{"text":"\\section{Introduction}\nThe interaction of protostellar outflows with the ambient molecular cloud occurs through radiative\nshocks that compress and heat the gas, which in turn cools down through line emission at\ndifferent wavelengths. In the dense medium where the\nstill very embedded protostars (the so called class 0 sources) are located, shocks are primarily non-dissociative, and hence the cooling is mainly through emission from \nabundant molecules. Molecular hydrogen is by far the most abundant species in these environments, \nand although H$_2$\\, emits only through quadrupole transitions with low radiative rates,\nit represents the main gas coolant in flows from young protostars. H$_2$\\, shocked emission in outflows has been widely studied in the past mainly through its\nro-vibrational emission in the near-IR (e.g. Eisl{\\\"o}ffel et al. 2000; Giannini et al. 2004; Caratti o Garatti et al. 2006) that traces the dense\ngas at T$\\sim$2000-4000 K. Most of the thermal energy associated with the shocks is however radiated away through\nthe emission of H$_2$\\ rotational transitions of the ground state vibrational level at $\\lambda \\le 28\\mu$m\n(e.g. Kaufman \\& Neufeld 1996). \nMid-IR H$_2$\\, lines are easily excited at low densities and temperatures between 300 and 1500 K: therefore they\nare very good tracers of the molecular shocks associated with the acceleration of ambient gas by\nmatter ejection from the protostar. Given the low excitation temperature, \nthey can also probe regions where H$_2$\\, has not yet reached the ortho-to-para equilibrium, thus \ngiving information on the thermal history of the shocked gas (Neufeld et al. 1998; Wilgenbus et al. 2000).\nIn addition, given the different excitation temperature and critical densities of the v=0--0 and v$\\ge$ 1 H$_2$\\ lines,\nthe combination of mid-IR with near-IR observations is a very powerful tool to constrain \nthe global physical structure and the shock conditions giving rise to the observed emission.\n\nThe study of the 0--0 rotational emission in outflows started in some detail with the \n\\emph{Infrared Space Observatory}. Thanks to the observations performed with the SWS and ISOCAM instruments,\nthe shock conditions, the ortho-para ratio and the global H$_2$\\, cooling have been derived in a handful of \nflows (e.g. Neufeld et al. 1998; Nisini et al. 2000; Molinari et al. 2000; Lefloch et al. 2003). More recently, the \\emph{Infrared Spectrometer} (Houck et al. 2004) on board\n\\emph{Spitzer}, with its enhanced spatial resolution and sensitivity with respect to the ISO spectrometers, \nhas been used to obtain detailed images of the H$_2$\\, rotational emission, from S(0) to S(7), of several\noutflows, from which maps of important physical parameters, such as temperature, column density and o\/p\nratio, have been constructed (Neufeld et al. 2006; Maret et al. 2009; Dionatos et al. 2010). \nIn this framework, Neufeld et al. (2009, hereafter N09) have recently presented IRS spectroscopic maps observations of five young \nprotostellar outflows at wavelengths between 5.2 and 37$\\mu$m, and discussed their averaged physical properties \nand overall energetics. In all the flows, the H$_2$\\, S(0)-S(7) emission has been detected and contributes to more than 95\\% of the \ntotal line luminosity in the 5.2-37$\\mu$m\\, range, while atomic emission, in the form of FeII and SI fine structure lines, accounts for only the remaining $\\sim$5\\%. \n\nIn the present paper, we will analyse the H$_2$\\, line maps obtained by N09 towards the L1157 outflow,\nwith the aim of deriving the main physical conditions pertaining to the molecular gas and their variations \nwithin the flow. This in turn will give information on the thermal history of the flow and on how \nenergy is progressively transferred from the primary ejection event to the slow moving ambient gas. \n\nFor this first detailed analysis, L1157 has been chosen among the sources observed by N09 given its uniqueness as a very active and well studied flow at different wavelengths. \nMore than 20 different chemical species have been indeed \ndetected in the shocked spots of this object (Bachiller \\& Perez Gutierrez 1997; Benedettini et al. 2007), some of them for the first time in outflows (e.g. HNCO, Rodr{\\'{\\i}}guez-Fern{\\'a}ndez et al. 2010, and complex organic molecules, Arce et al. 2008, Codella et al. 2009)\n, testifying for a rich shock induced chemistry. Warm H$_2$\\ shocked emission in L1157\nis also evidenced through near-IR maps (e.g. Davis \\& Eisl\\\"offel, 1995) and Spitzer-IRAC\nimages (Looney et al. 2007). The L1157\noutflow has been also recently investigated with the \\textit{Herschel Space Observatory}, showing\nto be very strong also at far-IR wavelengths (Codella et al. 2010, Lefloch et al. 2010,\nNisini et al. 2010).\n\nThe L1157 outflow extends about 0.7 pc in length. Its distance is uncertain and has been estimated between 250 and 440 pc. Here we will adopt D=440 pc for an easier comparison with other works.\nThe outflow is driven by a highly embedded, low mass class 0 source (L1157-mm or IRAS20386+6751) having $L_{bol} \\sim $ 8.3 $L_\\odot$ (Froebrich 2005). \nIt is a very nice example of an outflow driven by a precessing and pulsed jet, possessing an S-shaped structure and different cavities, whose morphology has been reproduced assuming that the outflow\nis inclined by $\\sim$80$^\\circ$ to the line of sight and the axis of the underlying jet precesses\non a cone of 6$^\\circ$ opening angle (Gueth et al. 1996). The episodic mass ejection events are\nevidenced by the presence, along the flow, of individual clumps that are symmetrically displaced \nwith respect to the central source.\nIt is therefore a very interesting target for a study of the physical conditions pertaining to these active regions through an H$_2$\\ excitation analysis.\n\nThe paper is organized as follow: the observations and the main results are summarized in \\S 2. In \\S 3 \nwe describe the analysis performed on the H$_2$\\ images to derive maps of temperature, column density and \northo-to-para ratio. A more detailed NLTE analysis on individual emission peaks is also presented here, where the \nSpitzer data are combined with near-IR data to further constrain the excitation conditions. The implications of these results for the shock conditions along the L1157 flow are discussed in \\S 4, together with an analysis of the global\nenergy budget in the flow. A brief summary follows in $\\S 5$.\n\n\\section{Observations and results \\label{analysis}}\n\nObservations of the L1157 outflow were obtained in November 2007 with the IRS instrument, during Cycle 4 of the Spitzer mission . \nThe full IRS spectral range (5.2-36.5$\\mu$m) was observed with the Long-High (LH), Short-High (SH) (R $\\sim$ 600) \nand Short-Low (SL) (R between 64 and 128) modules. \nThe L1157 outflow region was covered through 5 individual IRS maps of $\\sim$ 1\\arcmin x1\\arcmin\\, of size each, arranged along the outflow axis. Each map was obtained by stepping the IRS slit by half of its width in the direction perpendicular\nto the slit length. For the SH and LH modules the slit was stepped also parallel to its axis by 4\/5 (SL) and 1\/5 (LH) of its length. \nDetails on the data reduction that generated the individual line maps from the IRS scans are given in N09. The final maps have been resampled to a grid of 2\\arcsec\\, spacing allowing a pixel by pixel comparison of maps obtained with the different IRS modules.\nMaps of the brightest detected lines as well as the full spectrum in a representative position are shown in Fig.\\,7 and Fig.\\,12 of N09. As regards to H$_2$, all the pure rotational lines of the first vibrational levels, from S(0) to S(7),\n are detected at various intensity along the flow. Here we report, in Tab. \\ref{fluxes}, the H$_2$\\, brightness measured in a 20\\arcsec\\, FWHM Gaussian aperture towards different positions.\n\n Fig\\,1 shows the L1157 maps of the S(1) and S(2) lines while Fig.\\,2 displays the S(5) line with superimposed contours of the CO 2--1 emission from Bachiller et al. 2001.\n In the same figure, a map of the 2.12 $\\mu$m 1--0 S(1) line is also presented. The morphology of the 0--0 S(5) and 1--0 S(1) is very similar, with peaks of mid-IR emission \nlocated at the near-IR knots from A to D, as identified by Davis \\& Eisl\\\"offel (1995). \n When compared with the CO map, the mid-IR H$_2$\\, emission appears to follow the curved chain of clumps (labelled as B0-B1-B2 and R0-R1-R, for the blue-shifted and red-shifted lobes, respectively) that also correspond to peaks of SiO emission, as resolved in interferometric observation by Gueth et al. (1998) and Zhang et al. (2000). \n The L1157 outflow morphology has been suggested to delineate a precessing flow (Gueth et al. 1996), where the H$_2$\\, and SiO peak emission knots follow the location of the actual working surface of the precessing jet and are thus associated with the youngest ejection episodes. \nDiffuse H$_2$\\ emission is also detected in the S(1)-S(2) maps, that delineates the wall of a cavity that connects the central source with both\nthe R0 and B0 clumps. Such a cavity has been recognized in the CO 1-0 interferometric maps and it is likely created by the propagation of large bow-shocks.\nThe S(1)-S(2) maps of Fig.\\,1 show extended emission of H$_2$\\, also in the SE direction (i.e. where the B2 clump is located) and in the eastern edge of the northern lobe, that also follow quite closely the CO morphology: these regions at lower excitation might trace additional cavities created by an older ejection episode of the precessing jet.