension $r+1$ in $X$, and rational functions $g_i : W_i \\dashrightarrow \\bb{P}^1$, such that $\\alpha - \\alpha' = \\sum_i [g_i(0)] - [g_i(\\infty)]$ in $Z_{r, \\overline{p}}(X)$ (i.e., for all $i$, both $[g_i(0)]$ and $[g_i(\\infty)]$ are in $Z_{r, \\overline{p}}(X)$). \n\nIn contrast to the definition of $A_{r, \\overline p}(X)$, here we do \\textit{not} require that for every $t \\in \\mathbb{A}^1 \\subset \\bb{P}^1$, the cycle associated to the fiber $[g_i(t)]$ satisfies the perversity condition. We allow the $g_i$'s to have ``bad\" fibers, but we require that $\\alpha$ and $\\alpha'$ are related by ``good\" fibers. We could equivalently require that the condition of \\cite[Prop.~2.4(3)]{intsing} is satisfied with $\\mathbb{A}^1$ replaced by an open subset (containing at least two $k$-points) therein.\nWe let $A^w_{r, \\overline{p}}(X)$ denote the group of $r$-cycles of perversity $\\overline{p}$ modulo weak $\\overline{p}$-rational equivalence. \nThe main reason for introducing another equivalence relation is that our arguments ``naturally\" construct pairings on the groups $A^w_{r, \\overline p}(X)$.\nSince there is a canonical morphism $A_{r, \\overline{p}}(X) \\to A^w_{r, \\overline p}(X)$, we immediately obtain pairings on the groups $A_{r, \\overline p}(X)$. Note also the map $A_{r, \\overline p}(X) \\to IH^{\\overline p}_{2r}(X, \\mathbb{Z})$ factors through $A^w_{r, \\overline p}(X)$: the proof of \\cite[Prop.~2.5]{intsing} only requires $\\mathcal W \\hookrightarrow X \\times \\bb{P}^1$ to have good fibers over $0, \\infty \\in \\bb{P}^1$, not all $t \\in \\bb{P}^1$.\n\n\n\\medskip\n\n\\textbf{Resolutions.} Let $X$ be a variety over a field $k$ of characteristic zero. The celebrated result of Hironaka asserts the singularities of $X$ can be resolved by a sequence of blowups along smooth centers which are normally flat in their ambient spaces, i.e., the resolution can be expressed as a composition $\\pi: \\widetilde{X} = X_f \\to \\cdots \\to X_{n+1} \\to X_n \\to \\cdots \\to X = X_0$ where (for all $n = 0, \\ldots, f-1$) $X_{n+1} = \\Bl_{C_n} X_n$, $C_n \\hookrightarrow X_n$ is smooth, and $X_n$ is normally flat along $C_n$ \\cite{Hironaka}. We will call a resolution admitting such an expression a \\textit{strong resolution.}\n\n\\begin{definition} The exceptional locus $\\widetilde{E} \\hookrightarrow \\widetilde{X}$ of the resolution is simply the preimage of the singular locus of $X$. We say a subvariety $V \\hookrightarrow \\widetilde{X}$ is \\textit{exceptional} if it is contained in the exceptional locus of the resolution. We say an exceptional subvariety $V \\hookrightarrow \\widetilde{X}$ \\textit{first appears} on $X_{n+1} = \\Bl_{C_n}X_n$ if there is a subvariety $\\overline{V} \\hookrightarrow X_{n+1}$ such that the morphism $V \\hookrightarrow \\widetilde{X} \\to X_{n+1}$ factors through a birational morphism $V \\to \\overline{V}$, and there is no such subvariety of $X_n$. The concept of first appearance depends on the sequence of centers used to construct the resolution. \\end{definition}\n\n\\begin{notation} If $C_n \\hookrightarrow X_n$ is a center occurring in the resolution, let $\\widetilde{E}_{C_n} \\subseteq \\widetilde{E}$ denote the union of those exceptional divisors $E$ for which the canonical morphism $E \\to X$ factors through $C_n$, and let $\\widetilde{E}_{\\text{dom }C_n} \\subseteq \\widetilde{E}_{C_n}$ denote the union of those divisors for which the canonical morphism $E \\to C_n$ is dominant. \\end{notation}\n\n\\begin{definition}[stratification via resolution] \\label{strat res defn} \nWe describe several ways a center $C_n$ may contribute to a stratification of the variety being resolved.\nFirst we consider what happens ``below\" a center $C_n$. \n\\begin{BC}\n\\item If $C_n \\to X$ has generic dimension $g$ over its $m$-dimensional image $W \\hookrightarrow X$, then $W \\hookrightarrow X^{d-m}$; and for each $e \\geq 1$ the locus $W' \\hookrightarrow W$ over which $C_n \\to W$ has fiber dimension $\\geq g+e$ is placed in $X^{d-m+e+1}$. \n\\item If $g \\geq 1$, the generic fiber of $C_n \\to W$ is integral; the locus over which $C_n \\to W$ fails to have an integral fiber of dimension $g$ is placed in $X^{d-m+1}$.\n\\item The singularities of $W$ are placed in $X^{d-m+1}$.\n\\end{BC}\nNext we consider what happens ``above\" a center.\n\\begin{AC}\n\\item For a general point $c \\in C_n$, the number of irreducible components of ${(\\widetilde{E}_{\\text{dom }C_n})}_c$ is equal to the number of components of $\\widetilde{E}_{\\text{dom }C_n}$; where the former exceeds the latter is a closed set denoted $R_n \\hookrightarrow C_n$. \nIf the image $V$ of $R_n$ in $W$ is not dense, we place this image in $X^{d-m+e}$, where $e$ is the codimension of $V$ in $W$ (hence $d-m+e$ is the codimension of $V$ in $X$). \n\\end{AC}\nLet $\\pi : \\widetilde X \\to X$ be a resolution expressed as a sequence of blowups along $C_n \\hookrightarrow X_n$. For any of the four constructions listed above, we say a stratification of $X$ \\textit{satisfies condition (C) with respect to $\\pi$} if it refines the stratification obtained by applying construction (C) to all of the centers $C_n$ occurring in the resolution. \\end{definition}\n\n\\begin{remark} Let $\\pi : \\widetilde X \\to X$ be a resolution. If a stratification satisfies (BC-1) with respect to $\\pi$, then it refines the ``fiber dimension\" stratification given by specifying $X^i$ is the closed subset along which the fibers of $\\pi$ have dimension at least $i-1$.\n\nA stratification satisfying (BC-1), (BC-2), (BC-3), and (AC) may not have the property that $X^i \\setminus X^{i+1}$ is smooth since we do not, for example, necessarily place the incidences of the components of $X_{sing}$ in a smaller stratum. For an explicit example in which the incidences of the components of the singular locus are not detected by the conditions above, see the projective example of Section \\ref{need integral fibers}.\n\nIt would be very interesting to characterize intrinsically the strata which arise from the conditions described above.\n\\end{remark}\n\n\n\\medskip\n\\textbf{Error terms.} Suppose $\\alpha \\sim_{w, \\overline{p}} \\alpha'$ and \\begin{equation} \\label{alpha equiv} \\alpha - \\alpha' = \\sum_{f_i \\in k(S_i)} [f_i(0)] - [f_i(\\infty)] \\end{equation}\nis an equation of rational equivalence in $Z_{r, \\overline{p}}(X)$. Viewing the $f_i$'s as rational functions on the proper transforms of the $S_i$'s, we obtain an equation of rational equivalence $\\widetilde{\\alpha} + e_\\alpha - \\widetilde{\\alpha'} - e_{\\alpha'} = \\sum_{f_i \\in k(\\widetilde{S_i})} [f_i(0)] - [f_i(\\infty)]$ in $Z_r(\\widetilde{X})$, where $e_\\alpha$ is exceptional for some $\\widetilde{S_i} \\to S_i$ and is supported over the image of the exceptional locus of $\\widetilde{z} \\to z$ for $z$ some term in the equation $\\ref{alpha equiv}$ (and similarly for $e_{\\alpha'}$). In this situation we refer to $e_\\alpha$ and $e_{\\alpha'}$ as the ``error terms.\"\nWe write $e(\\alpha, \\alpha')$ for the union of the supports of the error terms, and typically $e_\\alpha$ will denote a component of the support of $e(\\alpha, \\alpha')$.\n\n\\medskip\n\\textbf{Normality.} There is no loss of generality in assuming $X$ is normal. For if $X$ is not necessarily normal and $\\nu : X^\\nu \\to X$ is the normalization, and $\\pi : \\widetilde{X^\\nu} \\to X^\\nu$ is the result of applying a resolution algorithm to $X^\\nu$, we simply declare the stratification of $X$ to be the image via $\\nu$ of the stratification of $X^\\nu$ (induced by $\\pi$), augmented by declaring $X^1$ is the singular locus of $X$. Note that $\\nu \\circ \\pi$ is probably not the result of applying a resolution algorithm to $X$ itself.\n\nFor $\\alpha \\in Z_{r, \\overline{p}}(X)$, let $\\alpha^\\nu \\in Z_r(X^\\nu)$ denote its proper transform. By definition, $\\alpha \\in Z_{r, \\overline{p}}(X)$ implies $\\alpha^\\nu \\in Z_{r, \\overline{p}}(X^\\nu)$. If $\\alpha \\in Z_{r, \\overline{p}}(X)$, then $\\dim ( \\nu^{-1}(\\alpha \\cap X^1)) = \\dim (\\alpha \\cap X^1) = r-1$ since $\\nu$ is finite and $\\alpha$ is not contained in $X^1$ (since $p_1=0$). Therefore $\\overline{p}$-rational equivalence on $X$ implies $\\overline{p}$-rational equivalence on $X^\\nu$. Let $\\nu^* : A_{r, \\overline{p}}(X) \\to A_{r, \\overline{p}}(X^\\nu)$ denote the morphism induced by proper transform.\n\nNow suppose we can define a pairing $A_{r, \\overline{p}}(X^\\nu) \\times A_{s, \\overline{q}}(X^\\nu) \\to A_{r+s-d}(X^\\nu)$ when $\\overline{p} + \\overline{q} = \\overline{t}$. Then the composition\n$$A_{r, \\overline{p}}(X) \\times A_{s, \\overline{q}}(X) \\xrightarrow{\\nu^* \\times \\nu^*} A_{r, \\overline{p}}(X^\\nu) \\times A_{s, \\overline{q}}(X^\\nu) \\to A_{r+s-d}(X^\\nu) \\xrightarrow{\\nu_*} A_{r+s-d}(X)$$\ndefines the pairing for $X$.\n\nMoreover, the condition $p_1=0$ implies the generic point of any error term factors through $X^2$: we have $\\dim (\\alpha \\cap X^1) \\leq r-1$, and $\\pi$ is finite over $X \\setminus X^2$, hence $\\dim (\\pi^{-1} (\\alpha \\cap X^1 \\setminus X^2)) \\leq r-1$ and $\\pi^{-1}(\\alpha \\cap X^1 \\setminus X^2)$ cannot support any error terms.\n\n\nIf $X$ is not normal, the stratification ``prescribed by the resolution\" is the one prescribed by $\\widetilde{X^\\nu} \\to X^\\nu \\to X$. We assume $X$ is normal and hence $X^2$ consists of the singular locus of $X$.