\n\n\\section{ H$_2$ Analysis }\n\\subsection{LTE 2D analysis of the rotational lines: maps of averaged parameters }\n\\label{sec:maps}\nWe have used the H$_2$\\, line maps to obtain the 2D distribution of basic H$_2$\\, physical parameters, through the analysis \nof the rotational diagrams in each individual pixel. As described in N09, who analysed the global \nH$_2$\\, excitation conditions in L1157, the distribution of upper level column densities of the S(0)-S(7) lines as a function of their \nexcitation energy, does not follow a straight line, indicating that a temperature stratification in the \nobserved medium exists. The exact form of this temperature stratification depends on the type of shock\nthe H$_2$\\, lines are tracing, as they probe the post-shock regions where the gas cools from $\\sim $ 1000 K \nto $ \\sim $ 200 K. \n\nThe simplest way to parametrize the post-shock temperature stratification\nis to assume a power-law distribution: this approach was applied by \nNeufeld \\& Yuan (2008), which also show that this type of distribution is expected \nin gas ahead of unresolved bow-shocks. On this basis, and following also N09, we fit the observations \nassuming a slab of gas where the H$_2$\\, column density in each layer at a given $T$ varies as \n$ dN \\propto T^{-\\beta}dT$.\n\nThis law is integrated, to find the total column density, between a minimum ($T_{min} $) and a maximum ($T_{max}$) temperatures . For our calculations \nwe have kept $T_{max}$ fixed at 4000 K, since gas at temperatures larger than this value is not expected to contribute to the emission of the observed pure rotational lines. $T_{min} $ was instead assumed to be equal, in each position, to the minimum temperature probed by the observed lines. This $T_{min} $ is taken as the excitation temperature giving\nrise to the observed ratio of the S(0) and S(1) column densities, assuming a Boltzman\ndistribution. \nThe $T_{min} $ value ranges between $\\sim$150 and 400 K.\n\nWe found that the approach of a variable $T_{min}$\nproduces always fits with a better $\\chi^2$ than assuming a fixed low value in all positions.\nWe also assume the gas is in LTE conditions. Critical densities of rotational lines from S(0) to S(7)\nrange between 4.9 cm$^{-3}$\\, (S(0)) and 4.4$\\times$10$^5$cm$^{-3}$(S(7)) at T=1000 K assuming only H$_2$\\, collisions\n(Le Bourlot et~al. 1999): critical densities decrease if collisions with H and He are not negligible. \nDeviations from LTE can be therefore expected only for the high-$J$ S(6) and S(7) transitions: the S\/N of these \ntransitions in the individual pixels is however not high enough to disentangle, in the rotational diagrams, NLTE \neffects from the effects caused by the variations of the other considered parameters. In particular, as also\ndiscussed in N09, there is a certain degree of degeneracy in the density and the $\\beta$ parameter \nof the temperature power law in a NLTE treatment that we are not able to remove in the analysis of the\nindividual pixels. This issue will be further discussed in \\S \\ref{sec:NLTE}. \nAn additional parameter of our fit is the ortho-to-para ratio (OPR) value. It is indeed recognized that the 0--0 H$_2$\\, lines are\noften far from being in ortho-to-para equilibrium, an effect that in a rotational diagram is evidenced by\na characteristic zigzag behavior in which column densities of lines with even-$J$ lie systematically above those of odd-$J$ lines. In order not to introduce too many parameters, we assume a single OPR value as a free parameter for the\nfit. In reality, the OPR value is temperature dependent (e.g. N09 and \\S \\ref{sec:NLTE}), and therefore\nthe high-$J$ lines might present an OPR value closer to equilibrium then the low-$J$ transitions. \nOur fit gives therefore only a value averaged over the temperature range probed by the lines considered (i.e. $\\sim$200-1500 K). \n\nIn summary, we have varied only three parameters, namely the total H$_2$\\, column density N(H$_2$), the OPR, and the temperature power law index $\\beta$, in order to obtain the best model fit through a $\\chi^2$ minimization procedure and assuming a 20$\\%$ flux uncertainty for all the lines. \nThe fit was performed only in those pixels where at least four lines with an S\/N larger\nthan 3 have been detected. \nBefore performing the fit, the line column densities were corrected for extinction assuming $A_v$ = 2 (Caratti o Garatti et al. 2006) and \nadopting the Rieke \\& Lebofsky (1985) extinction law. At the considered wavelengths,\nvariations of $A_v$ of the order of 1-2mag do not affect any of the derived results.\n\nFigure \\ref{fig:ncol} shows the derived map of the H$_2$\\, column density, while in Fig. \\ref{fig:b_tmin} and \\ref{fig:op_tcold} \nmaps of the OPR and $\\beta$ are displayed together with temperature maps relative to the ''cold'' and ''warm'' components ($T_{cold}$ and $T_{warm}$),\ni.e. the temperature derived from linearly fitting the S(0)-S(1)-S(2) and S(5)-S(6)-S(7) lines, once corrected for the derived OPR value. \nIn Fig. 6 we also show the individual excitation diagrams for selected positions along the flow, obtained from intensities\nmeasured in a 20$\"$ FWHM Gaussian aperture centered towards emission peaks (Tab. 1). Values of the\nfitted parameters in these positions are reported in Tab. \\ref{param}. In addition to $T_{cold}$ and $T_{warm}$,\nwe give in this table also the values of the average temperature in each knot, derived through a\nlinear fit through all the H$_2$\\ lines ($T_{med}$).\n\nThe maps show significant variations in the inferred parameters along the outflow. \nThe H$_2$\\ column density ranges between 5$\\times$10$ ^{19} $ and \n3$\\times$10$ ^{20} $cm$^{-3}$. The region at the highest column density is located towards the B1 molecular bullet (see Fig. \\ref{fig:h2co})\\footnote{Fig. \\ref{fig:h2co} shows that the molecular clumps B1, R0 and\nR coincide in position with the NIR H$_2$\\ knots A, C and D. In the paper, both nomenclature\nwill be used specifying if we refer to the NIR or mm condensations}. \nThis is consistent with the higher column density of CO found in B1 with respect \nto other positions in the blue lobe (Bachiller \\& Perez Gutierrez 1997), and might suggest that this is a zone where the outflowing gas is compressed due \nto the impact with a region of higher density (Nisini et al. 2007). Towards the NW, red-shifted \noutflow, the column density has a more uniform distribution, with a plateau at $ \\sim 10 ^{20} $ cm$^{-2}$\\ that follows the H$_2$\\ intensity distribution. \nThe N(H$_2$) decreases at the apex of the red-shifted outflow, with a value slightly below $ 10 ^{20} $ cm$^{-2}$\\ at the position\nof the D near-IR knot.\n\n$T_{cold}$ ranges between $ \\sim $ 250 and 550 K. The highest values are found at the tip of the northern outflow lobe, while local maxima corresponds to the positions of line intensity peaks.\n$T_{warm}$ ranges between $ \\sim $ 1000 and 1500 K. In this case the highest values are in the southern lobe, at the position of the A NIR knot. \nAs a general trend, the $T_{warm}$ value decreases going from the southern to the northern peaks of emission, with\nthe minimum value at the position of the D NIR knot. \n\n$\\beta$ values range between $ \\sim $4-4.5 in the blue-shifted lobe while it is larger in the red-shifted\nlobe, with maximum values of $ \\sim $ 5.5 at the tip of the flow. \nDue to the degeneracy between $\\beta$ and density discussed in the previous section, these\nvalues can be considered as upper limits because of our assumption of LTE conditions. \nNeufeld \\& Yuan (2008) have discussed the $\\beta$ index expectations in bow-shock excitation.\nA $\\beta$ index of $ \\sim $ 3.8 is expected in paraboloid bow shocks having a velocity at the bow apex\nhigh enough to dissociate H$_2$, in which case the temperature distribution\nextends to the maximum allowed temperature. Slower shocks that are not able to attain the maximum\ntemperature, produce steeper temperature distributions, i.e. with values of $ \\beta $ greater than 3.8.\nThis is consistent with our findings: low values of $ \\beta $ (of the order of 4) are found in the blue-shifted\nlobe: here, evidence of H$_2$\\ dissociation is given by the detection\nof atomic lines (i.e. [SiII], [FeII] and [SI]) in the IRS spectrum. In the red-shifted lobe, where values of $\\beta$ larger than 4 are derived in the LTE\nassumption, no atomic emission is detected and the $T_{warm}$ values\nare lower than those measured in the blue-shifted lobe, indicating a maximum temperature lower than in the blue flow .\n\nThe OPR varies significantly along the outflow, spanning from $ \\sim $ 0.6 to 2.8. Hence it is always below the equilibrium value of 3. Although a one-to-one correlation between temperature and OPR cannot be discerned, some trends can be\ninferred from inspection of Fig. \\ref{fig:op_tcold}. In general the OPR minima are observed in plateau regions between two consecutive\nintensity peaks, where also the cold temperature has its lowest values. At variance with this trend, the emission\nfilaments delineating the outflow cavity within $\\pm$ 20$ \\arcsec $ from the mm source, where the cold \ntemperature reaches a minimum value of $ \\sim $ 250 K, show rather high\nvalues of the OPR, $ \\sim $2.4-2.8. This might suggests that this region has experienced an older \nshock event that has raised the OPR, though not to the equilibrium value, and where the gas\nhad time to cool at a temperature close to the pre-shock gas temperature. \nOn the other hand, at the apex of both the blue- and red-shifted lobes, where the cold temperatures \nare relatively high (i.e. 500-550 K), the OPR is rather low, $ 1.5-2.0 $. Evidence of regions of low OPR and high \ntemperatures at the outflow tips has been already given in other flows (Neufeld et al. 2006; Maret et al. 2009). It has been suggested that these represent zones subject to recent shocks where the OPR has not had time yet to reach the\nequilibrium value. \n\n\n\\subsection{ NLTE analysis: constraints on H and H$_2$ particle density}\n\\label{sec:NLTE}\n\nAdditional constraints on the physical conditions responsible for the H$_2$\\, excitation, are provided by combining\nthe emission of the mid-IR H$_2$\\ pure-rotational lines from the ground vibrational level with the emission from near-IR ro-vibrational lines.