\n\n\\medskip\nThe following proposition demonstrates our proposal is sensible when the ``resolution\" is the blowup of a smooth variety along a smooth subvariety. The method of proof also shows the condition $\\overline p + \\overline q \\leq \\overline t$ cannot in general be weakened.\n\n\\begin{proposition} \\label{smooth case}\nLet $Y \\hookrightarrow X$ be a closed immersion of smooth varieties of codimension $c$. Let $\\alpha, \\beta \\hookrightarrow X$ be cycles of dimensions $r,s$ satisfying\n$$\\dim(\\alpha \\cap Y) \\leq r-c+p \\ , \\ \\dim(\\beta \\cap Y) \\leq s-c+q,$$\nwith $p+q \\leq c-1$. Let $\\pi: \\widetilde X \\to X$ be the blowup of $X$ along $Y$. Then $\\alpha \\cdot \\beta = \\pi_*( \\widetilde \\alpha \\cdot \\widetilde \\beta)$ as cycle classes on\n$|\\alpha| \\cap |\\beta|$ and hence on $X$.\n\\end{proposition}\n\n\\begin{proof} We have a cartesian diagram:\n\n$$\\xymatrix{ E \\ar[r]^-j \\ar[d]_-g & \\widetilde X\\ar[d]^-\\pi \\\\ Y \\ar[r] & X .\\\\}$$\nLet $z \\in A_*(E)$ denote the class of the canonical $g$-ample line bundle. Then $A_*(E)$ is $A_*(Y) [z]$ modulo a relation involving the Chern classes of the normal bundle $N_YX$; we will need that $g^*$ is a ring homomorphism, and that $g_* (z^n) = 0$ for $n \\leq c-2$.\n\n We set $\\pi^* \\alpha = \\widetilde \\alpha + e_\\alpha$ and $\\pi^* \\beta = \\widetilde \\beta + e_\\beta$ in $A_*(\\widetilde X)$; precise formulas for $e_\\alpha, e_\\beta$ are available \\cite[Thm.~6.7]{Ful}, but we will just need that $\\pi_* e_\\alpha = \\pi_*e_\\beta=0$. Since $\\alpha \\cdot \\beta = (\\pi_* \\pi^* \\alpha) \\cdot \\beta = \\pi_* (\\pi^* \\alpha \\cdot \\pi^* \\beta)$, it suffices to show the terms $\\widetilde \\alpha \\cdot e_\\beta, e_\\alpha \\cdot \\widetilde \\beta,$ and $e_\\alpha \\cdot e_\\beta$ vanish after $\\pi_*$ is applied.\n\nNow we analyze the term $\\widetilde \\alpha \\cdot e_\\beta$; the treatment of the term $e_\\alpha \\cdot \\widetilde \\beta$ merely involves interchanging the roles of $\\alpha$ and $\\beta$. We utilize the Chow rings of $Y$ and $E$ and show the vanishing of $g_* ( j^* (\\widetilde \\alpha) \\cdot e_\\beta )$, which implies $\\pi_* (\\widetilde \\alpha \\cdot e_\\beta) = 0$. Since $j^* (\\widetilde \\alpha)$ is an $(r-1)$-dimensional cycle supported over the ($\\leq (r-c+p)$-dimensional) set $\\alpha \\cap Y$, it can be expressed as\n$$j^* (\\widetilde \\alpha) = z^p g^*( a_{r-c+p} ) + z^{p-1} g^* (a_{r-c+p-1} ) + \\cdots + g^*( a_{r-c} )$$\nwhere $a_i$ is a class in $A_i(Y)$. Similarly, since $e_\\beta$ is an $s$-dimensional cycle supported over $\\beta \\cap Y$, we have an expression\n$$e_\\beta = z^{q-1} g^* (b_{s-c+q}) + z^{q-2} g^* (b_{s-c+q-1}) + \\cdots + g^* ( b_{s-c+1})$$\nwith $b_i \\in A_i(Y)$. Therefore $ j^* (\\widetilde \\alpha) \\cdot e_\\beta$ is a sum of terms of the shape $z^n g^* ( c_n )$ (with $c_n \\in A_{r+s-d+1+n-c}(Y)$) with $n \\leq p+q -1$. Now the projection formula for $g$, the formula $g_* (z^n) = 0$ for $n \\leq c-2$, and the hypothesis $p+q \\leq c-1$ together imply that $g_* ( j^* (\\widetilde \\alpha) \\cdot e_\\beta) =0 $, as desired.\n\nTo show the vanishing of the term $e_\\alpha \\cdot e_\\beta$, we note that\n$$e_\\alpha \\cdot_{\\widetilde X} e_\\beta = j_* ( c_1(N_{E} \\widetilde X) \\cap (e_\\alpha \\cdot_E e_\\beta) ),$$\nwhere we have used a subscript to indicate on which variety we calculate the product. The largest $z$-degree in $e_\\alpha \\cdot_E e_\\beta$ is $p+q-2$, and $ c_1(N_{E} \\widetilde X) = -z$, so the same argument as in the previous paragraph shows the vanishing of $g_* ( c_1(N_{E} \\widetilde X) \\cap (e_\\alpha \\cdot e_\\beta) )$, hence $\\pi_* (e_\\alpha \\cdot_{\\widetilde X} e_\\beta) = 0$.\n\\end{proof}\n\n\n\\begin{remark}\nThere does not seem to be a simple inductive argument which allows one to conclude Proposition \\ref{smooth case} holds for a composition of blowups along smooth centers $\\pi: X_2 = \\text{Bl}_{Y_1}(X_1) \\xrightarrow{\\pi_2} X_1 = \\text{Bl}_{Y_0}(X_0) \\xrightarrow{\\pi_1} X_0 =X$ for the following reason: if cycles on $X_0$ have controlled incidence with $Y_0$, it is not necessarily the case that their proper transforms via $\\pi_1$ will have suitably controlled incidence with $Y_1 \\hookrightarrow X_1$.\nFor example, suppose $Y_0$ is a linearly embedded $\\bb{P}^2$ in $X_0 = \\bb{P}^4$, and $Y_1$ is the preimage ${(\\pi_1)}^{-1}(L)$ of a line $L \\hookrightarrow Y_0$ via the blowup.\nThen $Y_1 \\to L$ is a smooth morphism between smooth varieties, and has geometrically integral fibers. The only sensible stratification is\n$$\\emptyset = X^4 \\hookrightarrow L = X^3 \\hookrightarrow Y_0 = X^2 = X^1 \\hookrightarrow X ;$$\nsince no lines are distinguished and no line has a distinguished point, there is no canonical way to define a non-empty finite set $X^4$. Now if $S \\hookrightarrow X$ satisfies $\\dim (S \\cap X^2) = \\dim (S \\cap X^3) = 0$, then $S$ satisfies the perversity condition $ \\overline p = (0, 1, 1)$ (and hence also the condition $\\overline q = (1,1,2)$). Therefore the pair $(S,S)$ satisfies a pair of complementary perversity conditions. However, the proper transform $S_1 \\hookrightarrow X_1$ satisfies $\\dim (S_1 \\cap Y_1) = 1$, so that $(S_1, S_1)$ violates the condition $p + q \\leq 1$ for the morphism $\\pi_2$ required by Proposition \\ref{smooth case}.\n\nThe self-intersection $S \\cdot S$ agrees with $\\pi_* (S_2 \\cdot S_2)$,\nbut the ``reason\" involves a feature of the morphism $X_2 \\to X_0$ which cannot be seen from its constituent factors: the ``error term\" $\\pi^* (S) - S_2$ is represented by surfaces supported over the finite set $S \\cap X^3$, and these can be moved away from each other (and from $S_2 \\cap \\pi^{-1}(S \\cap X^3)$) via rational functions on $L$.\nIn fact the same procedure works in the singular case, with possible modifications due to the more complicated geometry; see the penultimate paragraph of the proof of Proposition \\ref{intro smooth ex} for details.\n\\end{remark}\n\n\n\\section{Intersection with divisors} \\label{sec:divisors} In this section we construct a pairing between divisors and $r$-cycles (for $r \\geq 1$) by using a stratification obtained from a resolution of singularities. We make use of the ring structure on $A_* (X)$ for nonsingular $X$, and we use that products respect supports in the following sense: if $A, B \\hookrightarrow X$ are cycles of dimensions $r,s$, then $A \\cdot B \\in A_{r+s-d}( |A | \\cap | B|)$ is a well-defined cycle class \\cite[Ch.~8]{Ful}. Often we use that $|A| \\cap |B|$ (or its image via a resolution) cannot support a cycle of the relevant dimension.\n\nWe use $(-)^i$ to indicate the intersection $|(-)| \\cap |X^i|$ in the remainder of the paper.\n\n\n\\begin{theorem} \\label{divisor pairing}\nLet $X$ be a $d$-dimensional variety over $k$, and let $\\pi : \\widetilde{X} \\to X$ be a strong resolution of singularities. \nSuppose a stratification of $X$ satisfies (BC-1) and (AC) with respect to $\\pi$.\nLet $\\overline{p}, \\overline{q}$ be perversities such that $\\overline{p} + \\overline{q} = \\overline{t}$.\nThen the assignment\n$$(D, \\alpha) \\mapsto \\pi_* (\\widetilde{D} \\cdot \\widetilde{\\alpha})$$\ndetermines a well-defined pairing\n$$A^w_{d-1, \\overline{p}}(X)_\\mathbb{Q} \\otimes A^w_{r, \\overline{q}}(X)_\\mathbb{Q} \\to A_{r-1}(X)_\\mathbb{Q}$$\nfor any $r \\geq 1$.\n\\end{theorem}\n\n\\begin{proof} First we show the assignment is compatible with $\\overline{p}$-rational equivalences of divisors. If $D$ and $D'$ are divisors with $D \\sim_{w, \\overline{p}} D'$, then in particular $D - D' = \\divfcn (f)$ for some $f \\in k(X)$. Viewing instead $f \\in k (\\widetilde{X})$, we find $\\widetilde{D} - \\widetilde{D'} = e_D + \\divfcn(f)$, where the error term $e_D$ is an exceptional divisor for $\\pi$, and is supported over $(D \\cup D') \\cap X^2$. Since $\\widetilde{\\alpha}$ is not contained in the exceptional locus, the intersection $\\widetilde{\\alpha} \\cap e_D$ is proper.\n\nSince a stratification satisfying (BC-1) refines the fiber dimension stratification, the preimage $\\pi^{-1}(T)$ of a divisor $T \\hookrightarrow X^i \\setminus X^{i+1}$ has dimension no larger than $(d-i-1)+(i-1) = d-2$. Therefore $e_D$ is supported over the strata for which $p_i \\geq 1$. For these strata, $q_i \\leq i-2$, so $\\dim (\\alpha^i) \\leq r -2$. But now $\\pi_* ( e_D \\cdot \\widetilde{\\alpha})$ is an $(r-1)$-cycle supported in a subscheme of dimension $r-2$, hence $\\pi_* ( e_D \\cdot \\widetilde{\\alpha}) = 0$.\n\nThe vanishing of $\\pi_* (e_D \\cdot e_\\alpha)$ is proved by a similar argument: if $e_\\alpha \\not \\subseteq e_D$, then an identical argument applies. If $e_\\alpha \\subseteq e_D$, then $e_\\alpha$ is supported over a subscheme of dimension at most $r-2$ (since it is supported over those strata for which $q_i \\leq i-2$), so that $e_D \\cdot e_\\alpha$ is represented by a cycle having positive generic dimension over its image.\n\nNow we show the assignment is compatible with $\\overline{q}$-rational equivalences of $r$-cycles, i.e., the vanishing $\\pi_* (\\widetilde D \\cdot e_\\alpha)= 0$. If $\\alpha \\sim_{w, \\overline{q}} \\alpha'$, then $\\widetilde{\\alpha} - \\widetilde{\\alpha'} \\in A_r(\\widetilde{X})$ is represented by a cycle which is supported over $A \\cap X^2$, where $A$ is an $r$-cycle of perversity $\\overline{q}$. (Namely $A$ consists of the union of the supports of the terms appearing in the equation of rational equivalence, i.e., the image via $\\pi$ of $e(\\alpha, \\alpha')$.)