\nIt can be seen from Fig. \\ref{fig:h2co} that the 2.12$\\mu$m\\, emission follows quite closely the\nemission of the 0--0 lines at higher $J$. In addition to the 2.12$\\mu$m\\, data presented in Fig. \\ref{fig:h2co}, we have also considered\nthe NIR long-slit spectra obtained on the A and C NIR knots by Caratti o Garatti (2006). These knots, at the spatial\nresolution of the NIR observations (i.e. $\\sim$0.8\\arcsec), are separated in several different sub-structures that have been individually investigated with the long-slit spectroscopic observations. For our analysis we have considered the\ndata obtained on the brightest of the sub-structures, that coincide, in position, \nwith peaks of the 1--0 S(1) line.\n\nIn order to inter-calibrate in flux the Spitzer data with these NIR long-slit data, obtained with a slit-width of 0.5 arcsec, \nwe have proceeded as follows: we first convolved the \n2.12$\\mu$m\\, image at the resolution of the Spitzer images and than performed photometry on the\nA and C peaks positions with a 20\\arcsec\\, diameter Gaussian aperture, i.e. with the same\naperture adopted for the brightness given in Tab. \\ref{fluxes}. We have then scaled the fluxes of the individual lines given in \nCaratti o Garatti (2006) in order to match the 2.12$\\mu$m\\ flux gathered in the slit with that measured by the\nimage photometry. In doing this, we assumed that the average excitation conditions within the 20$\\arcsec$ aperture are\nnot very different from those of the A-C peaks. This assumption is \nobservationally supported by the fact that the ratios of different H$_2$\\ NIR lines \ndo not change significantly (i.e. less than 20\\%) in the A-C knot substructures separately\ninvestigated in Caratti o Garatti (2006).\nWe have considered only those lines detected with S\/N larger \nthan 5; in practice this means considering lines from the first four and three vibrational levels for knots A and C, respectively. \nThe excitation diagrams obtained by combining the Spitzer and NIR data for these two knots are displayed in Fig.\\ref{fig:ACfit}.\n\nIn order to model together the 0--0 lines and the near-IR ro-vibrational lines, we have implemented\n two modifications to the approach adopted previously. First of all, the NIR lines probe gas\nat temperatures higher than the pure rotational lines, of few thousands of K, at which values it is expected\nthat the OPR has already reached equilibrium. Thus, \nthe ortho-para conversion time as a function of temperature needs to be included in the\nfitting procedure, since lines excited at different temperatures have different OPRs.\nWe have here adopted the approach of N09 and used an analytical expression for the OPR as a function of the temperature, considering a gas that had an initial value of the ortho-to-para ratio OPR$_0 $ and has been heated to a temperature T for a time $ \\tau$. Assuming that the para-to-ortho conversion occurs through reactive collisions with atomic hydrogen, we have:\n\n\\begin{equation}\n\\frac{OPR(\\tau)}{1+OPR(\\tau)} = \\frac{OPR_0}{1+OPR_0}\\,e^{-n(H)k\\tau} + {\\frac{OPR_{LTE}}{1+OPR_{LTE}}}\\,\\left( 1 - e^{-n(H)k\\tau}\\right) \n\\end{equation}\n\nIn this expression, n(H) is the number density of atomic hydrogen and OPR$_{LTE}$ is the ortho-to-para ratio equlibrium value.\nThe parameter $ k $ is given by the sum of the rates coefficients for para-to-ortho conversion ($ k_{po} $), estimated\nas 8$ \\times $10$ ^{-11} $exp(-3900\/T) cm$ ^{3} $\\,s$ ^{-1} $, and for ortho-to-para conversion, $ k_{op} \\sim k_{po}\/3 $\n(Schofield et al. 1967). Thus the dependence of the OPR on the temperature is implicitly given by the dependence on T of the\n$ k $ coefficient. The inclusion of a function of OPR on $T$, introduces one additional parameter to our fit: while we have previously \nconsidered only the average OPR of the 0--0 lines, we will fit here the initial OPR$_0 $ value and the coefficient $ K = n(H)\\tau $.\n\nThe second important change that we have introduced with respect to the previous fitting procedure, \nis to include a NLTE treatment of the H$_2$\\, level column densities. In fact, the critical densities\nof the NIR lines are much higher than those of the pure rotational lines (see Le Bourlot et~al. 1999). For example, the \n$n _{crit} $ of the 1-0 S(1) 2.12$\\mu$m\\, line is 10$^7$cm$^{-3}$\\, assuming only collisions with H$_2$\\, and T=2000 K. \nTherefore the previously adopted LTE approximation might not be valid when combining lines from different \nvibrational levels.\nThis is illustrated in Fig.\\ref{fig:plot_dens}, where we plot the results obtained by varying the H$_2$\\ density between \n10$ ^{3} $ and 10$ ^{7} $cm$^{-3}$\\, while keeping the other model parameters fixed . The observed column densities in the A\nposition are displayed for comparison. \nFor the NLTE statistical equilibrium computation we have adopted the H$_2$\\, collisional rate coefficients given by Le Bourlot et~al. (1999) \\footnote{The rate coefficients for collisions with ortho- and para-H$_2$, HI and He, computed in Le Bourlot et~al. (1999) \nare available at the web-site: http:\/\/ccp7.dur.ac.uk\/cooling$\\_$by$\\_$h2\/} .\nThis figure demonstrates the sensitivity of the relative ratios between 0--0 and 1--0 transitions to density variations. For example, in this particular case, the ratio N(H$_2$)$_{0-0 S(7)}$\/N(H$_2$)$_{1-0 S(1)}$ is 64.6 at n(H$_2$)=10$ ^{4} $ cm$^{-3}$\\, and 1.9 at n(H$_2$) $ \\gtrsim $ 10$ ^{7} $ cm$^{-3}$. The figure also shows that the observational points display only a small misalignment in column densities between the 0--0 lines and the 1--0 lines, already indicating that the ro-vibrational lines are close to\nLTE conditions at high density.\n\nIn Fig.\\ref{fig:ACfit}, we show the final best-fit models for the combined mid- and near-IR column densities \nin the A and C positions. \nAs anticipated, the derived n(H$_2$) densities are large, of the order of 10$ ^{7} $ and 6$\\times$10$ ^{6}$ cm$^{-3}$, for the A and C positions, respectively. The two positions indeed show very similar excitation conditions: only the column density is a factor of 3 smaller in knot C. \nHence, we conclude that the lack of detection of rotational lines with v$ > $ 3 in knot C, in contrast to knot A (Caratti o Garatti et al. 2006), is due\npurely to a smaller number of emitting molecules along the line of sight and not to different excitation conditions.\n\nThe derived H$_2$\\, densities are much higher than previous estimates based on other tracers. Nisini et al. (2007)\nderive a density of 4$\\times$10$ ^{5} $ cm$^{-3}$\\, at the position of knot A from multi-line SiO observations, thus more than an order of magnitude smaller than those inferred from our analysis. The high-J CO lines observed along the blue-shifted lobe of L1157 by \nHirano et al. (2001), indicate a density even smaller, of the order of 4$\\times$10$ ^{4} $ cm$^{-3}$. \nSiO is synthesized and excited in a post-shock cooling zone where the maximum compression is reached, therefore\nit should trace post-shock regions at densities higher than H$_2$ (e.g. Gusdorf et al. 2008). \nOne possibility at the origin of the discrepancy is our assumption of collisions with only H$_2$\\ molecules, and thus of a \nnegligible abundance of H. This can be considered roughly true in the case of non-dissociative C-shocks, where H atoms are \nproduced primarely in the chemical reactions that form H$_2$O from O and H$_2$, with an abundance n(H)\/n(H$_2$+H) $ \\sim $ 10$^{-3}$\n(e.g. Kaufmann \\& Neufeld 1996). However, if the shock is partially dissociative, the abundance of H can increase considerably and \ncollisions with atomic hydrogen cannot be neglected, in view of its large efficiency in the H$_2$\\, excitation.\nThis situation cannot be excluded at least for the knot A, where atomic emission from [FeII] and [SI] has been \ndetected in our Spitzer observations.\nSince we cannot introduce the n(H) as an additional independent parameter of our fit, we have fixed n(H$_2$) at the value\nderived from SiO observations (4$\\times$10$ ^{5} $ cm$^{-3}$, Nisini et al. 2007) and varied the H\/H$_2$\\, abundance ratio. \nThe best fit is in this case obtained with a ratio H\/H$_2$=0.3: this indicates that our observational data are\nconsistent with previous H$_2$\\ density determinations only if a large fraction of the gas is in atomic form.\n\nTurning back to the inferred OPR variations with temperature, our fit implies that the OPR in the cold gas component at T=300 K\nis significantly below the equilibrium value, while the value of \nOPR=3 is reached in the hot gas at T=2000 K traced by the NIR lines. The parameter $K=n(H)\\tau$\nis constrained to be $\\sim$10$^6$ and 10$^7$ yr\\,cm$^{-3}$\\, for knots A and C, respectively. \nWe can also estimate the time needed for the gas to reach this distribution of OPR, from the limits on the\natomic hydrogen abundance previously discussed.\nOur data implies a high value of the n(H) density: a minimum value of n(H) $\\sim$0.6-1$\\times$10$^4$cm$^{-3}$, (for knots C and A, respectively) \n is given if we assume H\/H$_2$$\\sim$ 10$^{-3}$ (and thus the n(H$_2$) $\\sim$ 10$ ^{7} $cm$^{-3}$, given\n by our fit), while a maximum value of $\\sim$10$^5$cm$^{-3}$,\nis derived from the fit where n(H$_2$) is kept equal to 4$\\times$10$ ^{5} $ cm$^{-3}$.\nThe high abundance of atomic H ensures that conversion of para- to\northo-H$_2$\\, proceeds very rapidly: the fitted values of the K parameter indeed imply that the observed range of\nOPR as function of temperature have been attained in a timescale between 100 and 1000 yrs for both the knots.\n\nFinally, given the column density and particle density discussed above, we can estimate the H$_2$\\,\ncooling length ($L \\sim$ N(H$_2$)\/n(H$_2$)). If we consider the case of n(H$_2$) $\\sim$ 10$ ^{7} $cm$^{-3}$,\nand negligible n(H), we have $L \\sim$ 10$ ^{13} $ cm while a length of $\\sim$ 10$ ^{15} $ cm is inferred\nin the case of n(H$_2$)$\\sim$ 4$\\times$10$^5$cm$^{-3}$. \nAll the parameters derived from the above analysis are summarised in Tab. \\ref{shock} and they will be discussed\nin the next section in the framework of different shock models.\n\n\\section{Discussion}\n\n\\subsection{Shock conditions giving rise to the H$_2$\\, emission}\n\nThe copious H$_2$\\, emission at low excitation observed along the L1157 outflow \nindicates that the interaction of the flow with the ambient medium occurs\nprevalently through non-dissociative shocks. \nBoth the Spitzer IRS maps of N09, and the NIR narrow band images of Caratti o Garatti et al. 2006, show that significant gas dissociation in L1157 occurs only at the A peak, where both mid-IR \nlines from [FeII], [SII] and [SI] and weak [FeII] at 1.64$\\mu$m\\ have been detected. \nWeaker [SI]\\,25$\\mu$m\\ and [FeII]\\,26$\\mu$m\\ emission have been also detected on the C spot, but\noverall the atomic transitions give a negligible contribution to the total gas cooling,\nas pointed out in N09.\nThese considerations suggest that most of the shocks along the outflow occur at speeds\n below $\\sim$ 40km\\,s$^{-1}$, as this is the velocity limit above which H$_2$\\ is expected to be dissociated.\nThe knot A is the only one showing a clear bow-shock structure. Here the velocity at\nthe bow apex is probably high, causing H$_2$\\ dissociation and atomic line excitation,\nconsistent with the fitted temperature power law $\\beta$ of $\\sim$ 4, as discussed in \\S 3.1,\nwhile the bulk of the H$_2$\\, emission comes from shocks at lower velocities originating in\nthe bow wings.\n\nConstraints on the shock velocity that gives rise to the molecular emission in L1157\n have been already given in previous works. The sub-mm SiO emission and\nabundances, measured in different outflow spots, suggest shock velocities of the \norder of 20-30km\\,s$^{-1}$\\ (Nisini et al. 2007). The comparison of SiO and H$_2$\\ \nemission against detailed shock models performed by Gusdorf et al. (2008) confirm a similar\nrange of velocities in the NIR-A knot, although the authors could not find a unique \nshock model that well represents both the emissions. \n\nCabrit et al. (1999) found that the column density of the mid-IR H$_2$\\, emission lines, \nfrom S(2) to S(8), observed by ISO-ISOCAM was consistent either with C-shocks having\nvelocities of $\\sim$\\,25km\\,s$^{-1}$\\, or with J-shocks at lower velocity, of the order\nof $\\sim$\\,10km\\,s$^{-1}$. Gusdorf et al. (2008), however, conclude\nthat stationary shock models, either of C- or J-type, are not able to reproduce the observed\nrotational diagram on the NIR-A position, constructed combining ISOCAM data and \nNIR vibrational lines emission. A better fit was obtained \nby these authors considering non-stationary shock models, which have developed a magnetic precursor but which retain a J-type discontinuity (the so-called CJ shocks, Flower et al. 2003). \nSimilar conclusions, but on a different outflow, have been reached by Giannini et al. 2006\nwho studied the H$_2$\\ mid- and near-IR emission in HH54: in general, stady-state C- and J-type \nshocks fail to reproduce simultaneously the column densities of both the ro-vibrational \nand the v=0, pure-rotational H$_2$\\ levels. \n\nA different way to look at the issue of the prevailing shock conditions in the observed\nregions, is to compare the set of physical parameters that we have inferred from our analysis \nto those expected from different shocks. With this aim, we summarize in \nTab. \\ref{shock} the physical properties derived on the A and C H$_2$\\, knots. \nIn addition to the parameters derived from the NLTE analysis \n reported in Section \\ref{analysis}, namely H$_2$ post-shock density, H\/H$_2$\\ fraction, \n initial OPR, cooling length and time, the table reports also the average values of \nOPR and rotational temperature, as they are measured from a simple linear fit\n of the rotational diagrams presented in Fig. \\ref{fig:fit_nir}.\n\n\nAs mentioned in \\S 3.2, the high fraction of atomic hydrogen inferred by our analysis rule out excitation in a pure\n C-shock. In fact, dissociation in C-shocks is always too low to have a \nH\/H$_2$ ratio higher than 5$\\times$10$^{-3}$, irrespective from the shock velocity and magnetic field strength \n(Kaufman \\& Neufeld 1996; Wilgenbus et al. 2000).\nC-shocks are not consistent with the derived parameters even if we consider the model fit with\nthe high H$_2$\\ post-shock density of the order of 10$^{7}$ cm$^{-3}$ and negligible atomic hydrogen: in this case we derive an emission length of 10$^{13}$ cm,\nwhich is much lower than the cooling length expected in C-shocks, which, although \ndecreasing with the pre-shock density, is never less than 10$^{15}$ cm (Neufeld et al. 2006) .\n\nStationary J-shock models better reproduce some of our derived parameters.\nFor example, in J-type shocks the fraction of hydrogen in the post-shocked gas\ncan reach the values of 0.1-0.3 we have inferred, provided that the shock velocity is larger\nthan $\\sim$ 20 km\\,s$^{-1}$. In general, a reasonable agreement with the inferred post-shock density and \nH\/H$_2$\\ ratio is achieved with models having $v_s$=20-25 km\\,s$^{-1}$ and pre-shock densities of 10$^3$cm$^{-3}$\n(Wilgenbus et al. 2000). Such models predict a shock flow time of the order of 100 yr or less, which\nis also in agreement with the value estimated in our analysis at least in knot A.\n In such models, however, the cooling length is an order of magnitude smaller than the \ninferred value of $ \\sim $10$ ^{15} $ cm. In addition, the gas temperature remains high for most of the post-shocked region: the average rotational temperature of \nthe v=0 vibrational level is predicted to be, according to the Wilgenbus et al. (2000)\ngrid of models, always about 1600 K or larger, as compared with the value of about 800-900 K inferred from observations. \nThe consequence of the above inconsistencies is that J-type shocks tend to underestimate \nthe column densities of the lowest H$_2$\\, rotational levels in L1157, an effect already pointed \nout by Gusdorf et al. (2008). \n\nAs mentioned before, Gusdorf et al. (2008) conclude that the H$_2$\\ pure rotational emission in L1157 \nis better fitted with a non-stationary C+J shock model with either $v_s$ between 20 and 25 km\\,s$^{-1}$\\ and pre-shock densities $n_H = 10^4$ cm$^{-3}$, or with $v_s \\sim 15$ km\\,s$^{-1}$\\, and\nhigher pre-shock densities of $n_H = 10^5$ cm$^{-3}$. Such models, however, still underestimate\nthe column densities of the near-IR transitions: the post-shocked H$_2$\\ gas density \nremains lower than the NIR transitions critical density and the atomic hydrogen\nproduced from H$_2$\\ dissociation is not high enough to populate the vibrational\nlevels to equilibrium conditions.\n\nThe difficulty of finding a suitable single model that reproduce the derived physical \nconditions is likely related to possible geometrical effects and to the fact that multiple shocks with different velocities might be present along the line of sight. It would be indeed interesting to explore whether bow-shock models might be able to predict the averaged physical characteristics \nalong the line of sight that we infer from our analysis.\n\n\\subsection{Flow energetics}\n\nH$_2$\\ emission represents one of the main contributor to the energy radiated\naway in shocks along outflows from very young stars. Kaufman \\& Neufeld (1996) predicted that\nbetween 40 and 70\\% of the total shock luminosity is emitted in H$_2$\\, lines\nfor shocks with pre-shock density lower than 10$^5$ cm$^{-3}$\\ and shock velocities larger\nthan 20 km\\,s$^{-1}$, the other main contributions being in CO and H$_2$O rotational emission. \nThis has been also observationally tested by Giannini et al. (2001) who measured the \nrelative contribution of the different species to the outflow cooling in a sample of class 0 \nobjects observed with ISO-LWS. \n\nWe will discuss here the role of the H$_2$\\, cooling in the global radiated energy of the L1157 outflow.\nFrom the best fit model obtained for the knots A and C, we have derived the total, extinction corrected,\nH$_2$\\ luminosity by integrating over all the ro-vibrational transitions considered by our\nmodel. $L_{\\rm H_2}$ is found\nto be 8.4$\\times$10$^{-2}$ and 3.7$\\times$10$^{-2}$ L$_\\odot$ for the A and C knots, respectively. Out of this total\nluminosity, the contribution of only the rotational lines is 5.6$\\times$10$^{-2}$(A) and 2.7$\\times$10$^{-2}$(C) L$_\\odot$,\nwhich means that in both cases they represent about 70\\% of the total H$_2$\\, luminosity.\n\nN09 have found that the total luminosity of the H$_2$\\ rotational lines from S(0) to\nS(7), integrated over the entire L1157 outflow, \namount to 0.15 L$_\\odot$. If we take into account an additional 30\\% of contribution from the v$>$0 \nvibrational levels, we estimate a total H$_2$\\ luminosity of 0.21 L$_\\odot$. This is a 30\\% larger\nthan the total H$_2$\\ luminosity estimated by Caratti o Garatti (2006) in this outflow, \nassuming a single component gas at temperature between 2000 and 3000 K that fit the NIR H$_2$\\ lines. \n\nIf we separately compute the H$_2$\\ luminosity\nin the two outflow lobes, we derive $L_{\\rm H_2}$ = 8.5$\\times$10$^{-2}$ L$_\\odot$ in the blue lobe and 1.3$\\times$10$^{-1}$ L$_\\odot$ in the red\nlobe. Comparing these numbers with those derived in the individual A and C knots, \nwe note that the A knot alone contributes to most of the H$_2$\\ luminosity in the blue lobe. By contrast,\nthe H$_2$\\ luminosity of the red lobe is distributed among several peaks of similar values. \nThis might suggest that most of the energy carried out by the blue-shifted jet is\nreleased when the leading bow-shock encounters a density enhancement at the position of the A knot. \nOn the other hand, the red-shifted gas flows more freely without large density discontinuities, and \nthe corresponding shocks are internal bow-shocks, all with similar luminosities.\n\nThe integrated luminosity radiated by CO, H$_2$O and OI in L1157 has been estimated, through ISO and\nrecent Herschel observations, as $\\sim$0.2 L$_\\odot$ (Giannini et al. 2001, Nisini et al. 2010), \nwhich means that H$_2$\\ alone contributes about 50\\% of the\ntotal luminosity radiated by the outflow. \nIncluding all contributions, the total shock cooling along the L1157 outflow amounts to about 0.4 L$_\\odot$, i.e. $L_{cool}\/L_{bol}$ $\\sim$ 5$\\times$10$^{-2}$, assuming $L_{bol}$=8.4 L$_\\odot$ \nfor L1157-mm (Froebrich 2005). This ratio is consistent with the range of values derived from other class 0 sources \nfrom ISO observations (Nisini et al. 2002).