\n\nFirst assume $q_2 = 0$, so $\\alpha$ and $\\alpha'$ meet the singular locus $X^2$ properly. Then any component $e_\\alpha$ of $e(\\alpha, \\alpha)$ is an $r$-cycle having generic dimension at least $2$ over its image, and this image is contained in the (at most) $(r-2)$-dimensional set $A \\cap X^2$. Therefore $ \\widetilde D \\cdot e_\\alpha$ is represented by an $(r-1)$-cycle having generic dimension at least $1$ over its image, and so $\\pi_* ( \\widetilde D \\cdot e_\\alpha) =0$.\n\nFinally assume $p_2 = 0$; this is the most interesting case. We may assume $e := e_\\alpha$ is integral and first appears on $X_{n+1} = \\Bl_{C_n} X_n$. Let $\\overline{e} \\hookrightarrow X_{n+1}$ denote the ($r$-dimensional) image of $e$, and let $e_n \\hookrightarrow C_n \\hookrightarrow X_n$ denote the image of $e$. Since $e$ first appears on $X_{n+1}$, there is an exceptional divisor $E'$ for the morphism $X_{n+1} \\to X_n$ which contains $\\overline e$ and dominates $C_n$. Since $e \\to e_n$ is birational, there exists a component $E \\subseteq \\widetilde E \\hookrightarrow \\widetilde X$ which contains $e$ and maps birationally onto $E'$. (Since the exceptional locus of $X_{n+1} \\to X_n$ is flat over $C_n$, every exceptional component of $X_{n+1} \\to X_n$ must dominate $C_n$.)\nLet $R_n \\hookrightarrow C_n$ denote the locus determined by the reducible fibers over $C_n$ as described in condition (AC) in Definition $\\ref{strat res defn}$; we will use that $R_n$ contains the reducible fiber locus of $E \\to C_n$. Let $V \\hookrightarrow W$ denote the image of $R_n$ in $X$. \n\nNote that $\\dim (e_n) \\leq r-1$, hence if $e_n$ has positive generic dimension over its image in $X^2$, then we may proceed as in the case $q_2=0$. So we may assume $e_n$ is generically finite over its image $\\overline{e_n} \\hookrightarrow W \\hookrightarrow X$. Here $W$ denotes the image of $C_n$; say $\\dim W = d-c$, so that $W$ is a component of $X^c \\hookrightarrow X^2$. We have $\\dim ( \\overline{e_n}) = r-1 = (r-c) + (c-1)$, hence $q_c = c-1$ and $p_c = 0$, so $D$ meets $W$ properly.\n\nLet $D_n \\hookrightarrow X_n$ denote the proper transform of $D$ via $X_n \\to X$, and $\\widetilde{D} \\hookrightarrow \\widetilde{X}$ its proper transform via $\\widetilde{X} \\to X$. Since $D$ does not contain $W$, $D_n$ does not contain $C_n$ and $\\widetilde{D}$ does not contain $E$.\nThe situation is summarized by the following commutative diagram:\n\n\n$$\\xymatrix{ & e \\ar[r] \\ar[d] & E \\ar[r]^-i \\ar[d]_{\\pi _E } & \\widetilde{X} \\ar[d]^-\\pi & \\widetilde{D} \\ar[l] \\ar[d] \\\\\nR_n \\ar@\/_1pc\/[rr]|\\hole \\ar[d] & e_n \\ar[r] \\ar[d] & C_n \\ar[d] \\ar[r]^-{i_n} & X_n \\ar[d] & D_n \\ar[l] \\ar[d] \\\\\nV \\ar@\/_1pc\/[rr] & \\overline{e_n} \\ar[r] & W \\ar[r] & X & D \\ar[l] }$$\n\n\\medskip\n\nWe assume that $R_n$ is not contained in $D_n$, and later we handle the case $D_n \\supseteq R_n$ by a separate argument. \nA priori we have $\\pi_E^{-1} |D_n \\cap C_n| \\supseteq | \\widetilde{D} \\cap E |$, but $D_n \\not \\supset R_n$, together with the condition on fiber integrality, implies the generic points of $\\pi_E^{-1} |D_n \\cap C_n|$ coincide with those of $|\\widetilde{D} \\cap E|$. \nSince $C_n$ is smooth, $D_n \\cap C_n$ is the support of a Cartier divisor in $C_n$. Set $M := \\mathcal{O}_{C_n} (|D_n \\cap C_n|) \\in \\pic(C_n)$, and $L := \\mathcal{O}_{\\widetilde{X}} (\\widetilde{D}) \\in \\pic(\\widetilde{X})$. The coincidence of the generic points implies $i^*L$ and ${\\pi_E}^* M$ are rational multiples of one another in $\\pic (E)_\\mathbb{Q}$; say $m \\cdot {\\pi_E}^* M = i^* L$ for some $m \\in \\mathbb{Q}$.\n\nSince ${\\pi_E}_* (e) = 0$, the projection formula for $\\pi_E$ implies\n\\begin{equation} \\label{push 0} {\\pi_E}_* (c_1 ( {\\pi_E}^*(M)) \\cdot e) = c_1(M) \\cdot {\\pi_E}_*(e) = 0. \\end{equation}\nThe projection formula for $i$ implies the equality of cycle classes\n\\begin{equation} \\label{proj i} i_* ( c_1 ( i^* L ) \\cdot e) = c_1(L ) \\cdot e \\in A_{r-1}(\\widetilde{X}). \\end{equation}\nThe functoriality of proper push-forward and the relation in $\\pic (E)_\\mathbb{Q}$ imply\n\\begin{equation} \\label{last one} \\pi_* i_* ( c_1 ( i^* L ) \\cdot e) = {i_n}_* {\\pi_E}_* ( c_1 ( i^* L ) \\cdot e) = m \\cdot {i_n}_* {\\pi_E}_* ( c_1({ \\pi_E}^* ( M )) \\cdot e). \\end{equation}\nThese equations together imply the vanishing $\\pi_* ( c_1(L) \\cdot e) = 0$ in $A_{r-1}(X)_\\mathbb{Q}$.\n\nWe return to the case in which $D_n \\supseteq R_n$. Then $D$ contains the image $V$ of $R_n$ in $X$, and $V$ is a component of $X^{d-c+m}$ (for some $m \\geq 1$, since $p_c=0$). Then $p_{d-c+m} \\geq 1$, hence $q_{d-c+m} \\leq d-c+m -2$ and therefore $\\dim (A \\cap V) \\leq r-2$.\nWe have a commutative diagram relating push-forward and pull-back\n$$\\xymatrix{A_{r-1} (\\widetilde{X}) \\ar[d] \\ar[r]^-{\\pi_*} & A_{r-1}(X) \\ar[d]^-\\cong \\\\\nA_{r-1} (\\widetilde{X} \\setminus \\pi^{-1} | A \\cap V| ) \\ar[r] & A_{r-1} (X \\setminus | A \\cap V|) }$$\nso it suffices to show $\\pi_* ( \\widetilde{D} \\cdot e) $ vanishes upon restriction to $X \\setminus | A \\cap V|$. \nAway from $V$, the generic points of $\\pi_E^{-1} |D_n \\cap C_n|$ coincide with those of $|\\widetilde{D} \\cap E|$ (as in the case $R_n \\not \\subset D_n$), and the vanishing of $\\pi_* (\\widetilde D \\cdot e)$ in $A_{r-1}(X)$ follows.\n\\end{proof}\n\\begin{remark} Outside of the case $p_2=0$, one can work with the coarser fiber dimension stratification, and with integral coefficients. \\end{remark}\n\n\\section{Intersections on a variety with one-dimensional singular locus} \\label{sec:onedimsing}\n\nIn this section we show a resolution may be used to defined intersection pairings on a variety with one-dimensional singular locus. At one place we need the generic smoothness of a morphism between smooth integral $k$-schemes, so our results in this section are aimed at the case $\\charct k = 0$, though they apply to special situations in positive characteristic. We do not require a strong resolution, but we require the smoothness of the components of the exceptional locus $\\widetilde E \\hookrightarrow \\widetilde X$. Resolutions as in Theorem $\\ref{one dim sing}$ exist by \\cite[Thm.~1.6(2)]{BierMil}, or, since we do not use the smoothness of the centers, \\cite{Kollar:res}.\n\n\\medskip\n\\textbf{The stratification.} Let $X$ be a variety over $k$. Suppose $\\pi : \\widetilde X \\to X$ is a resolution of the singularities of $X$ such that all exceptional components are smooth over $k$, and generically smooth over their images in $X$. We set $X^{d-1}$ equal to the singular locus of $X$, and we define $X^d \\hookrightarrow X^{d-1}$ to be the smallest set with the following properties:\n\\begin{enumerate}\n\\item $X^d$ contains the singularities of $X^{d-1}$, i.e., contains the singularities of each component of $X^{d-1}$, and contains the points at which the components intersect;\n\\item $X^d$ contains the image of every exceptional divisor $E \\subseteq \\widetilde E \\hookrightarrow \\widetilde X$ that is contracted to a point by the resolution $\\pi : \\widetilde X \\to X$; and\n\\item for every exceptional divisor $E$ with 1-dimensional image $E_1$, suppose the morphism $E \\to E_1$ has smooth generic fiber (e.g., $\\charct k =0$); then $X^d$ contains the image of the singular fibers of $E \\to E_1$.\n\\end{enumerate}\n\n\\begin{remark} \nCondition (1) is a mild strengthening of condition (BC-3) of Definition \\ref{strat res defn}, since we additionally take into account the incidences of the components of $X^{d-1}$. Condition (BC-1) implies condition (2) must be satisfied. Condition (3) above is a strengthening of condition (AC).\n\\end{remark}\n\n\\begin{theorem} \\label{one dim sing} Suppose $X$ is a variety over a field $k$ such that $\\dim(X_{sing}) =1$, and suppose $\\pi : \\widetilde X \\to X$ is a resolution of singularities such that the exceptional components are \n\\begin{itemize}\n\\item$k$-smooth, and \n\\item generically smooth over their images in $X$.\n\\end{itemize}\nLet $\\overline p, \\overline q$ be perversities such that $\\overline p + \\overline q = \\overline t$.\nWith respect to the stratification defined above, the assignment $\\alpha, \\beta \\mapsto \\pi_* ( \\widetilde \\alpha \\cdot \\widetilde \\beta)$ determines a well-defined pairing $A^w_{r, \\overline p}(X) \\otimes A^w_{s, \\overline q}(X) \\to A_{r+s-d} (X)$. \\end{theorem}\n\nAs a matter of notation, we mostly let $\\alpha$ denote the factor in which error terms are considered, so that $e_\\alpha \\in Z_*(\\widetilde X)$ is a cycle which arises by taking the proper transform of an equivalence (respecting some perversity condition) relating (say) $\\alpha$ to $\\alpha'$ on $X$.\n\n\\begin{proof} The claim is obvious if $r+s-d \\geq 2$, for then $e_\\alpha \\cdot \\widetilde \\beta, \\widetilde \\alpha \\cdot e_\\beta,$ and $e_\\alpha \\cdot e_\\beta$ are all represented by cycles of dimension $\\geq 2$ supported over $X^{d-1}$, hence have generic dimension $\\geq 1$ over their images, so that all vanish after $\\pi_*$ is applied. This case requires no perversity condition at all.