\n\nThe total kinetic energy of the L1157 molecular outflow \nestimated by Bachiller et al. (2001) amounts to 0.2 L$_\\odot$ without any correction for the\noutflow inclination angle, or to 1.2 L$_\\odot$ if an inclination angle of 80 degrees \nis assumed. Considering that the derivation of the L$_{kin}$ value has normally\nan uncertainty of a factor of five (Downes \\& Cabrit 2007), we conclude that the mechanical energy flux \ninto the shock, estimated as $L_{cool}$, is comparable to the kinetic energy of the swept-out \noutflow and thus that the shocks giving rise to the H$_2$ emission have \nenough power to accelerate the molecular outflow.\n\nThe total shock cooling derived above can be also used to infer the momentum flux \nthrough the shock, i.e. $\\dot{P}$ = 2$L_{cool}$\/V$_s$, where V$_s$ is the shock velocity that\nwe can assume, on the basis of the discussion in the previous section, to be of the order of 20 km\\,s$^{-1}$.\nComputing the momentum flux separately for the blue and red outflow lobe, we derive \n$\\dot{P}_{red} \\sim$ 1.7$\\times$10$^{-4}$ and $\\dot{P}_{blue} \\sim$ 1.1$\\times$10$^{-4} $M$_\\odot$ yr$^{-1}$ km\\,s$^{-1}$. \nIn this calculation, we have assumed that the contribution from cooling species\n different from H$_2$, as estimated by ISO and Herschel, is distributed among the two lobes \n in proportion to the H$_2$\\ luminosity. If we assume that the molecular \n outflow is accelerated at the shock front through momentum conservation, then the\n above derived momentum flux should results comparable to the thrust of the outflow,\n derived from the mass, velocity and age measured through CO observations.\nThe momentum flux measured in this way by Bachiller et al. (2001) is 1.1$\\times$10$^{-4}$and 2$\\times$10$^{-4}$ M$_\\odot$ yr$^{-1}$ km\\,s$^{-1}$\nin the blue and red lobes, respectively, i.e. comparable to our derived values. It is interesting to note that \nthe $\\dot{P}$ determination from the shock luminosity confirms the asymmetry between\nthe momentum fluxes derived in two lobes. As shown by Bachiller et al. (2001), the L1157 red lobe \nhas a 30\\% smaller mass with respect to the blue lobe, but a higher momentum flux due to the larger flow velocity. The northern red lobe is in fact more extended than the southern lobe:\nhowever, given the higher velocity of the red-shifted gas, the mean kinematical ages of the two lobes is very\nsimilar. \n\n\n\\section{Conclusions \\label{conclusions}}\n\nWe have analysed the H$_2$\\ pure rotational line emission, from S(0) to S(7),\nalong the outflow driven by the L1157-mm protostar, mapped with the Spitzer - IRS \ninstrument. The data have been analysed assuming a gas temperature stratification where the H$_2$\\ column \ndensity varies as $T^{-\\beta}$ and 2D maps of the H$_2$\\ column density,\northo-to-para ratio (OPR) and temperature spectral index $\\beta$\nhave been constructed. \nFurther constraints on the physical conditions of the shocked gas have been derived \nin two bright emission knots by combining the Spitzer observations with near-IR \ndata of H$_2$\\ ro-vibrational emission. Finally, the global H$_2$\\ radiated energy of the\noutflow has been discussed in comparison with the energy budget of the associated\nCO outflow.\n\nThe main conclusions derived by our analysis are the following:\n\\begin{itemize}\n\\item H$_2$\\ transitions with $J_{lower} \\le$ 2 follows the morphology of the CO molecular\noutflow, with peaks correlated with individual CO clumps and more diffuse\nemission that delineates the CO cavities created by the precessing jet. \nLines with higher $J$ are localized on the shocked peaks, presenting a morphology\nsimilar to that of the H$_2$\\, 2.12$\\mu$m\\, ro-vibrational emission.\n\\item Significant variations of the derived parameters are observed along the flow. \nThe H$_2$\\ column density ranges between 5$\\times$10$^{19} $ and 3$\\times$10$^{20} $cm$^{-2}$: \nthe highest values are found in the blue-shifted lobe, suggesting that here the outflowing\ngas is compressed due to the impact with a high density region.\nGas components in a wide range of temperature values, from $ \\sim $ 250 to $ \\sim $ 1500 K\ncontribute to the H$_2$\\ emission along individual lines of sight. The largest range\nof temperature variations is derived towards the intensity peaks closer to the \ndriving source, while a more uniform temperature distribution, with $ T $\nbetween $ 400 $ and $ 1000$ K, is found at the tip of the northern outflow lobe.\n\\item The OPR is in general lower than the equilibrium value at high temperatures and spans a range from $\\sim$0.6 to 2.8, with the lowest values\nfound in low temperature plateau regions between consecutive intensity peaks. \nAs in previous studies, we also found the presence of regions at low OPR (1.5-1.8) \nbut with relatively high temperatures. These might represent zones subject to recent shocks \nwhere the OPR has not had time yet to reach the equilibrium value.\n\n\\item Additional shock parameters have been derived in the two bright near-IR \nknots A and C, located \nin the blue- and red-shifted outflow lobes, where the mid- and near-IR H$_2$\\ \ndata have been combined. \nThe ratio between mid- and near-IR lines is very sensitive to the molecular plus atomic hydrogen particle density. A high \nabundance of atomic hydrogen (H\/H$_2$ $\\sim$ 0.1-0.3) is implied by the \nthe observed H$_2$\\ column densities if we assume n(H$_2$) values as derived by independent \nmm observations. With this assumption, the cooling lengths of the shock result \nof the order of 7$\\times$10$ ^{14} $ and 10$ ^{15} $ cm for the A and C knot, respectively.\nThe distribution of OPR values as a function of temperature and the \nderived abundance of atomic hydrogen, implies that the shock passing time is of the\norder of 100 yr for knot A and 1000 yr for knot C, given the assumption that the para-to-ortho\nconversion occurs through reactive collisions with atomic hydrogen.\nWe find that planar shock models, either of C- or J-type, are\nnot able to consistently reproduce all the physical parameters derived from our analysis \nof the H$_2$\\ emission. \n\\item Globally, H$_2$\\ emission contributes to about 50\\% of the total shock radiated energy in the L1157 outflow. We find that the momentum flux through the shocks derived from the radiated luminosity is\ncomparable to the thrust of the associated molecular outflow, supporting a scenario\nwhere the working surface of the shocks drives the molecular outflow. \n\\end{itemize}\n\n\\acknowledgments\n\nThis work is based on observations made with the Spitzer Space Telescope, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under a contract with NASA. Financial support from contract ASI I\/016\/07\/0 is acknowledged. \n\n\\bibliographystyle{plainnat}\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} {"text":"\\section{Introduction}\n\nFrames are overcomplete (or redundant) sets of vectors that serve to faithfully represent signals. They were introduced in $1952$ by Duffin and Schaeffer \\cite{duscha52}, and reemerged with the advent of wavelets \\cite{Christensen:2003ab, Dau92,Ehler:2007aa, Ehler:2008ab, hw89}. \nThough the overcompleteness of frames precludes signals from having unique representation in the frame expansions, it is, in fact, the driving force behind the use of frames in signal processing \\cite{Casazza:2003aa, koche1, koche2}.\n\nIn the finite dimensional setting, frames are exactly spanning sets. However, many applications require ``custom-built'' frames that possess additional properties which are dictated by these applications. As a result, the construction of frames with prescribed structures has been actively pursued. For instance, a special class called {\\it finite unit norm tight frames } (FUNTFs) that provide a Parseval-type representation very similar to orthonormal bases, has been customized to model data transmissions \\cite{Casazza:2003aa, Goyal:2001aa}. Since then the characterization and construction of FUNTFs and some of their generalizations have received a lot of attention \\cite{Casazza:2003aa, koche1, koche2}. Beyond their use in applications, FUNTFs are also related to some deep open problems in pure mathematics such as the Kadison-Singer conjecture \\cite{cftw}. FUNTFs appear also in statistics where, for instance, Tyler used them to construct $M$-estimators of multivariate scatter \\cite{Tyler:1987}. We elaborate more on the connection between the $M$-estimators and FUNTFs in Remark~\\ref{remark:M estimator FUNTF}. These $M$-estimators were subsequently used to construct maximum likelihood estimators for the the wrapped Cauchy distribution on the circle in \\cite{KentTaylor1994} and for the angular central Gaussian distribution on the sphere in \\cite{Tyler:1988}. \n\nFUNTFs are exactly the minimizers of a functional called the frame potential \\cite{Benedetto:2003aa}. This was extended to characterize all finite tight frames in \\cite{Waldron:2003aa}. Furthermore, in \\cite{fjko, jbok}, finite tight frames with a convolutional structure, which can be used to model filter banks, have been characterized as minimizers of an appropriate potential. All these potentials are connected to other functionals whose extremals have long been investigated in various settings. We refer to \\cite{cokum06, Delsarte:1977aa, Seidel:2001aa, Venkov:2001aa, Welch:1974aa} for details and related results. \n\nIn the present paper, we study objects beyond both FUNTFs and the frame potential. In fact, we consider a family of functionals, the {\\it $p$-frame potentials}, which are defined on sets $\\{x_{i}\\}_{i=1}^{N}$ of unit vectors in $\\R^d$; see Section~\\ref{section:pfp}. These potentials have been studied in the context of spherical $t$-designs for even integers $p$, cf.~Seidel in \\cite{Seidel:2001aa}, and their minimizers are not just FUNTFs but FUNTFs that inherit additional properties and structure. Common FUNTFs are recovered only for $p=2$. In the process, we extend Seidel's results on spherical $t$-designs in \\cite{Seidel:2001aa} to the entire range of positive real $p$. \n\nIn Section~\\ref{section:estimates}, we give lower estimates on the $p$-frame potentials, and prove that in certain cases their minimizers are FUNTFs, which possess additional properties and structure. In particular, if $0
2$. Finally in Section \\ref{section:intro prob}, we introduce {\\it probabilistic $p$-frames} that generalize the concepts of frames and $p$-frames. We characterize the minimizers of {\\it probabilistic $p$-frame potentials} in terms of probabilistic $p$-frames. The latter problem is solved completely for $0
2}\nLet $\\{x_i\\}_{i=1}^N\\subset S^{d-1}$, $N\\geq d$, and $2
2}\n\\FP_{p, N}(\\{x_{i}\\}_{i=1}^{N}) \\geq N(N-1)\\big(\\frac{N-d}{d(N-1)}\\big)^{p\/2}+N,\n\\end{equation}\n and equality holds if and only if $\\{x_i\\}_{i=1}^N$ is an equiangular FUNTF.\n\\end{proposition}\n\n\n\\begin{proof}\nFor $\\frac{1}{2}=\\frac{1}{p}+\\frac{1}{r}$, H\\\"older's inequality yields\n\\begin{equation}\\label{eq:hoelder numer 1}\n\\|(\\langle x_i,x_j\\rangle)_{i\\neq j}\\|_{\\ell_2} \\leq \\|(\\langle x_i,x_j\\rangle)_{i\\neq j}\\|_{\\ell_p}(N(N-1))^{1\/r}.\n\\end{equation}\nRaising to the $p$-th power and applying $\\frac{1}{r}=\\frac{1}{2}-\\frac{1}{p}$ leads to \n\\begin{equation}\\label{eq:raising to the power}\n\\|(\\langle x_i,x_j\\rangle)_{i\\neq j}\\|_{\\ell_2}^p \\leq \\|(\\langle x_i,x_j\\rangle)_{i\\neq j}\\|_{\\ell_p}^p(N(N-1))^{p\/2-1}.\n\\end{equation}\nTherefore, \n$$\n \\sum_{i\\neq j}|\\langle x_i,x_j\\rangle |^p \\geq \\big(\\sum_{i\\neq j}|\\langle x_i,x_j\\rangle |^2\\big)^{p\/2} (N(N-1))^{1-p\/2}.$$\nUsing the fact that $\\sum_{i\\neq j}|\\langle x_i,x_j\\rangle |^2\\geq \\frac{N^2}{d}-N$ (see Theorem~\\ref{theorem:Benedetto Fickus}) implies that\n$$ \\sum_{i\\neq j}|\\langle x_i,x_j\\rangle |^p \\geq \\big(N(\\frac{N}{d}-1)\\big)^{p\/2} (N(N-1))^{1-p\/2} = N(N-1)\\bigg(\\frac{N-d}{d(N-1)}\\bigg)^{p\/2},$$\nwhich proves~\\eqref{eq:potential for p>2}. \n\nTo establish the last part of the Proposition, we recall that an equiangular FUNTF $\\{x_{k}\\}_{k=1}^{N} \\subset \\R^d$ satisfies \n\\begin{equation}\\label{eq:equi}\n|\\langle x_i,x_j\\rangle | = \\sqrt{\\frac{N-d}{d(N-1)}},\\quad \\text{ for all }i\\neq j\n\\end{equation} \nsee, \\cite{Casazza:2008ab,Sustik:2007aa}, for details. Consequently, if $\\{x_{k}\\}_{k=1}^{N}$ is an equiangular FUNTF, then~\\eqref{eq:potential for p>2} holds with equality. \n\nOn the other hand, if equality holds in \\eqref{eq:potential for p>2}, then $\\sum_{i\\neq j}|\\langle x_i,x_j\\rangle |^2 = \\frac{N^2}{d}-N$ and $\\{x_i\\}_{i=1}^N$ is a FUNTF due to Theorem \\ref{theorem:Benedetto Fickus}. Moreover, the H\\\"older estimate \\eqref{eq:hoelder numer 1} must have been an equality which means that $|\\langle x_i,x_j\\rangle|=C$ for $i\\neq j$, and some constant $C\\geq 0$. Thus, the FUNTF must be equiangular.\n\\end{proof}\n\n\n By comparing \\eqref{eq:Welch} with \\eqref{eq:potential for p>2}, it is easily seen that the Welch bound is not optimal for small $N$: \n \n \n \\begin{proposition}\\label{prop:second one}\n Let $\\{x_i\\}_{i=1}^N\\subset S^{d-1}$ and $p=2k>2$ be an even integer. If $d 2}, so we focus on $p\\in (0,2)$. \n\nOne easily verifies that, for $p_0=\\frac{\\log(\\frac{d(d+1)}{2})}{\\log(d)}$, an orthonormal basis plus one repeated vector and an equiangular FUNTF have the same $p_0$-frame potential $\\FP_{p_{0}, d+1}$. Under the assumption that those two systems are exactly the minimizers of $\\FP_{p_{0}, d+1}$, the next result will give a complete characterization of the minimizers of $\\FP_{p, d+1}$, for $0 1$. According to Proposition \\ref{prop:potential for p>2}, the minimizers of the $p$-frame potential for $2 2} and \\ref{prop:second one} still hold for complex vectors $\\{z_i\\}_{i=1}^N\\subset \\C^d$ that have unit norm. The constraints on $N$ and $d$ that allow for the existence of a complex FUNTF are slightly weaker than in the real case~\\cite{Sustik:2007aa}.\n\\end{remark}\n\n\n\n\n\n\n\n\n\n \n\\section{The probabilistic $p$-frame potential}\\label{section:intro prob}\n\nThe present section is dedicated to introducing a probabilistic version of the previous section. We shall consider probability distributions on the sphere rather than finite point sets. Let $\\mathcal{M}(S^{d-1},\\mathcal{B})$ denote the collection of probability distributions on the sphere with respect to the Borel sigma algebra $\\mathcal{B}$. \n\nWe begin by introducing the probabilistic $p$-frame which generalizes the notion of probabilistic frames introduced in~\\cite{Ehler:2010aa}. \n\n\\begin{definition}\\label{probpframe}\nFor $0 0$ such that\n \\begin{equation}\\label{ppframeineq}\nA\\|y\\|^p\\leq \\int_{S^{d-1}} |\\langle x,y\\rangle|^p d\\mu(x) \\leq B\\|y\\|^p, \\quad\\forall y\\in\\R^d.\n \\end{equation}\n We call $\\mu$ a \\emph{tight probabilistic $p$-frame} if and only if we can choose $A=B$.\n \\end{definition}\n Due to Cauchy-Schwartz, the upper bound $B$ always exists. \nConsequently, in order to check that $\\mu$ is a probabilistic $p$-frame one only needs to focus on the lower bound $A$. \n\n Since the uniform surface measure $\\sigma$ on $S^{d-1}$ is invariant under orthogonal transformations, one can easily check that it constitutes a tight probabilistic $p$-frame, for any $0 2$. Then, for all $y\\neq 0 \\in \\R^d$, \n\\begin{align*}\nA\\|y\\|^p&\\leq \\int_{S^{d-1}} |\\langle x,y\\rangle|^p d\\mu(x)\\\\\n&=\\int_{S^{d-1}} |\\langle x,y\\rangle|^2\\, |\\langle x,y\\rangle|^{p-2}\\, d\\mu(x)\\\\\n& \\leq \\int_{S^{d-1}} \\|x\\|^{p-2}\\, \\|y\\|^{p-2}\\, |\\langle x,y\\rangle|^2 d\\mu(x) \\\\\n&= \\|y\\|^{p-2}\\, \\int_{S^{d-1}} |\\langle x,y\\rangle|^2 d\\mu(x),\n\\end{align*}from which it follows that $$A\\|y\\|^2 \\leq \\int_{S^{d-1}} |\\langle x,y\\rangle|^2 d\\mu(x).$$\n\nIf $\\mu$ is a probabilistic $p$-frame for some $p<2$. Then, for all $y\\neq 0 \\in \\R^d$, \n$$\\|y\\|^{2}=|\\langle Sy,S^{-1}y\\rangle_{\\R^{d}}|=|\\langle F^{*}Fy,S^{-1}y\\rangle_{\\R^{d}}|=|\\langle Fy,FS^{-1}y\\rangle_{L_{p}\\to L_{p'}}|,$$ which can be estimated by \n$$\\|y\\|^{2} \\leq \\|Fy\\|_{L_{p}}\\|FS^{-1}y\\|_{L_{p'}}\\leq C \\|Fy\\|_{L_{2}}\\|y\\|,$$ where we have used the fact that for $p<2$, $L_{2}(S^{d-1}, \\mu) \\subset L_{p}(S^{d-1}, \\mu)$. This conclude the proof of a). \n\n\nb) If $\\mu$ is a probabilistic $p$-frame for some $1\\leq p < \\infty,$ then by a) $\\mu$ is a probabilistic frame. In this case, $\\tilde{\\mu}$ is known to be a probabilistic frame, cf.~\\cite{Ehler:2010aa}, and thus a probabilistic $p$-frame. \n\\end{proof}\n\n\n \n\n\n\nWe are particularly interested in tight probabilistic $p$-frame potentials, which we seek to characterize in terms of minimizers of appropriate potentials. This motivates the following definition: \n\n \\begin{definition}\\label{profframpot}\nFor $0 0$. One can check that the measure $\\nu$ defined by \n\\begin{equation*}\n\\nu(E) := m\\delta_{y_2}(E)-\\mu(E\\cap K),\\quad E\\in\\mathcal{B},\n\\end{equation*}\nsatisfies $\\nu(S^{d-1})=0$, and $\\mu +\\epsilon \\nu \\geq 0$. Hence, $\\PFP(\\mu,\\nu,p)\\geq 0$. On the other hand, we can estimate\n$$\n\\PFP(\\mu,\\nu,p) = \\int_{S^{d-1}} P_\\mu(y)d\\nu(y)= P_\\mu(y_2)m - \\int_K P_\\mu(y)d\\mu(y)= am - \\int_K P_\\mu(y)d\\mu(y)$$ and so\n\n$$\\PFP(\\mu,\\nu,p) \\leq am - (b-\\frac{b-a}{2})m = - \\frac{b-a}{2}m <0.$$\n\nThis is a contradiction to $\\PFP(\\mu,\\nu,p)\\geq 0$ and implies that there is a constant $C$ such that $P_\\mu(y)=C$, for all $y\\in\\supp(\\mu)$. \nWe still have to verify that the constant $C$ is in fact $\\PFP(p)$: \n\\begin{align*}\n\\PFP(p) = \\PFP(\\mu,p) & = \\int_{S^{d-1}} P_\\mu(y)d\\mu(y) \\\\\n& = \\int_{\\supp(\\mu)} P_\\mu(y)d\\mu(y)\\\\\n& = \\int_{\\supp(\\mu)} C d\\mu(y) = C.\n\\end{align*}\n\nThe proof of $(2)$ is similar to the one above, and so we omit it. \n\\end{proof}\nThe following result is an immediate consequence of Proposition~\\ref{prop:1}. \n\n\\begin{corollary}\\label{theorem:tight p frame is necessary}\nLet $0 0$ and $\\delta_\\varepsilon>0$ such that \n \\begin{itemize}\n \\item[(a)] $B_\\varepsilon(v)\\cap B_\\varepsilon(w)=\\emptyset$ and $\\mu(B_\\varepsilon(v)), \\mu(B_\\varepsilon(w)) \\geq \\delta_\\varepsilon$. \n \\item[(b)] for all $x\\in B_\\varepsilon(v)$ and $y\\in B_\\varepsilon(w)$, $|\\langle x,y\\rangle |^p\\geq |\\langle x,y\\rangle |^2+\\varepsilon$.\n \\end{itemize}\nBy using $B=B_\\varepsilon(v)\\times B_\\varepsilon(w)$, this implies\n\\begin{align*}\n\\PFP(\\mu,p) & = \\int_{B} |\\langle x,y\\rangle|^p d\\mu(x) d\\mu(y) + \\int_{S^{d-1}\\times S^{d-1}\\setminus B } |\\langle x,y\\rangle|^p d\\mu(x) d\\mu(y)\\\\\n& \\geq \\int_{B} (|\\langle x,y\\rangle|^2+\\varepsilon) d\\mu(x) d\\mu(y) + \\int_{S^{d-1}\\times S^{d-1}\\setminus B } |\\langle x,y\\rangle|^2 d\\mu(x) d\\mu(y)\\\\\n& = \\PFP(\\mu,2) + \\varepsilon \\mu(B_\\varepsilon(v)) \\mu(B_\\varepsilon(w))\\\\\n&\\geq \\PFP(\\mu,2) +\\varepsilon \\delta_\\varepsilon^2 > \\PFP(\\mu,2),\n\\end{align*}\n which is a contradiction. Thus, we have verified that $|\\langle x,y\\rangle|\\in \\{0,1\\}$, for all $x,y\\in \\supp(\\mu)$. Distinct elements in $\\supp(\\mu)$ are then either orthogonal to each other or antipodes. According to Corollary \\ref{theorem:tight p frame is necessary}, $\\supp(\\mu)$ is complete in $\\R^d$. Thus, there must be an orthonormal basis $\\{x_i\\}_{i=1}^d$ such that\n \\begin{equation*}\n \\{x_1,\\ldots,x_d\\} \\subset \\supp(\\mu) \\subset \\{\\pm x_1,\\ldots,\\pm x_d\\}.