\n\nConsider the case $r+s-d=1$. If the error term $e_\\alpha$ is supported over a finite set, then $e_\\alpha \\cdot \\widetilde \\beta$ is represented by a 1-cycle which is contracted to a finite set by $\\pi$ (since $e_\\alpha$ is so contracted), hence $\\pi_* (e_\\alpha \\cdot \\widetilde \\beta) = 0$. The same argument works with $e_\\beta$ in the place of $\\widetilde \\beta$. Therefore we may assume $\\dim (\\alpha^{d-1}) =1$, so that $p_{d-1} = d - r$ and $q_{d-1} \\leq r-2$, and hence $\\dim (\\beta^{d-1}) \\leq 0$. Working one component at a time, we may assume $e_\\alpha$ is contained in a single (smooth) exceptional component $i :E \\hookrightarrow \\widetilde X$. Now $e_\\alpha \\cdot \\widetilde \\beta = i_* (e_\\alpha \\cdot i^* (\\widetilde \\beta))$, and $i^* (\\widetilde \\beta)$ is represented by an $(s-1)$-dimensional cycle supported over the finite set $\\beta^{d-1}$. Therefore $e_\\alpha \\cdot i^* (\\widetilde \\beta)$ is represented by a 1-cycle which is contracted to a finite set, so $ ( \\pi \\circ i)_* ( e_\\alpha \\cdot i^* (\\widetilde \\beta) ) = 0$. Therefore $\\pi_* (e_\\alpha \\cdot \\widetilde \\beta) = 0$, as desired. Since one of $e_\\alpha, e_\\beta$ must be supported over a finite set in $X^{d-1}$, the term $\\pi_* (e_\\alpha \\cdot e_\\beta)$ vanishes for the same reason.\n\nNow we consider the case $r+s-d=0$. Now $\\dim (\\alpha^{d-1}) =1$ implies $\\dim (\\beta^{d-1}) < 0$, so the interesting situation is when both $\\alpha^{d-1}$ and $\\beta^{d-1}$ are finite; only one of $\\alpha, \\beta$ is allowed to meet $X^d$. As in the previous paragraph, we choose a (smooth) exceptional component $i : E \\hookrightarrow \\widetilde X$ containing $e_\\alpha$. Both $e_\\alpha$ and $i^* (\\widetilde \\beta)$ are supported over finite sets in $X^{d-1}$, and their incidence must occur over points in $X^{d-1} \\setminus X^d$. \nAgain working one component at a time, the cycles $e_\\alpha$ and $i^* (\\widetilde \\beta)$ are disjoint unless they are supported over the same point $x \\in X^{d-1} \\setminus X^d$. Let $j : E_x \\hookrightarrow E$ denote the inclusion of the exceptional fiber over $x$. By construction of the stratification, $E_x$ is smooth. \n\nSet $\\beta' := i^* (\\widetilde \\beta)$, and let $N$ denote the normal bundle of the embedding $j$. Using the projection formula and the self-intersection formula (\\cite[Cor.~6.3]{Ful}), we find:\n$${j}_* (e_\\alpha) \\cdot_E {j}_* ( \\beta' ) = j_* (e _\\alpha \\cdot_{E_x} j^* j_* \\beta' ) = j_* (e_\\alpha \\cdot_{E_x} ( c_1 (N) \\cap \\beta')).$$\nBut $E_x$ is principal, hence $c_1 (N) = 0$, and therefore\n$${j}_* (e_\\alpha) \\cdot_E {j}_* (\\beta') = e_\\alpha \\cdot_ E i^* (\\widetilde \\beta) =0.$$\nThe projection formula for $i$ implies $e_\\alpha \\cdot \\widetilde \\beta = 0 \\in A_0( \\widetilde X)$.\n\nThe vanishing of $e_\\alpha \\cdot e_\\beta$ holds for a similar reason: the nontrivial case is when both $e_\\alpha$ and $e_\\beta$ are supported over the same point $x \\in X^{d-1} \\setminus X^d$. But then\n$j_* (e _\\alpha) \\cdot_E j_* (e_\\beta) = j_* ( e_\\alpha \\cdot_{E_x} ( c_1(N) \\cap e_\\beta ) )$, and the vanishing of $c_1(N)$ allows us to conclude. \\end{proof}\n\n\\begin{remark}\nThe basic obstacle to extending the above analysis to the case $\\dim(X_{sing})=2$ is that the singular fibers of $E \\to E_2$ (here $E$ is a component of the singular locus with 2-dimensional image $E_2 \\hookrightarrow X$) may be supported over a divisor in $E_2$, and, in the case $\\alpha^{d-1}$ and $\\beta^{d-1}$ are both finite, the perversity conditions do not rule out the incidence being supported in a singular fiber.\n\\end{remark}\n\n\n\\section{2-cycles on a fourfold} \\label{sec:4fold} In this section we work on a normal quasi-projective $4$-dimensional variety $X$; more precisely we assume the singularities of $X$ occur in codimension at least $2$. As in the previous cases, the resolutions we require exist in characteristic zero.\nWe describe more explicitly the construction of the stratification satisfying (BC-1), (BC-2), (BC-3), and (AC) with respect to a strong resolution. We start by declaring $X^2$ is the singular locus of $X$.\n\n\\begin{enumerate}\n\n\\item (BC-1) For every (smooth, integral) two-dimensional center $S$ such that the composition $S \\to X^2$ is dominant over a component of $X^2$, the image of the positive-dimensional (i.e., one-dimensional) fibers of $S \\to X^2$ is a finite set in $X^2$. Place this set in $X^4$.\n\n\\item For every two-dimensional center $S$ such that the composition $S \\to X^2$ has one-dimensional image $W \\hookrightarrow X^2$,\n\n\\subitem (BC-1) place $W$ in $X^3$, and\n\n\\subitem (BC-3) place the singularities of $W$ in $X^4$.\n\n\\subitem (BC-2) The morphism $p : S \\to W$ has integral generic fiber. Place the zero-dimensional set $ \\{ w \\in W | p^{-1}(w) \\text{ is reducible} \\}$ in $X^4$.\n \n\\item (BC-1) If the two-dimensional center $S$ of a blowup has zero-dimensional image in $X^2$ (so it lies over a single point $x \\in X^2$), then $x$ is placed in $X^4$.\n\n\\item (AC) Let $p : \\widetilde{E}_{\\text{dom }S} \\to S$ denote the canonical morphism. Suppose $\\widetilde{E}_{\\text{dom }S}$ has $t$ components. Then the image in $X^2$ of the closed set $\\{ s \\in S | p^{-1}(s) \\text{ has more than $t$ components} \\}$\nis placed in $X^3$.\n\\item For every one-dimensional center $C$ which is generically finite onto its image in $X^2$,\n\n\\subitem (BC-1) place the image curve in $X^3$, and\n\n\\subitem (BC-3) place its singularities in $X^4$.\n\n\\subitem (BC-1) If a one-dimensional center $C$ has zero-dimensional image in $X^2$ (so it lies over a single point $x \\in X^2$), then $x$ is placed in $X^4$.\n\n\\subitem (AC) Place in $X^4$ the closed set in $C$ over which $\\widetilde{E}_{\\text{dom }C} \\to C$ has more components than does $\\widetilde{E}_{\\text{dom }C}$ itself.\n\n\\item (BC-1) The image in $X^2$ of every zero-dimensional center $Z$ is placed in $X^4$.\n\n\\end{enumerate}\n\nThe instances of condition (BC-1) are necessary to guarantee the stratification refines the stratification by fiber dimension. In the course of the proof we point out where the other conditions are used; the condition (BC-3) in (5) does not seem to be necessary, but (BC-3) in (2) is used.\n\nWe will use notation from the following diagram.\n\n$$\\xymatrix{ \\widetilde{X} \\ar[r] & \\Bl_{Z_n}X_n \\ar[r] & X_n \\ar[r] & \\ldots \\ar[r] & X \\\\\n\\widetilde{E} \\ar[u] & E_n \\ar[u] \\ar[r]^-{p_n} & Z_n \\ar[u] & \\ldots & X^2 \\ar[u] \\\\ }$$\nThe center $Z_n$ will be written as $S_n$ when it is two-dimensional and as $C_n$ when it is one-dimensional. Note that each $Z_n$ is smooth, and each $p_n : E_n \\to Z_n$ is a flat morphism. Zero-dimensional centers play no essential role since incidences in $X^4$ are forbidden by the perversity condition.\nBlowups along three-dimensional centers are finite morphisms (by normal flatness) and do not influence the stratification.\n\n\n\\begin{theorem} \\label{4fold} Let $X$ be a quasi-projective fourfold over $k$, and let $\\pi : \\widetilde X \\to X$ be a strong resolution of singularities.\nSuppose a stratification of $X$ satisfies (BC-1), (BC-2), (BC-3), and (AC) with respect to $\\pi$.\nLet $\\overline{p}, \\overline{q}$ be perversities such that $\\overline{p} + \\overline{q} = \\overline{t}$. The assignment $(\\alpha, \\beta) \\mapsto \\pi_* (\\widetilde{\\alpha} \\cdot \\widetilde{\\beta})$ determines a well-defined pairing $A^w_{2, \\overline{p}} (X)_\\mathbb{Q} \\otimes A^w_{2, \\overline{q}} (X)_\\mathbb{Q} \\to A_0(X)_\\mathbb{Q}.$\n\\end{theorem}\n \n\nThere are essentially three complementary pairs of perversities to analyze, and these are handled in the next three propositions. For each pair the strategy is to consider possible locations of the generic points of error terms, and for each location we move the error term away from the other 2-cycle. If the error term first appears on the blowup of $X_n$ along $Z_n$, then the move is achieved by finding a suitable rational function on $Z_n$. \n\n\\begin{proposition} Theorem \\ref{4fold} is true for $\\overline{p}=\\overline{0}, \\overline{q} = \\overline{t}$; and for $\\overline{p}=(0,0,1), \\overline{q}=(1,2,2)$. \\end{proposition}\n\n\\begin{proof} If $\\alpha \\sim_{w, \\overline{p}} \\alpha'$ and $z$ is a cycle appearing in the equation relating $\\alpha$ to $\\alpha'$, the dimension of $\\pi^{-1}(z^i)$ is at most $2-i + (i-1) =1$. Since the preimage of the exceptional part is $1$-dimensional, it cannot support a $2$-cycle and the error terms vanish. Therefore $\\widetilde{\\alpha} \\sim \\widetilde{\\alpha'}$ and clearly then $\\pi_* (\\widetilde{\\alpha} \\cdot \\widetilde{\\beta}) \\sim \\pi_* (\\widetilde{\\alpha'} \\cdot \\widetilde{\\beta})$.\n\nIt remains to check the compatibility with $\\overline{q}$-rational equivalence in $\\beta$, i.e., the vanishing $\\pi_* (\\widetilde{\\alpha} \\cdot e_\\beta) =0$, where $e_\\beta$ is a component of the error term $e(\\beta, \\beta')$. Now $\\alpha^2$ is a finite set contained in the smooth part $X^2 \\setminus X^3$ of $X^2$, and $\\alpha^3$ is empty. If the generic point of $e_\\beta$ lies over $X^3$, then $\\widetilde{\\alpha} \\cap e_\\beta = \\emptyset$ and we are done.\n\nTherefore we suppose the generic point of $e_\\beta$ lies over $X^2 \\setminus X^3$, and let $E \\subseteq \\widetilde{E} \\hookrightarrow \\widetilde{X}$ be a component of the exceptional locus which contains $e_\\beta$. If the image of $E$ has dimension less than or equal to $1$, then $E$ has generic dimension at least $2$ over its image, contradicting the assumption that the generic point of $e_\\beta$ lies over $X^2 \\setminus X^3$. Therefore $e_\\beta$ first appears on some blowup $\\Bl_{S_n}X_n$ where $S_n$ is a smooth surface on some intermediate variety $X_n$. Now let $E$ denote an exceptional divisor which contains $e_\\beta$ (as above), and maps birationally onto a divisor which is exceptional for the morphism $\\Bl_{S_n}X_n \\to X_n$ (i.e., $E$ also first appears on $\\Bl_{S_n}X_n$).\n\n Let $a : S_n \\to X^2$ denote the canonical morphism. Note that ${a}^{-1}(\\alpha^2)$ is finite by our construction of the stratification.\nLet $e_n$ denote the (one-dimensional) image of $e_\\beta$ in $S_n$. There exists a $1$-cycle $C$ on $S_n$ which is rationally equivalent to $e_n$, and with the property that $C \\cap {a}^{-1}(\\alpha^2) = \\emptyset$. In other words, there is a rational function $g_n: S_n \\dashrightarrow \\mathbb{P}^1$ such that $[g_n(0)] = e_n + Z$, and such that both $Z$ and $[g_n(\\infty)]$ are disjoint from ${a}^{-1}(\\alpha^2)$.\n\nNow consider the composition $g: E \\to S_n \\dashrightarrow \\mathbb{P}^1$. By the fiber integrality hypothesis (i.e., since (AC) in (4) places the image of the reducible fibers of $E \\to S_n$ into $X^3$), the support of $e_\\beta$ coincides (set-theoretically) with the preimage of $e_n$ by the morphism $E \\to S_n$. \nTherefore $g$ provides a rational equivalence between $m \\cdot e_\\beta + Z'$ and $Z''$, where $Z'$ is supported over $Z$, and $Z''$ is supported over $[g_n(\\infty)]$. Hence both $Z'$ and $Z''$ are disjoint from $\\widetilde{\\alpha}$. Therefore $\\widetilde{\\alpha} \\cdot (m \\cdot e_\\beta) \\sim \\widetilde{\\alpha} \\cdot (Z'' - Z') = 0$ and so $\\widetilde{\\alpha} \\cdot e_\\beta$ is zero in $A_0(\\widetilde{X})_\\mathbb{Q}$, as desired.\n\nAn identical argument handles the pair $\\overline{p}=(0,0,1), \\overline{q}=(1,2,2)$ since it imposes the same conditions on $\\alpha$ and $\\beta$. \\end{proof}\n \n\\begin{proposition} \\label{intro smooth ex} Theorem \\ref{4fold} is true for $\\overline{p}=(0,1,1), \\overline{q} = (1,1,2)$. \\end{proposition}\n\n\\begin{proof} In this case $\\alpha^2$ is finite and $\\alpha^4$ is empty, hence if $e(\\alpha, \\alpha')$ is nonempty it consists of several surfaces lying over some zero-cycle $Z \\hookrightarrow X^3 \\setminus X^4$. Thus a component $e_\\alpha$ of $e(\\alpha, \\alpha')$ is an irreducible surface which first appears on $\\Bl_{C_n}X_n$, the blowup of some intermediate variety $X_n$ along a smooth one-dimensional center $C_n$. Let $E$ denote an exceptional divisor that contains $e_\\alpha$ and first appears when $e_\\alpha$ does (so that $E \\hookrightarrow \\widetilde{E}_{C_n}$).\nThe image of $e_\\alpha$ is a point $c \\in C_n$; since $C_n$ is finite over its image in $X^2$, the proper transform of $\\beta$ via $X_n \\to X$ meets $C_n$ in a finite set.\n\nThere exists a rational function $g_n: C_n \\to \\mathbb{P}^1$ such that $[g_n(0)] = c + Z$, and such that both $Z$ and $[g_n(\\infty)]$ are disjoint from the proper transform of $\\beta$. Consider the rational function $g: E \\to C_n \\to \\mathbb{P}^1$. Since the image of $c \\in C_n$ lies in $X^3 \\setminus X^4$, the fiber of $E \\to C_n$ over $c$ is irreducible (by (AC) in (5)), and therefore $e_\\alpha$ coincides set-theoretically with $E \\cap \\pi^{-1}(c)$.\nTherefore $g$ provides a rational equivalence between $m \\cdot e_\\alpha + Z'$ and $Z''$, where $Z'$ is supported over $Z \\hookrightarrow X^3 \\setminus X^4$, and $Z''$ is supported over $[g_n(\\infty)]$. Consequently $e_\\alpha \\cdot \\widetilde{\\beta} = 0$ in $A_0(\\widetilde{X})_\\mathbb{Q}$. Note $e_\\beta$ must be supported over a finite set in $C_n$, so the same construction moves $e_\\alpha$ away from $e_\\beta$.\n\nNow we show the vanishing of $\\widetilde{\\alpha} \\cdot e(\\beta, \\beta')$. First we consider components $e_\\beta$ of $e(\\beta, \\beta')$ whose generic points lie over $X^2 \\setminus X^3$. Then $e_\\beta$ first appears on the blowup of a two-dimensional center $S_n$ which is dominant over a component of $X^2$. The image of $e_\\beta$ in $S_n$ is a subvariety of dimension $1$, and $e_n := p_n (e_\\beta)$ is not contracted by $a: S_n \\to X^2$. (If $e_\\beta$ were supported over a subvariety contracted by $a$, then it would have generic dimension $2$ over its image.) We choose an exceptional divisor $E \\supset e_\\beta$ as usual.\n\n\nSince $\\alpha^4$ is empty, ${a}^{-1}(\\alpha^2)$ is finite. Therefore $e_n$ is rationally equivalent to a $1$-cycle on $S_n$ which is disjoint from ${a}^{-1}(\\alpha^2)$, i.e., there exists a rational function $g_n: S_n \\dashrightarrow \\mathbb{P}^1$ such that $[g_n(0)] = e_n + Z'$, and such that both $Z'$ and $[g_n(\\infty)]$ are disjoint from ${a}^{-1}(\\alpha^2)$. Now we consider the composition $g := g_n \\circ p_n : E \\to S_n \\dashrightarrow \\mathbb{P}^1$. By the integrality condition on the fibers (condition (AC) in (4)), $[g(0)]$ is a multiple of $e_\\beta$ plus a 2-cycle supported over $Z'$, and $[g(\\infty)]$ is a 2-cycle supported over $[g_n(\\infty)]$. In particular, some multiple of $e_\\beta$ is rationally equivalent to a cycle disjoint from $\\widetilde{\\alpha}$.\n\n\nNext we consider error terms $e_\\beta$ whose generic points are supported over $X^3 \\setminus X^4$. There are two possibilities for first appearance.\nSuch a term may first appear on the blowup along a (smooth) one-dimensional center $C_n$ (with one-dimensional image in $X^2$); \nthis is handled by finding a rational function on $C_n$ which, upon precomposing with a canonical morphism from an exceptional divisor to $C_n$, gives a rational equivalence between some multiple of $e_\\beta$ and a $2$-cycle which is disjoint from the preimage of $\\alpha^3$ (as $\\alpha^3$ necessarily meets $C_n$ in a finite set) and hence from $\\widetilde{\\alpha}$.\n\nThe other possibility is that the error term $e_\\beta$ (with generic points supported over $X^3 \\setminus X^4$) first appears on the blowup along a (smooth) two-dimensional center $S_n$, in which case the one-dimensional image of $e_\\beta$ inside $S_n$ is contracted to a point by the morphism $S_n \\to X^2$: either the center $S_n$ has one-dimensional image in $X^2$, or $e_\\beta$ is supported over some exceptional part of the generically finite morphism $S_n \\to X^2$. But $e_\\beta$ is exceptional for $S_n \\to X^2$ implies the generic point of $e_\\beta$ is supported over $X^4$, so we may assume the two-dimensional center $S_n$ has one-dimensional image $W \\hookrightarrow X^2$. By our definition of $X^4$, (the finite sets) $\\alpha^3 \\cap W$ and $\\beta^3 \\cap W$ are supported in the smooth locus of $W$, and in the locus over which $S_n \\to W$ has integral fibers (by (BC-2) and (BC-3) as in (2)). Let $q$ denote the canonical morphism $E \\to S_n \\to W$, and let $w \\in W$ be the image of $e_\\beta$ via $q$. Note that $e_\\beta$ coincides set-theoretically with $q^{-1}(w)$. There exists a rational function $g: W \\dashrightarrow \\mathbb{P}^1$ such that $[g(0)] = w + Z$, and such that both $Z$ and $[g(\\infty)]$ are disjoint from $\\alpha^3$. Then the rational function $E \\to S_n \\to W \\xrightarrow{g} \\mathbb{P}^1$ provides a rational equivalence between some multiple of $e_\\beta$ and a cycle which is disjoint from the preimage of $\\alpha^3$, so that $\\widetilde{\\alpha} \\cdot e_\\beta \\sim 0$ as desired.\n\nFinally, components $e_\\beta$ of $e(\\beta, \\beta')$ supported over $X^4$ are automatically disjoint from $\\widetilde{\\alpha}$. In particular this applies if $e_\\beta$ is supported over some exceptional part of the generically finite morphism $S_n \\to X^2$. \\end{proof}\n\n\\begin{proposition} Theorem \\ref{4fold} is true for $\\overline{p}=(0,1,2), \\overline{q} = (1,1,1)$. \\end{proposition}\n\n\\begin{proof} In this case $\\alpha^2$ is finite (but may have support in $X^4$), $\\beta^2$ is $1$-dimensional, $\\beta^3$ is finite, and $\\beta^4$ is empty. The error terms appearing in $e(\\alpha, \\alpha')$ cannot be supported over $X^2 \\setminus X^3$ since $p_2=0$. The error terms supported over $X^3 \\setminus X^4$ are handled (moved away from the finite set $\\beta^3$) as in the previous case. The error terms supported over $X^4$ are disjoint from $\\widetilde{\\beta}$ since $\\beta^4 = \\emptyset$. The same reasoning applies with $\\widetilde \\beta$ replaced by $e_\\beta$.\n\nNow we show the vanishing $\\pi_* (\\widetilde \\alpha \\cdot e_\\beta) =0$.\nThe error terms in $e(\\beta, \\beta')$ lying over $X^3$ are handled as in the previous case. It remains to show $\\widetilde{\\alpha} \\cdot e_\\beta = 0$ when $e_\\beta$ is a component of $e(\\beta, \\beta')$ whose generic point is supported over $X^2 \\setminus X^3$. We let $e_0$ denote the image of $e_\\beta$ in $X^2$, and let $X^2_e$ denote the component of $X^2$ that contains $e_0$.\n\nLet $S$ denote the first two-dimensional center that dominates $X^2_e$. Then $S \\to X^2_e$ is a resolution of singularities, and it (Stein) factors as $S \\to S' \\to X^2_e$, where $b: S' \\to X^2_e$ omits those blowups along centers landing in $X^4$. (While $S$ is a closed subvariety in some intermediate $X_j$, the variety $S'$ may not admit a closed immersion into any $X_j$.) Thus $b$ is a finite morphism, the singular set of $S'$ is contained in $b^{-1}(X^4)$, and the proper transform $e_0^{'} \\hookrightarrow S'$ of $e_0$ via $S' \\to X^2_e$ is supported in the smooth locus of $S'$.