\n \\end{equation*} \nConsequently, there is a density $f:S^{d-1}\\rightarrow\\R$ that vanishes on $S^{d-1}\\setminus \\supp(\\mu)$ such that $\\mu(x)=f(x)\\nu_{\\pm x_1,\\ldots,\\pm x_d}(x)$. \n \nTo verify that $f$ satisfies (ii), let us define $\\tilde{f}:S^{d-1}\\rightarrow \\R$ by \n\\begin{equation*}\n\\tilde{f}(x)=\\begin{cases} f(x)+f(-x),& x\\in\\{x_1,\\ldots,x_d\\}\\\\\n0,& \\text{ otherwise. } \n\\end{cases}\n\\end{equation*}\nThis implies that $\\tilde{\\mu}(x)=\\tilde{f}(x)\\nu_{x_1,\\ldots,x_d}(x)$ is also a minimizer of $\\PFP(\\cdot,2)$. But the minimizers of the probabilistic frame potential for $p=2$ have been investigated in~\\cite[Section 3]{Ehler:2010aa}. We can follow the arguments given there to obtain $\\tilde{f}(x_i)=\\frac{1}{d}$, for all $i=1,\\ldots,d$. \n\\end{proof}\n\n\n \n For even integers $p$, we can give the minimum of $\\PFP(\\mu, p)$ and characterize its minimizers. The following theorem generalizes Theorem \\ref{theorem:p even integer discrete}. Moreover, note that the bounds are now sharp, i.e., for any even integer $p$, there is a probabilistic tight $p$-frame: \n \n \\begin{theorem}\\label{theorem:p even integer}\n Let $p$ be an even integer. For any probability distribution $\\mu$ on $S^{d-1}$, \n \\begin{equation*}\n \\PFP(\\mu, p)=\\int_{S^{d-1}}\\int_{S^{d-1}} |\\langle x,y\\rangle|^p d\\mu(x) d\\mu(y) \\geq \\frac{1\\cdot 3\\cdot 5\\cdots(p-1)}{d(d+2)\\cdots (d+p-2) },\n \\end{equation*}\nand equality holds if and only if $\\mu$ is a probabilistic tight $p$-frame. \n \\end{theorem}\n \n\n\n\n \\begin{proof}\n Let $\\alpha=\\frac{d}{2}-1$ and consider the Gegenbauer polynomials $\\{C_{n}^{\\alpha}\\}_{n\\geq 0}$ defined by \n \\begin{equation*}\n C_0^\\alpha(x) = 1, \\qquad C_1^\\alpha(x) = 2 \\alpha x,\n \\end{equation*}\n \\begin{align*}\nC_{n}^\\alpha(x) &= \\frac{1}{n}[2x(n+\\alpha-1)C_{n-1}^\\alpha(x) - (n+2\\alpha-2)C_{n-2}^\\alpha(x)]\\\\\n&= C_n^{(\\alpha)}(z)=\\sum_{k=0}^{\\lfloor n\/2\\rfloor} (-1)^k\\frac{\\Gamma(n-k+\\alpha)}{\\Gamma(\\alpha)k!(n-2k)!}(2z)^{n-2k}.\n\\end{align*}\n$\\{C_{n}^{(\\alpha)}\\}_{n=1}^s$ is an orthogonal basis for the collection of polynomials of degree less or equal to $s$ on the interval $[-1,1]$ with respect to the weight\n\\begin{equation*}\nw(z) = \\left(1-z^2\\right)^{\\alpha-\\frac{1}{2}},\n\\end{equation*} \ni.e., for $m\\neq n$,\n \\begin{equation*}\n \\int_{-1}^1 C_n^{(\\alpha)}(x)C_m^{(\\alpha)}(x)w(x)\\,dx = 0.\n \\end{equation*}\n They are normalized by\n \\begin{equation*}\n \\int_{-1}^1 \\left[C_n^{(\\alpha)}(x)\\right]^2(1-x^2)^{\\alpha-\\frac{1}{2}}\\,dx = \\frac{\\pi 2^{1-2\\alpha}\\Gamma(n+2\\alpha)}{n!(n+\\alpha)[\\Gamma(\\alpha)]^2}.\n \\end{equation*}\nThe polynomials $t^p$, $p$ an even integer, can be represented by means of\n\\begin{equation*}\nt^p=\\sum_{k=0}^p \\lambda_k C^{\\alpha}_k(t).\n\\end{equation*}\nIt is known (see, e.g.,~\\cite{Bachoc:2005aa,Delsarte:1977aa}) that $\\lambda_i> 0$, $i=0,\\ldots,p$, and $\\lambda_0$ is given by\n \\begin{equation*}\n\\lambda_0= \\frac{1}{c}\\int_{-1}^1 t^p w(t) dt,\n \\end{equation*}\n where \n \\begin{equation*}\n c = \\frac{\\pi 2^{d+3}\\Gamma(d-2) }{(\\frac{d}{2}-1)\\Gamma(\\frac{d}{2}-1)^2}.\n \\end{equation*}\n Moreover, $C^\\alpha_k$ induces a positive kernel, i.e., for $\\{x_i\\}_{i=1}^N\\subset S^{d-1}$ and $\\{u_i\\}_{i=1}^N\\subset \\R$,\n \\begin{equation*\n \\sum_{i,j=1}^{N} u_iC^{\\alpha}_k (\\langle x_i,x_j\\rangle )u_j \\geq 0, \\quad \\forall k=0,1,2,...\n \\end{equation*}\n see~\\cite{Bachoc:2005aa,Delsarte:1977aa}. Note that the probability measures with finite support are weak star dense in $\\mathcal{M}(S^{d-1},\\mathcal{B})$. Since $C^{\\alpha}_k$ is continuous, we obtain, for all $\\mu\\in \\mathcal{M}(S^{d-1},\\mathcal{B})$, \n \\begin{equation*} \n \\int_{S^{d-1}} \\int_{S^{d-1}} C^{\\alpha}_k (\\langle x,y\\rangle )d\\mu(x) d\\mu(y) \\geq 0, \\quad \\forall k=0,1,2,...\n \\end{equation*}\nWe can then estimate\n\\begin{align*}\n\\int_{S^{d-1}} \\int_{S^{d-1}} |\\langle x,y\\rangle|^p d\\mu(x) d\\mu(y) & = \\int_{S^{d-1}} \\int_{S^{d-1}}\\sum_{k=0}^p \\lambda_k C^{\\alpha}_k (\\langle x,y\\rangle )d\\mu(x) d\\mu(y)\\\\\n& = \\sum_{k=0}^p \\lambda_k \\int_{S^{d-1}} \\int_{S^{d-1}}C^{\\alpha}_k (\\langle x,y\\rangle ) d\\mu(x) d\\mu(y) \\geq \\lambda_0.\n\\end{align*}\nFrom the results in \\cite{Seidel:2001aa}, one can deduce that \n \\begin{equation*}\n\\lambda_0= \\frac{1\\cdot 3\\cdot 5\\cdots(2t-1)}{d(d+2)\\cdots (d+2t-2) },\n \\end{equation*}\nwhich provides the desired estimate.\n\nWe still have to address the ``if and only if'' part. Equality holds if and only if $\\mu$ satisfies \n \\begin{equation*}\n \\int_{S^{d-1}}\\int_{S^{d-1}} C^{\\alpha}_k (\\langle x,y\\rangle ) d\\mu(x) d\\mu(y) = 0, \\quad \\forall k=1,\\ldots, p. \n \\end{equation*}\nWe shall follow the approach outlined in \\cite{Venkov:2001aa} in which the analog of Theorem~\\ref{theorem:p even integer discrete} was addressed for finite symmetric collections of points. In this case, the finite symmetric sets of points lead to finite sums rather than integrals as above. The key ideas that we need in order to use the approach presented in \\cite{Venkov:2001aa} are: First, $\\tilde{\\mu}(E):=\\frac{1}{2}(\\mu(E)+\\mu(-E))$, for $E\\in\\mathcal{B}$, satisfies $\\PFP(\\tilde{\\mu},p) = \\PFP(\\mu,p)$. Thus, we can assume that $\\mu$ is symmetric. Secondly and more critically, the map \n\\begin{equation*}\ny\\mapsto \\int_{S^{d-1}} |\\langle x,y\\rangle |^p d\\mu(x)\n\\end{equation*}\nis a polynomial in $y$. In fact, the integral resolves in the polynomial's coefficients. These two observations enable us to follow the lines in \\cite{Venkov:2001aa}, and we can conclude the proof.\n\\end{proof}\n \n \\begin{remark}\nOne may speculate that Theorem \\ref{theorem:p even integer} could be extended to $p\\geq 2$ that are not even integers. This is not true in general. For $d=2$ and $p=3$, for instance, the equiangular FUNTF with $3$ elements induces a smaller potential than the uniform distribution. The uniform distribution is a probabilistic tight $3$-frame, but the equiangular FUNTF is not.\n\\end{remark}\n\n\n\n\n\\section*{Acknowledgements}\nThe authors would like to thank C.~Bachoc, W.~Czaja, C.~Wickman, and W.~S.~Yu for discussions leading to some of the results presented here. M.~Ehler was supported by the Intramural Research Program of the National Institute of Child Health and Human Development and by NIH\/DFG Research Career Transition Awards Program (EH 405\/1-1\/575910). K.~A.~Okoudjou was partially supported by ONR grant N000140910324, by RASA from the Graduate School of UMCP, and by the Alexander von Humboldt foundation. \n\n\n\n\n\\bibliographystyle{plain}\n ","meta":{"redpajama_set_name":"RedPajamaArXiv"}}
{"text":"\\section{Introduction}\n\\thispagestyle{empty}\n\nOne equivalent characterization of the amenability of an infinite group $G$, called the \\textit{F{\\o}lner condition}, is that the isoperimetric constant (also known as Cheeger constant) of its Cayley graph should be $0$.\nThat constant is defined as the infimum of $\\frac{|\\partial F|}{|F|}$ over all finite sets $F\\subset G$ with $|F|\\leq\\frac{1}{2}|G|$.\nAs the quotient cannot reach $0$, amenability of infinite groups is therefore characterized by the existence of a sequence of sets $F_n$ such that $\\frac{|\\partial F_n|}{|F_n|}$ converges towards $0$, also known as a \\textit{F{\\o}lner sequence}.\nOne natural direction for studying the possible F{\\o}lner sequences on a given group is to ask how small the sets can be.\nWe consider the F{\\o}lner function.\nIt has classically been defined using the inner boundary:\n\\begin{equation}\\label{defdin}\n\t\\partial_{in}F=\\left\\{g\\in F:\\exists s\\in S\\bigcup S^{-1}:gs\\notin F\\right\\}.\n\\end{equation}\n\\begin{defi}\\label{foldef}\n\tThe \\textit{F{\\o}lner function} $\\Fol$ (or $\\Fol_S$; or $\\Fol_{G,S}$) of a group $G$ with a given finite generating set $S$ is defined on $\\N$ by\n\t$$\\Fol(n)=\\min\\left(|F|:F\\subset G,\\frac{|\\partial_{in}F|}{|F|}\\leq\\frac{1}{n}\\right).$$\n\\end{defi}\n\nRemark that $\\Fol(1)=1$ and that the values of the function are finite if and only if $G$ is amenable.\nIts values clearly depend on the choice of a generating set, but the functions arising from different generating sets (and more generally, functions arising from quasi-isometric spaces) are asymptotically equivalent.\nTwo functions are asymptotically equivalent if there are constants $A$ and $B$ such that $f(x\/A)\/B 2^n$, the result follows immediately.\nAssume that $|V(\\overline{G})|\\leq2^n$, and thus $c_K\\leq n$, and $c_i\\leq n-K+i$.\nThen\n$$|E(\\overline{G})|=\\sum_{i=1}^Kc_i2^{c_i}+\\sum_{i \\)}u_0(z)dz, \n\\end{equation}\nwith symmetric matrices $M_1, M_2,P\\in \\mathcal S_d(\\R)$. \nExperience shows that no linear term is needed in this formula, since\nthe potential is exactly quadratic (see\ne.g. \\cite{CLSS08}). \n\\smallbreak\n\nWe compute:\n\\begin{align*}\n i\\d_t u & = -i\\frac{\\dot h}{h}u -\\frac{1}{2}\\<\\dot M_1(t)y,y\\>u\\\\\n&\\quad \n +\\frac{1}{h}\\int e^{\\frac{i}{2}\\(\\dots\\)} \\(-\\frac{1}{2}\\<\\dot\n M_2(t)z,z\\>-\\<\\dot P(t)y,z\\>\\)u_0(z)dz,\n\\end{align*}\n\\begin{align*}\n \\d_{j}^2 u &= \\frac{1}{h}\\int e^{\\frac{i}{2}\\(\\dots\\)}\n \\(-\\(\\(M_1(t)y\\)_j + \\(P(t)z\\)_j\\)^2 -i\\(M_1\\)_{jj}\\)u_0(z)dz,\n\\end{align*}\nhence\n\\begin{align*}\n & i\\d_tu+\\frac{1}{2}\\Delta u = -i\\frac{\\dot h}{h}u\n +\\frac{i}{2}\\operatorname{tr} M_1 - \\frac{1}{2}\\<\\dot M_1(t)y,y\\>u\\\\\n&+ \\frac{1}{2h}\\int\n e^{\\frac{i}{2}\\(\\ \\)}u_0(z)\\times\\\\\n&\\times\\( \n-\\<\\dot M_2(t)z,z\\>-2\\<\\dot\nP(t)y,z\\>-|M_1(t)y|^2 -|P(t)z|^2 -2\n\\2$ seems essentially sharp in order to have\nglobal in time Strichartz estimates. The result remains true for $\\mu\n=2$ (\\cite{BPST03,BPST04}), but in \\cite{GoVeVi06}, the authors\nprove that for repulsive potentials which are homogeneous of degree\nsmaller than $2$, global Strichartz estimates fail to exist.\n\n\n\n\n\n\\section{Quantum scattering}\n\\label{sec:quant}\n\nIn this section, we prove Theorem~\\ref{theo:scatt-quant}. Since the\ndependence upon $\\eps$ is not measured in\nTheorem~\\ref{theo:scatt-quant}, we shall \nconsider the case $\\eps=1$, corresponding to \n\\begin{equation}\n \\label{eq:psi}\n i\\d_t \\psi +\\frac{1}{2}\\Delta \\psi = V\\psi + |\\psi|^{2\\si}\\psi.