\n\nSuppose $e_\\beta$ first appears on the blowup of the two-dimensional center $S_n$, and consider the Stein factorization $S_n \\to S_n' \\to S'$ of the morphism $S_n \\to S'$. (Note $S_n \\to S$ is dominant since the generic point of $e_\\beta$ is supported over $X^2 \\setminus X^3$, and of course $S_n=S$ is possible.) Again choose an exceptional component $E \\supset e_\\beta$ dominating $S_n$. Let $e_n$ denote the image of $e_\\beta$ in $S_n$, and $e_n'$ its image in $S_n'$. Since $S_n' \\to S' \\to X^2_e$ is finite, the preimage of $X^4$ in $S_n'$ is finite.\n\nThe morphism $S_n \\to S_n'$ contracts exactly those curves lying over curves contracted by $S \\to S'$. Since the curves contracted by $S \\to S'$ (more precisely the zero-dimensional image in $X^2$ of such curves) are disjoint from $e_0$, the morphism $S_n \\to S_n'$ is an isomorphism in a neighborhood of $e_n$, hence $S_n'$ is smooth in a neighborhood of $e_n'$. Now we proceed as usual. We find a rational function on $S_n'$ moving $e_n'$ away from the finite preimage of $X^4$. Then the composite rational function $E \\to S_n \\to S_n' \\dashrightarrow \\mathbb{P}^1$ moves a multiple of $e_\\beta$ away from $\\widetilde{\\alpha}$, as required.\n\\end{proof}\n\n\\section{Necessity of condition (AC)} \\label{need integral fibers}\n\nSince the arguments of the previous section may seem intricate, one may ask if there is a simpler description of a stratification for which intersecting on the resolution induces a well-defined pairing. In this section we show the stratification according to the fiber dimension of the resolution is insufficiently fine to obtain a well-defined intersection product, even when all the strata are smooth. The example applies to both equivalence relations $\\sim_{\\overline p}$ and $\\sim_{w, \\overline p}$.\nWe work over a field $k$ of characteristic $\\neq 2$.\n\n\\medskip\n\\textbf{An affine example.} Let $X \\hookrightarrow \\mathbb{A}^4$ be the hypersurface defined by the vanishing of $x^2 - y^2 + tz^2$. Then $X$ is singular along the line $L$ defined by $x=y=z=0$, and we claim a resolution $\\pi : \\widetilde{X} \\to X$ is obtained by blowing up $X$ along $L$. On the patch where $x$ generates (say $y = y'x$ and $z = z'x$), the blowup is defined by the vanishing of $1- {(y')}^2 + t{(z')}^2$, which is smooth. On the patch where $y$ generates, the blowup is defined by the vanishing of ${(x')}^2 -1 + t{(z')}^2$, which is also smooth. On the patch where $z$ generates, the blowup is defined by the vanishing of ${(x')}^2 - {(y')}^2 +t$, and this too is smooth.\n\nNote that all of the fibers of $\\pi^{-1}(L) \\to L$ are one-dimensional. Since $L$ is regular, this implies $\\pi^{-1}(L) \\to L$ is flat and so $L \\hookrightarrow X$ is normally flat. Taking into account only the fiber dimensions in the resolution and the singularities of the centers, we are led to the stratification $X^3 = \\emptyset \\hookrightarrow X^2 = X^1 = L \\hookrightarrow X$. The relevant feature (which this stratification ignores) is that the fiber of $\\pi$ over $(0,0,0,0)$ has two components ($x'+y' = 0$ and $x'-y' =0$), whereas the fiber over $(t_0, 0, 0, 0)$ is irreducible for $t_0 \\neq 0$.\n\n\nLet $D \\hookrightarrow X$ be the divisor defined by the ideal $(x+y,t)$, let $\\alpha_0$ be the $1$-cycle defined by $(x-y,t,z)$, and let $\\alpha_1$ be the $1$-cycle defined by $(x-y,t-1,z)$. Each of the cycles $D, \\alpha_0$, and $\\alpha_1$ meets $L$ in a finite set, so $D$ has perversity $\\overline 0$ and the $\\alpha$'s have perversity $\\overline 1 := (1,1,1)$. Furthermore $\\alpha_0 \\sim \\alpha_1$ since the $\\alpha$'s arise as preimages of distinct values of the regular function $t$ on the surface $S$ cut out by $(x-y,z)$. The equivalence respects the perversity condition $\\overline 1$ since $t : S \\to \\mathbb{A}^1$ maps $L$ isomorphically onto $\\mathbb{A}^1$, so that each fiber of $t$ meets $L$ exactly once.\n\nNevertheless, $\\widetilde{D} \\cap \\widetilde{\\alpha_0}$ consists of a single (reduced) point: on the patch where $z$ generates, the intersection occurs at $x' = y' = t = z =0$. On the other hand, $\\widetilde{D} \\cap \\widetilde{\\alpha_1} = \\emptyset$ (in fact $D \\cap \\alpha_1 = \\emptyset$). \nWe conclude that the stratification is too coarse for the resolution to determine a well-defined intersection product.\nOf course the example disappears if we use instead the stratification $X^3 = (0,0,0,0) \\hookrightarrow X^2=X^1 = L \\hookrightarrow X$, as implied by Theorem \\ref{divisor pairing}.\n\n\\medskip\n\\textbf{Behavior at infinity.} For the reader who does not take degrees of zero-cycles on an affine variety too seriously, we now show the behavior persists on the projective closure.\n\nLet $\\underline {X} \\hookrightarrow \\bb{P}^4_{S,T,X,Y,Z}$ be the hypersurface cut out by $SX^2 - SY^2 + T Z^2$. The singular locus of $\\underline X$ consists of three components, each abstractly isomorphic to $\\bb{P}^1$:\n\\begin{itemize}\n\\item $\\Sigma_1 = Z(X,Y,Z)$, which is the closure of $L$ in the affine example;\n\\item $\\Sigma_2 = Z(X - Y, S, Z)$; and\n\\item $\\Sigma_3 = Z(X + Y, S, Z)$.\n\\end{itemize}\n\nThe intersection point $p = [0 : 1 : 0 : 0 : 0] = \\Sigma_1 \\cap \\Sigma_2 \\cap \\Sigma_3$ (which did not appear in the affine example) may be placed in ${\\underline X}^3$ (because the singular locus of $\\underline X$ is itself singular there, or because the resolution has different behavior over $p$). Our constructions will preserve any perversity condition which forbids incidence with $p$, hence will show\n$${\\underline X}^3 = p \\hookrightarrow {\\underline X}^2 = {\\underline X}^1 = \\Sigma_1 \\cup \\Sigma_2 \\cup \\Sigma_3 \\hookrightarrow \\underline X$$\nis not fine enough for the intersection product to be well-defined. As in the affine example, ${\\underline X}^3$ must include $[1:0:0:0:0]$ as well.\n\nA resolution is obtained by blowing up $\\underline X$ along $\\Sigma_1$, then blowing up the proper transforms $\\widetilde{\\Sigma_2}, \\widetilde{\\Sigma_3} \\hookrightarrow B_{\\Sigma_1} (\\underline X)$; since $\\widetilde{\\Sigma_2}$ and $\\widetilde{\\Sigma_3}$ are disjoint on $B_{\\Sigma_1} (\\underline X)$, the order is irrelevant. Initially blowing up $p$ does not improve the situation, in the sense that one finds a copy of the original singularity on $B_p (\\underline X)$. This is perhaps not too surprising since one finds Whitney umbrellas along $x=0$ and $y=0$, and these too are not resolved by blowing up the ``worst\" point in the singular locus.\n\n\\begin{proposition} \\label{ex strong res} The morphism\n$$B_{\\widetilde{\\Sigma_2} \\cup \\widetilde{\\Sigma_3}} (B_{\\Sigma_1} (\\underline X ))\\to B_{\\Sigma_1} (\\underline X) \\to \\underline X$$\nis a strong resolution of the singularities of $\\underline X$. \\end{proposition}\n\\begin{proof}\nFirst we verify that the blowup along $\\Sigma_1 \\cup \\Sigma_2 \\cup \\Sigma_3 - p$ resolves $\\underline X - p$; for this calculation the order of the blowup is irrelevant. Then we analyze the behavior near $p$.\n\nFor the blowup along $\\Sigma_1 - p$, we work on $S \\neq 0$ and recover the affine example: the blowup of $x^2 - y^2 + tz^2$ along $x=y=z=0$. Hence $B_{\\Sigma_1 - p} (\\underline X - p)$ is smooth above $\\Sigma_1 - p$. The analysis of the other blowups is similar (work on $X \\neq 0$) and we omit the details.\n\nHaving analyzed the morphism away from $p$, we analyze the blowup along $\\Sigma_1$ where $T \\neq 0$, so that we are blowing up $sx^2 - sy^2 + z^2 = 0$ along $x=y=z=0$. Where $x$ generates, the blowup is defined by $s - s {(y')}^2 + {(z')}^2 =0$, and is singular along $1-{(y')}^2 = s = z' = 0$; these two singular points are exactly where the proper transforms $\\widetilde{\\Sigma_2}, \\widetilde{\\Sigma_3}$ meet the fiber over $p$. Where $y$ generates, the blowup is defined by $s{(x')}^2 -s + {(z')}^2 =0$, and is singular along ${(x')}^2-1=s=z'=0$ (so that we see the same two points). Where $z$ generates, the blowup is smooth.\n\nThe blowup of $B_{\\Sigma_1}(\\underline X)$ along $\\widetilde{\\Sigma_2}$ is covered by two charts:\n\\begin{itemize}\n\\item the blowup of $s - s{(y')}^2 + {(z')}^2 =0$ along $1-y' = s = z' = 0$, and\n\\item the blowup of $s{(x')}^2 -s + {(z')}^2 =0$ along $x'-1 = s= z' =0$;\n\\end{itemize}\nboth of these are smooth. The blowup along $\\widetilde{\\Sigma_3}$ is similar, except one uses the centers $1+y' = s = z' = 0$, and $x'+1 = s= z' =0$ in these charts.\nAll three centers are normally flat in their ambient spaces since the exceptional divisors over them are irreducible, and the centers are regular and one-dimensional.\n\\end{proof}\n\n\n\\begin{lemma} \nThe stratification\n$${\\underline X}^3 = p \\hookrightarrow {\\underline X}^2 = {\\underline X}^1 = \\Sigma_1 \\cup \\Sigma_2 \\cup \\Sigma_3 \\hookrightarrow \\underline X$$\nsatisfies (BC-1), (BC-2), and (BC-3) with respect to the resolution of Proposition \\ref{ex strong res}. With respect to this stratification,\n\\begin{itemize}\n\\item the divisor $\\underline D = Z(X+Y, T) \\hookrightarrow \\underline X$ satisfies the perversity condition $\\overline 0$, and\n\\item the 1-cycles $\\underline \\alpha_0 = Z(X-Y,T,Z)$ and $\\underline \\alpha_1 = Z(X-Y, T-S,Z)$ satisfy the perversity condition $\\overline 1$ and are equivalent as 1-cycles of perversity $\\overline 1$.\n\\end{itemize} \nNevertheless, $\\deg (\\pi_* ( \\widetilde{\\underline D} \\cdot \\widetilde{\\underline \\alpha_0}) ) =1$ and $\\widetilde{\\underline D} \\cap \\widetilde{\\underline \\alpha_1} = \\emptyset$.\n\\end{lemma}\n\\begin{proof}\nThe first claim follows from the following incidence properties: $\\underline D \\cap \\Sigma_1 = [ 1 : 0 : 0 : 0 : 0 ] \\ , \\ \\underline D \\cap \\Sigma_2 = \\emptyset \\ , \\ \\underline D \\cap \\Sigma_3 = [ 0 :0:1:1:0]$.\n\nNext we analyze the rational equivalence relating $\\underline \\alpha_0 = Z(X-Y,T,Z)$ and $\\underline \\alpha_1 = Z(X-Y, T-S,Z)$. Both of these 1-cycles are lines in $\\bb{P}^2 \\cong Z(X - Y, Z) =: P \\hookrightarrow \\underline X$. Note $\\Sigma_1 \\cup \\Sigma_2 \\hookrightarrow P$, so we must ensure the equivalence $\\underline \\alpha_0 \\sim \\underline \\alpha_1$ can be chosen to respect the perversity condition. Since $\\Sigma_3 \\cap P = p$, the component $\\Sigma_3$ poses no difficulty beyond that presented by $\\Sigma_1 \\cup \\Sigma_2$.\n\nWe claim there exists an $\\mathbb{A}^1$-relative 1-cycle $A \\hookrightarrow \\underline X \\times \\mathbb{A}^1$ with the following properties:\n\\begin{enumerate}\n\\item $A_0 = \\underline \\alpha_0, A_1 = \\underline \\alpha_1$;\n\\item all of the fibers $A_t$ are disjoint from $p$; and\n\\item all of the fibers $A_t$ meet $\\Sigma_1 \\cup \\Sigma_2$ in finitely many points (in fact, in exactly two points).\n\\end{enumerate}\nThis means exactly that $A$ determines an $\\mathbb{A}^1$-family of 1-cycles of perversity $\\overline 1$, so that $\\underline \\alpha_0 \\sim_{\\overline 1} \\underline \\alpha_1$. (Of course $\\underline \\alpha_0$ and $\\underline \\alpha_1$ are then weakly equivalent as 1-cycles of perversity $\\overline 1$.)\n\nWe consider $[\\underline \\alpha_0], [ \\underline \\alpha_1], [\\Sigma_1],$ and $[\\Sigma_2]$ as points of $\\check P$, the $\\bb{P}^2$ dual to $P$. The lines $ \\underline \\alpha_0, \\underline \\alpha_1, \\Sigma_2$ meet at $[0:0:1:1:0]$, and this point does not belong to $\\Sigma_1$. Therefore $[\\underline \\alpha_0], [ \\underline \\alpha_1],$ and $[\\Sigma_2]$ lie on a line $\\ell \\hookrightarrow \\check P$, and $[\\Sigma_1] \\notin \\ell$. Then the family $A$ is the family of lines corresponding to the canonical morphism $\\ell - [\\Sigma_2] \\to \\check P$. Since this family avoids $[\\Sigma_1]$ and $[\\Sigma_2]$, all of the fibers $A_t$ meet $\\Sigma_1 \\cup \\Sigma_2$ exactly twice. Furthermore, exactly one point $[L] \\in \\ell$ corresponds to a line containing $p$; this is $[\\Sigma_2]$, hence all of the fibers $A_t$ are disjoint from $p$.\n\nThe incidences are contained in the locus where $S \\neq 0$, so are captured by the affine situation.\n\\end{proof}\n\n\n\\begin{remark}\nIn this example, the family $A$ of 1-cycles cannot be extended to a $\\bb{P}^1$-family respecting the perversity condition $\\overline 1$. However, by performing the blowups in a different order, one can find an example of the same essential phenomenon, and so that the $\\mathbb{A}^1$-family relating the 1-cycles extends to a $\\bb{P}^1$-family, all of whose fibers satisfy the perversity condition $\\overline 1$. Namely, we first blow up $\\Sigma_2$ on $\\underline X$, then blow up the proper transform of $\\Sigma_3$. The resulting variety $\\underline X'$ is singular exactly along ${\\Sigma'_1}$, the proper transform of $\\Sigma_1$. There are no incidences $\\underline D' \\cap \\underline \\alpha_i'$ in the exceptional divisors over $\\Sigma_2 \\cup \\Sigma_3$, so the incidences are captured by the affine situation. The variety $\\underline X'$ is resolved by blowing up $\\Sigma_1$, and $B_{\\Sigma'_1} (\\underline X')$ has one-dimensional fibers over $\\Sigma'_1$, all of which are irreducible except one (the same which appears in the affine example).\n\nNote $\\Sigma_2 \\hookrightarrow P$ is already Cartier, so the proper transform of $P$ via the blowup along $\\Sigma_2$ is isomorphic to $P$. Since $\\Sigma_3$ meets $P$ in a single point, the transform $P'$ of $P$ on $\\underline X'$ is isomorphic to $\\bb{P}^2$ blown up at a single point. \n\nThe blown up point is not contained in any of the lines $\\Sigma_1', \\underline \\alpha'_0, \\underline \\alpha'_1$. The rational function $P \\dashrightarrow \\bb{P}^1$ relating $\\underline \\alpha_0$ to $\\underline \\alpha_1$ may be considered as a rational function $P' \\dashrightarrow \\bb{P}^1$ relating $\\underline \\alpha'_0$ to $\\underline \\alpha'_1$; since $\\Sigma'_1$ does not occur as a fiber of this map, we may think of this function as a $\\bb{P}^1$-family of 1-cycles of perversity $\\overline 1$ for the stratification ${(\\underline X')}^3 = \\emptyset \\hookrightarrow {(\\underline X')}^2 = {(\\underline X')}^1 = \\Sigma'_1 \\hookrightarrow \\underline X'$.\n\\end{remark}\n\n\\section{Further questions} \\label{sec:further}\n\n\\textbf{Independence of resolution.} Given two resolutions $\\pi_1 :X_1 \\to X \\ , \\ \\pi_2 : X_2 \\to X$ of the variety $X$, we have defined two stratifications $S_1, S_2$ of $X$, and (in certain situations) products $\\bullet_i : A_{r, \\overline p}(X, S_i) \\otimes A_{s, \\overline q}(X, S_i) \\to A_{r+s-d}(X)$ (for $i=1,2$) by pushing forward the intersection formed on the resolution. (We make explicit the dependence of the group $A_{r, \\overline p}(X)$ on the stratification, and for simplicity we drop the superscript $w$ and the coefficients.) For any stratification $S$ that refines both $S_1$ and $S_2$, we have canonical morphisms $C^i_{r, \\overline p}: A_{r, \\overline p}(X, S) \\to A_{r, \\overline p}(X, S_i)$ (for $i=1,2$), and thus it makes sense to ask whether $\\bullet_1 (C^1_{r, \\overline p} \\otimes C^1_{s, \\overline q})$ and $\\bullet_2 (C^2_{r, \\overline p} \\otimes C^2_{s, \\overline q} )$ coincide as morphisms $A_{r, \\overline p}(X, S) \\otimes A_{s, \\overline q}(X, S) \\to A_{r+s-d}(X)$. We assume we are in one of the situations in which the resolution is known to induce a well-defined product.\n\nThe simplest case is when $\\pi_2$ is obtained from $\\pi_1$ by a sequence of blowups of $X_1$ along smooth centers. In this case we have a canonical morphism \n$C_{r, \\overline p} : A_{r, \\overline p}(X, S_2) \\to A_{r, \\overline p}(X, S_1)$ (i.e., the stratification induced by $\\pi_2$ refines the one induced by $\\pi_1$). \n\n\\begin{proposition} With the notation and hypotheses as above, suppose $f : X_2 \\to X_1$ is a composition of blowups along smooth centers, and set $\\pi_2 = \\pi_1 \\circ f : X_2 \\to X$. Then we have $\\bullet_2 = \\bullet_1 (C_{r, \\overline p} \\otimes C_{s, \\overline q}) : A_{r, \\overline p}(X,S_2) \\otimes A_{s, \\overline q}(X, S_2) \\to A_{r+s-d}(X)$. \\end{proposition}\n\n\\begin{proof} Let $(-)_i$ denote the proper transform of a cycle on $X$ via $\\pi_i : X_i \\to X$. We have $f^* (\\alpha_1) = \\alpha_2 + e_\\alpha$ and $f^* (\\beta_1) = \\beta_2 + e_\\beta$. Since $\\alpha_1 \\cdot \\beta_1 = f_* (f^* (\\alpha_1) \\cdot f^* (\\beta_1))$, it follows that\n$${(\\pi_1)}_* (\\alpha_1 \\cdot \\beta_1) = {(\\pi_2)}_* (\\alpha_2 \\cdot \\beta_2) + {(\\pi_2)}_*(\\alpha_2 \\cdot e_\\beta + e_\\alpha \\cdot \\beta_2 + e_\\alpha \\cdot e_\\beta).$$\nThe cycles $e_\\alpha$ and $e_\\beta$ may be thought of as error terms which first appear on one of the blowups occurring in the morphism $f$. Therefore our arguments apply to show the vanishing of the cycle class of ${(\\pi_2)}_*(\\alpha_2 \\cdot e_\\beta + e_\\alpha \\cdot \\beta_2 + e_\\alpha \\cdot e_\\beta)$, and therefore we obtain the equality ${(\\pi_1)}_* (\\alpha_1 \\cdot \\beta_1) = {(\\pi_2)}_* (\\alpha_2 \\cdot \\beta_2).$\n\\end{proof}\n\nIf $k$ is an algebraically closed field of characteristic zero and $X$ is complete, then two resolutions of $X$ are related by a sequence of smooth blowups and blowdowns by the weak factorization theorem of Abramovich-Karu-Matsuki-W{\\l}odarczyk \\cite[Thm.~0.1.1]{AKMW}. Since the intermediate varieties do not necessarily admit morphisms to $X$, it is not clear how to obtain a comparison of the products formed via two resolutions. If the strong factorization conjecture \\cite[Conj~0.2.1]{AKMW} holds, however, then there is variety $Y$ admitting morphisms $f_i : Y \\to X_i$ which are compositions of blowups along smooth centers, and such that $\\pi_1 \\circ f_1 = \\pi_2 \\circ f_2 : Y \\to X$. In this case we could conclude that the products defined using $X_1$ and $X_2$ agree upon restriction to $A_{r, \\overline p}(X, S_Y) \\otimes A_{s, \\overline q}(X, S_Y)$, i.e., the group formed using the stratification induced by the resolution $Y \\to X$.\n\n\n\n\n\\medskip\n\\textbf{Comparison with Goresky-MacPherson product.} It would be interesting to know whether our intersection product (when it is defined) agrees with the intersection pairing defined by Goresky-MacPherson via the cycle class mapping. The difficulty in making the comparison is that there is no obvious way to take the proper transform of a topological cycle. A cycle on $X$ gives rise to a canonical cycle on $\\widetilde X$ relative to $\\widetilde E$, and using this one can show the products agree after composing with the canonical map $H_*(X) \\to H_*(X, X^1)$. For pairs of cycles with supports intersecting properly in each stratum, or more generally any pair which is weakly $(\\sim_{\\overline p}, \\sim_{\\overline q})$-equivalent to such a pair, our intersection product agrees with that of Goresky-MacPherson since in this case both may be described as taking the closure of an intersection product formed on the smooth locus.\n\n\\medskip\n\\textbf{Refinements for small perversities.} If $\\overline p + \\overline q < \\overline t$, is there a refinement\n$$A_{r, \\overline p}(X) \\otimes A_{s, \\overline q}(X) \\to A_{r+s - d, \\overline p + \\overline q}(X) \\ \\ ?$$\nThis seems difficult to achieve using resolutions. For example, suppose $\\alpha \\in A_{r, \\overline 0}(X)$ and $D \\in A_{d -1, \\overline 0}(X)$. The fibers of $\\widetilde \\alpha \\cap \\pi^{-1}(X^i) \\to \\alpha \\cap X^i$ are typically $(i-1)$-dimensional (the source is typically $(r-1)$-dimensional, and the target is typically $(r-i)$-dimensional). To find a representative of $\\pi_* (\\widetilde \\alpha \\cdot \\widetilde D)$ in $A_{r-1, \\overline 0}(X)$, one would seek a divisor $\\widetilde D_1 \\sim \\widetilde D \\hookrightarrow \\widetilde X$ such that the image of $\\widetilde \\alpha \\cap \\pi^{-1}(X^i) \\cap \\widetilde D_1 \\to \\alpha \\cap X^i$ is not dense. This means $\\widetilde D_1$ misses most fibers of the morphism $\\widetilde \\alpha \\cap \\pi^{-1}(X^i) \\to \\alpha \\cap X^i$. One might try the following technique for moving divisors, at least in the quasi-projective case: find some effective divisor $D$ (with better incidence properties) such that $\\widetilde D + D$ is ample, then use $\\widetilde D_1 = H-D$ for some $H \\in |\\widetilde D + D|$.\nUnfortunately, the divisor $H$, being ample, will meet \\textit{every} fiber of the morphism $\\widetilde \\alpha \\cap \\pi^{-1}(X^i) \\to \\alpha \\cap X^i$ if $i-1 \\geq 1$.\n\n\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\nLet $\\BB^n=\\{z\\in\\CC^n\\ |\\ |z|<1\\}$ be the unit ball in $\\CC^n$ and\ndenote by $Rat(\\BB^n, \\BB^N)$ the set of all proper holomorphic\nrational maps from $\\BB^n$ to $\\BB^N$. We say that $f, g\\in\nRat(\\BB^n, \\BB^N)$ are {\\it holomorphically equivalent} (or {\\it\nequivalent}, for short) if there are $\\sigma\\in Aut(\\BB^n)$ and\n$\\tau\\in Aut(\\BB^N)$ such that $f=\\tau\\circ g\\circ \\sigma$. By a\nwell-known result of Cima-Suffridge \\cite{CS}, $F$ extends\nholomorphically across the boundary $\\p {\\BB}^n$.\n\nFor the equal dimensional case $N=n$, Alexander \\cite{A77} proved\nthat $Rat(\\BB^n, \\BB^n)$ must be automorphisms for $n>1$.\nSubsequently, much effort has been paid to the classification of\n$Rat(\\BB^n, \\BB^N)$ with $N>n$. When $N \/ n$ is not too large, it\nturns out that the maps are equivalent to relatively simple ones. In\nfact, the classification problem had been done for $N\\leq 3n-3$ and\nthe maps turn out to be all monomial maps. The systematic\ninvestigations on the precise classification of $Rat(\\BB^n, \\BB^N)$\ncan be found in the work of\n\\cite{F82,Hu99,HJ01,Ha05,HJX06,CJY18,JY18}, etc. In \\cite{FHJZ10},\nFaran-Huang-Ji-Zhang constructed a family of maps in $Rat(\\BB^n,\n\\BB^{3n-2})$, which cannot be equivalent to any polynomial maps.\nThis indicates\nthat the maps could be quite\ncomplicated when $N\\geq 3n-2$.\n\n\\medspace\n\nTo study maps in $Rat(\\BB^n, \\BB^N)$ , there are two geometric\nproblems which are of fundamental importance. The first one is the\nD'Angelo conjecture.\nFor any rational holomorphic map $H=\\frac{(P_1, ..., P_m)}{Q}$ on\n$\\CC^n$ where $P_j, Q$ are holomorphic polynomials with $(P_1, ...,\nP_m, Q)=1$, the {\\it degree} of $H$ is defined, as in algebraic\ngeometry, to be $deg(H):=\\max\\{\\deg(P_j), deg(Q), 1\\le j\\le m\\}$.\nThe D'Angelo conjecture is as follows: For any $F\\in Rat(\\BB^n,\n\\BB^N)$, it should have\n\\[\ndeg(F) \\le \\begin{cases}\n2N-3,\\ & if\\ n=2,\\\\\n\\frac{N-1}{n-1},\\ & if\\ n\\ge 3.\n\\end{cases}\n\\]\nThere are several partial results supporting this conjecture. The\nconjecture is true for all monomial maps, as demonstrated by\nD'Angelo-Kos-Riehl \\cite{DKR03} for the case $n=2$ and\nby Lebl-Peter \\cite{LP12} for the case $n \\ge 3$. If $F$ is a\nrational map with geometric rank one, this conjecture was proved in\n\\cite[corollary 1.3]{HJX06}. If the conjecture is proved, it would\nbe sharp due to the known examples. Also, it is proved that\n$deg(F)\\le \\frac{N(N-1)}{2(2n-3)}$ holds for any $F\\in Rat(\\BB^n,\n\\BB^N)$ with $n=2$ in \\cite{Me06} and with $n\\ge 2$ in \\cite{DL09}.\n\n\\medskip\n\nAnother geometric problem is the gap conjecture.\nFor any integer $k$ with $1\\le k\\le K(n)$ where $K(n):=\\max\\{t\\in\n\\ZZ^+\\ |\\ \\frac{t(t+1)}{2}2$. The second gap\ninterval is ${\\cal I}_2=(2n, 3n-3)$. When $N \\le 3n-3$ (and $n\\ge\n4$), we know from \\cite{AHJY15} that $deg(F)\\le 2$. These results\nconfirm the D'Angelo conjecture for the first and the second gap\nintervals.\n\n\\medskip\n\nThe third gap interval is ${\\cal I}_3=(3n, 4n-6)$. If D'Angelo\nconjecture is true, we would have $deg(F)\\le 3$ for any $F\\in\nRat(\\BB^n, \\BB^{4n-6})$ because $deg(F)\\le\n\\frac{4n-6-1}{n-1}=4-\\frac{3}{n-1}$. This is confirmed by our main\nresult of this paper as follows.\n\n\\medskip\n\n\\begin{thm}\\label{mainthm}\n If $F\\in Rat(\\BB^n, \\BB^{4n-6})$ with $n \\ge 7$, then $deg(F)\\le 3$.\n\\end{thm}\n\nThe rest of the paper is organized as follows. In Section 2, we\nintroduced some known properties for Rat$(\\HH_n,\\HH_N)$, especially\nfor maps of geometric rank $2$. Section 3 was devoted to the proof\nof our main theorem assuming Proposition \\ref{propdeg}. In Sections\n4-7, we gave a detailed proof of Proposition \\ref{propdeg} according\nto four different cases.\n\n\n\\medspace\n\n\n\\medskip\n\n\\section{Preliminaries}\n\n\n\nLet $\\HH_n=\\{(z, w)\\in\\CC^{n-1}\\times \\CC\\ |\\ \\text{Im}(w)>|z|^2\\}$\nbe the Siegel upper half space and denote by $Rat(\\HH_n, \\HH_N)$ the\nset of all proper holomorphic rational maps from $\\HH_n$ to $\\HH_N$.\nBy the Cayley transform, we can identify $\\BB^n$ with $\\HH_n$ and\nidentify $Rat(\\BB^n,\\BB^N)$ with $Rat(\\HH_n, \\HH_N)$. In what\nfollows, we will prove Theorem \\ref{mainthm} through the properties\nof $Rat(\\HH_n, \\HH_N)$.\n\n\n\nLet $F=(f,\\phi,g)=(\\widetilde{f}, g)= (f_1,\\cdots$, $f_{n-1}$,\n$\\phi_1,\\cdots$, $\\phi_{N-n},g)\\in Rat(\\HH_n, \\HH_{N})$. For each\n$p\\in \\p\\HH_n$, define $\\sigma^0_p\\in \\hbox{Aut}(\\HH_n)$ and\n$\\tau^F_p\\in\\hbox{Aut}(\\HH_N)$ as follows:\n\\begin{equation*}\\begin{split}\n&\\sigma^0_p(z,w)=(z+z_0, w+w_0+2i \\langle z,\\overline{z_0}\n\\rangle),\\\\\n&\\tau^F_p(z^*,w^*)=(z^*-\\widetilde{f}(z_0,w_0),w^*-\\overline{g(z_0,w_0)}-\n2i \\langle z^*,\\overline{\\widetilde{f}(z_0,w_0)} \\rangle).\n\\end{split}\\end{equation*}\n Then\n$F$ is equivalent to $F_p:=\\tau^F_p\\circ F\\circ\n\\sigma^0_p=(f_p,\\phi_p,g_p)$ and $F_p(0)=0$. In \\cite{Hu99}, Huang\nconstructed an automorphism $\\tau^{**}_p\\in\n {Aut}_0({\\HH}_N)$ such that\n$F_{p}^{**}:=\\tau^{**}_p\\circ F_p$ satisfies the following\nnormalization:\n\\[\nf^{**}_{p}=z+{\\frac{i}{ 2}}a^{**(1)}_{p}(z)w+o_{wt}(3),\\ \\phi_p^{**}\n={\\phi_p^{**}}^{(2)}(z)+o_{wt}(2), \\ g^{**}_{p}=w+o_{wt}(4).\n\\]\n\nWrite\n$\\mathcal{A}(p):=-2i(\\frac{\\partial^2(f_p)^{\\ast\\ast}_l}{\\partial\nz_j\n \\partial w}|_0)_{1\\leq j,l\\leq n-1}$. The {\\it geometric rank} of $F$ at\n$p$ is defined to be the rank of the $(n-1)\\times (n-1)$ matrix\n$\\mathcal{A}(p)$, which is denoted by $Rk_F(p)$. Now we define the\n{\\it geometric rank} of $F$ to be $\\kappa_0(F)=max_{p\\in\n\\partial\\HH_n} Rk_F(p)$. \\medspace\n\nWhen a map in $ Rat({\\HH}_n,{\\HH}_N)$ is not of full rank (i.e.,\n$\\kappa_0\\le n-2$), by the works of \\cite{Hu03} and \\cite{HJX06},\nit can further be normalized to the following form:\n\\begin{thm}\n \\label{normalize **}\n Suppose that $F\\in Rat({\\HH}_n,{\\HH}_N)$\n has geometric rank $1\\le\\kappa_0\\le n-2$ with $F(0)=0$. Then there are\n $\\sigma\\in \\hbox{Aut}({\\HH}_n)$ and\n $\\tau\\in \\hbox{Aut}({\\HH}_N)$ such that\n $\\tau\\circ F\\circ \\sigma $ takes\n the following form, which is still denoted by $F=(f,\\phi,g)$ for\n convenience of notation:\n\n \n \\begin{equation}\n \\left\\{\n \\begin{array}{l}\n f_l=\\sum_{j=1}^{\\kappa_0}z_jf_{lj}^*(z,w),\\ l\\le\\kappa_0,\\\\\n f_j=z_j,\\ \\text{for} \\ \\kappa_0+1\\leq j\\leq n-1,\\\\\n \\phi_{lk}=\\mu_{lk}z_lz_k+\\sum_{j=1}^{\\kappa_0}z_j\\phi^*_{lkj}\\ \\text{for\n } \\ \\ (l,k)\\in {\\cal S}_0,\\\\\n \\phi_{lk}=O_{wt}(3),\\ \\ (l,k)\\in {\\cal S}_1,\\\\\n g=w;\\\\\n f_{lj}^*(z,w)=\\delta_l^j+\\frac{i\\delta_{l}^j\\mu_l}{2}w+b_{lj}^{(1)}(z)w+O_{wt}(4),\\\\\n \\phi^*_{lkj}(z,w)=O_{wt}(2),\\ \\ (l,k)\\in {\\cal S}_0,\\\\\n \\phi_{lk}=\\sum_{j=1}^{\\kappa_0}z_j\\phi_{lkj}^*=O_{wt}(3)\\ \\ for\\\n (l,k)\\in {\\cal S}_1.\n \\end{array}\\right.\n \\label{eqn:hao}\n \\end{equation}\n Here, for $1\\le \\kappa_0\\le n-2$, we write ${\\cal S} ={\\cal S}_0\\cup\n {\\cal S}_1$, the index set for all components of $\\phi$, where\n ${\\cal S}_{0}=\\{(j,l): 1\\le j\\leq \\kappa_0, 1\\leq l\\leq n-1, j\\leq\n l\\}$, $ {\\cal S}_1=\\Big\\{(j, l): j=\\kappa_0+1, \\kappa_0+1\\le l \\le \\kappa_0 +\n N-n-\\frac{(2n-\\kappa_0-1)\\kappa_0}{2} \\Big\\}$, and\n \\begin{equation}\n \\label{mui and mujk} \\mu_{jl}=\\begin{cases}\\sqrt{\\mu_j+\\mu_l} &\\\n for\\ j