\n\\end{equation}\nWe split the proof of Theorem~\\ref{theo:scatt-quant} into two\nsteps. First, we solve the Cauchy problem with data prescribed at\n$t=-\\infty$, that is, we show the existence of wave operators. Then,\ngiven an initial datum at $t=0$, we show that the (global) solution to\n\\eqref{eq:psi} behaves asymptotically like a free solution, which\ncorresponds to asymptotic completeness. \n\\smallbreak\n\nFor each of these two steps, we first show that the nonlinearity is\nnegligible for large time, and then recall that the potential is\nnegligible for large time (linear scattering). This means that for any $\\tilde \\psi_-\\in\nH^1(\\R^d)$, there exists a unique $\\psi\\in \n C(\\R;H^1(\\R^d))$ solution to \\eqref{eq:psi} such that\n \\begin{equation*}\n \\|\\psi(t)-e^{-itH}\\tilde \\psi_-\\|_{H^1(\\R^d)}\\Tend t {-\n \\infty} 0,\n \\end{equation*}\nand for any $\\varphi\\in H^1(\\R^d)$, there exist a unique $\\psi\\in\n C(\\R;H^1(\\R^d))$ solution to \\eqref{eq:psi} and a unique $\\tilde\\psi_+\\in\n H^1(\\R^d)$ such that\n\\begin{equation*}\n \\|\\psi(t)-e^{-itH}\\tilde \\psi_+\\|_{H^1(\\R^d)}\\Tend t {+\n \\infty} 0.\n \\end{equation*}\nThen, we recall that the potential $V$ is negligible for large\ntime. We will adopt the following notations for the propagators,\n\\begin{equation*}\n U(t)=e^{i\\frac{t}{2}\\Delta},\\quad U_V(t)= e^{-itH}. \n\\end{equation*}\n\n\n\nIn order to construct wave operators which show that the nonlinearity\ncan be neglected for large time, we shall work with an $H^1$\nregularity, on the Duhamel's formula associated to \\eqref{eq:psi} in\nterms of $U_V$, with a prescribed asymptotic behavior as $t\\to\n-\\infty$:\n\\begin{equation}\n \\label{eq:duhamel-}\n \\psi(t) = U_V(t)\\tilde \\psi_- -i\\int_{-\\infty}^t\n U_V(t-s)\\(|\\psi|^{2\\si}\\psi(s)\\)ds. \n\\end{equation}\nApplying the gradient to this formulation brings up the problem of\nnon-commutativity with $U_V$. The worst term is actually the linear\none, $U_V(t)\\tilde \\psi_-$, since\n\\begin{equation*}\n \\nabla \\(U_V(t)\\tilde \\psi_-\\) = U_V(t)\\nabla \\tilde \\psi_-\n -i\\int_0^t U_V(t-s)\\((U_V(s)\\tilde \\psi_-)\\nabla V\\)ds.\n\\end{equation*}\nSince the construction of wave operators relies on the use of\nStrichartz estimates, it would be necessary to have an estimate of\n\\begin{equation*}\n \\left\\|\\nabla \\(U_V(t)\\tilde \\psi_-\\)\\right\\|_{L^qL^r}\n\\end{equation*}\nin terms of $\\psi_-$, for admissible pairs\n$(q,r)$. Proposition~\\ref{prop:StrichartzRS} yields\n\\begin{equation*}\n \\left\\|\\nabla \\(U_V(t)\\tilde \\psi_-\\)\\right\\|_{L^qL^r} \\lesssim \\|\\nabla \\tilde\n \\psi_-\\|_{L^2} + \\|(U_V(t)\\tilde \\psi_-)\\nabla V\\|_{L^{\\tilde\n q'}L^{\\tilde r'}},\n\\end{equation*}\nfor any admissible pair $(\\tilde q,\\tilde r)$. In the last factor,\ntime is present only in the term $U_V(t)\\tilde \\psi_-$, so to be able\nto use Strichartz estimates again, we need to consider $\\tilde\nq=2$, in which case $\\tilde r=2^*:=\\frac{2d}{d-2}$:\n\\begin{equation*}\n \\|(U_V(t)\\tilde \\psi_-)\\nabla V\\|_{L^2L^{{2^*}'}}\\le \\|U_V(t)\\tilde\n \\psi_-\\|_{L^2L^{2^*}}\\|\\nabla V\\|_{L^{d\/2}},\n\\end{equation*}\nwhere Assumption~\\ref{hyp:V} implies $\\nabla V\\in L^{d\/2}(\\R^d)$ as\nsoon as $\\mu>1$. Using the endpoint Strichartz estimate from\nProposition~\\ref{prop:StrichartzRS}, we have\n\\begin{equation*}\n \\|U_V(t)\\tilde\n \\psi_-\\|_{L^2L^{2^*}} \\lesssim \\|\\tilde \\psi_-\\|_{L^2}, \n\\end{equation*}\nand we have:\n\\begin{lemma}\\label{lem:stri2}\n Let $d\\ge 3$. Under the assumptions of\n Proposition~\\ref{prop:Morawetz}, for all admissible pair $(q,r)$, \n \\begin{equation*}\n \\|e^{-itH}f\\|_{L^q(\\R;W^{1,r}(\\R^d))}\\lesssim \\|f\\|_{H^1(\\R^d)}. \n \\end{equation*}\n\\end{lemma}\nWe shall rather use a vector-field, for we believe this approach may be\ninteresting in other contexts.\n\n\n\\subsection{Vector-field}\n\\label{sec:vector-field}\n\n We\nintroduce a vector-field which naturally commutes with $U_V$, and\nis comparable with the gradient. \n\\smallbreak\n\nFrom Assumption~\\ref{hyp:V}, $V$ is bounded, so there exists $c_0\\ge\n0$ such that $V+c_0\\ge 0$. We shall consider the operator\n\\begin{equation*}\n A = \\sqrt{H+c_0}=\\sqrt{-\\frac{1}{2}\\Delta +V+c_0}.\n\\end{equation*}\n\\begin{lemma}\\label{lem:A}\n Let $d\\ge 3$, and $V$ satisfying Assumption~\\ref{hyp:V} with\n $V+c_0\\ge 0$. For every $1
$ satisfies\n \\begin{equation}\\label{eq:decayQ}\n \\left\\|\\frac{d^\\alpha}{dt^\\alpha}Q(t)\\right\\|\\lesssim\n \\
\n\\end{equation*}\nthe time-dependent Hamiltonian present in \\eqref{eq:u}. Like in the\nquantum case, we show that the nonlinearity is negligible for large\ntime by working on Duhamel's formula associated to \\eqref{eq:u} in\nterms of $H_Q$. Since $H_Q$ depends on time, we recall that the\npropagator $U_Q(t,s)$ is the operator which maps $u_0$ to $u_{\\rm lin}(t)$,\nwhere $u_{\\rm lin}$ solves\n\\begin{equation*}\n i\\d_t u_{\\rm lin} +\\frac{1}{2}\\Delta u_{\\rm lin} =\n \\frac{1}{2}\\
u_{\\rm lin};\\quad u_{{\\rm lin}}(s,y)=u_0(y). \n\\end{equation*}\nIt is a unitary dynamics, in the sense that $U_Q(s,s)=1$, and\n$U_Q(t,\\tau)U_Q(\\tau,s)=U_Q(t,s)$;\nsee e.g. \\cite{DG}. Then to prove the existence of wave operators, we consider the\nintegral formulation\n\\begin{equation}\n \\label{eq:duhamel-wave-class}\n u(t) = U_Q(t,0)\\tilde u_--i\\int_{-\\infty}^t U_Q(t,s)\\(|u|^{2\\si}u(s)\\)ds.\n\\end{equation}\nA convenient tool is given by Strichartz estimates associated to\n$U_Q$. Local in time Strichartz estimates follow from general results\ngiven in \\cite{Fujiwara}, where local dispersive estimates are\nproven for more general potential. To address large time, we take\nadvantage of the fact that the \npotential is exactly quadratic with respect to the space variable, so\nan explicit formula is available for $U_Q$, entering the general\nfamily of Mehler's formulas (see e.g. \\cite{Feyn,HormanderQuad}). \n\\subsection{Mehler's formula}\n\\label{sec:mehler}\n\nConsider, for $t_0\\ll -1$,\n\\begin{equation*}\ni\\d_tu+\\frac{1}{2}\\Delta u=\\frac{1}{2}\\< Q(t)y,y\\> u\\quad\n;\\quad u(t_0,y)=u_0(y).\n\\end{equation*}\nWe seek a solution of the form\n\\begin{equation}\n \\label{eq:mehler}\n u(t,y) = \\frac{1}{h(t)}\\int_{\\R^d}\n e^{\\frac{i}{2}\\(\\
+t^2 \\IM \\int_{\\R^d} \\
.\n\\end{align*}\nOn the other hand, we still have\n\\begin{align*}\n \\frac{d}{dt}\\|u(t)\\|_{L^{2\\si+2}}^{2\\si+2}& =2 (\\si+1)\\int\n |u|^{2\\si}\\RE \\(\\bar u\\d_tu\\) = 2 (\\si+1)\\int\n |u|^{2\\si}\\RE \\(\\bar u \\times\\frac{i}{2}\\Delta u\\) ,\n\\end{align*}\nand so,\n\\begin{align*}\n \\frac{d}{dt}\\(\\frac{1}{2}\\|J(t)u\\|_{L^2}^2\n +\\frac{t^2}{\\si+1}\\|u(t)\\|_{L^{2\\si+2}}^{2\\si+2}\\)\n &=\\frac{t}{\\si+1}(2-d\\si)\\|u(t)\\|_{L^{2\\si+2}}^{2\\si+2}\\\\\n+ \n t\\RE\\int_{\\R^d} \\
& +t^2 \\IM \\int_{\\R^d} \\
. \n\\end{align*}\nThus for $t\\ge 0$ and $\\si\\ge\\frac{2}{d}$, \\eqref{eq:decayQ} implies\n\\begin{equation*}\n \\frac{d}{dt}\\(\\frac{1}{2}\\|J(t)u\\|_{L^2}^2\n +\\frac{t^2}{\\si+1}\\|u(t)\\|_{L^{2\\si+2}}^{2\\si+2}\\)\n \\lesssim \n \\
^{-\\mu}\\nabla u\\|_{L^1(t,\\infty;L^2)}. \n\\end{align*}\nThe right hand side goes to zero as $t\\to \\infty$, hence the\nproposition. \n\\end{proof}\n\n\\begin{remark}\n As pointed out in the previous section, it would be possible to\n prove the existence of wave operators by using an adapted vector\n field $\\mathcal A$. On the other hand, if $Q(t)$ is not proportional\n to the identity matrix, it seems that no (exploitable) analogue of\n the pseudo-conformal conservation law is available in terms of\n $\\mathcal A$ rather than in terms of $J$. \n\\end{remark}\n\\subsection{Conclusion}\n\\label{sec:concl-class}\n\nLike in the case of quantum scattering, we use a stronger version of\nthe linear scattering theory:\n\\begin{proposition}\\label{prop:Cook-class}\n Let $d\\ge 1$, $V$ satisfying Assumption~\\ref{hyp:V} with $\\mu>1$. Then \nthe strong limits\n \\begin{equation*}\n \\lim_{t\\to \\pm \\infty} U_Q(0,t)U(t) \\quad \\text{and}\\quad \\lim_{t\\to\n \\pm\\infty} U(-t) U_Q(t,0) \\quad \\text{and}\\quad \n\\end{equation*}\nexist in $\\Sigma$. \n\\end{proposition}\n\\begin{proof}\n For the first limit (existence of wave operators), again in view of\n Cook's method, we prove that for all $\\varphi\\in \n \\Sch(\\R^d)$, \n\\begin{equation*}\n t\\mapsto \\left\\| U_Q(0,t) \\U(t)\\varphi\\right\\|_{\\Sigma}\\in\n L^1(\\R). \n \\end{equation*}\nFor the $L^2$ norm, we have, in view of \\eqref{eq:decayQ},\n\\begin{equation*}\n \\left\\| U_Q(0,t) \\
U(t)\\varphi\\right\\|_{L^2} \\lesssim\n \\
U(t)\\varphi\\right\\|_{L^2} \\lesssim\n \\
U_Q(t,0)\\varphi\\right\\|_{\\Sigma}\\in\n L^1(\\R). \n \\end{equation*}\nFor the $L^2$ norm, we have\n\\begin{align*}\n \\left\\| U(-t) \\
U_Q(t,0)\\varphi\\right\\|_{L^2}&= \\left\\|\n \\
U_Q(t,0)\\varphi\\right\\|_{L^2}\\\\\n& \\lesssim\n \\
$ is exactly quadratic in space, and so\ndifferentiating it three times with any space variables yields zero. \n\\end{proof}\n\n\n\n\\section{Proof of Theorem~\\ref{theo:cv}}\n\\label{sec:cv}\n\nThe main result of this section is:\n\\begin{theorem}\\label{theo:cv-unif}\n Let $d=3$, $\\si=1$, $V$ as in Theorem~\\ref{theo:scatt-quant}, and\n $u_-\\in \\Sigma^7$. Suppose that Assumption~\\ref{hyp:flot} \n is satisfied. Let $\\psi^\\eps$ be given by\n Theorem~\\ref{theo:scatt-quant}, $u$ be given by\n Theorem~\\ref{theo:scatt-class}, $\\varphi^\\eps$ defined by\n \\eqref{eq:phi}. We have\n the uniform error \n estimate:\n \\begin{equation*}\n \\sup_{t\\in \\R}\\|\\psi^\\eps(t)-\\varphi^\\eps(t)\\|_{L^2(\\R^3)} =\n \\O\\(\\sqrt\\eps\\). \n \\end{equation*}\n\\end{theorem}\nTheorem~\\ref{theo:cv} is a direct consequence of the above\nresult, whose proof is the core of\nSection~\\ref{sec:cv}. From now on, we assume $d=3$ and $\\si=1$. \n\\subsection{Extra properties for the approximate solution}\n\\label{sec:extra-u}\n\nFurther regularity and localization properties on $u$ will be\nneeded. \n\\begin{proposition}\\label{prop:extra-u}\n Let $\\si=1$, $1\\le d\\le 3$, $k\\ge 2$ and $V$ satisfying\n Assumption~\\ref{hyp:V} for some $\\mu>1$. If $u_-\\in \\Sigma^k$, then\n the solution $u\\in C(\\R;\\Sigma)$ provided by Theorem~\\ref{theo:scatt-class}\n satisfies $u\\in C(\\R;\\Sigma^k)$. The momenta\n of $u$ satisfy\n \\begin{equation*}\n \\lVert \\lvert y\\rvert^\\ell u(t,y)\\|_{L^2(\\R^d)}\\le C_\\ell\n \\
\\)\\Big|_{y=\\frac{x-q(t)}{\\sqrt\\eps}}\n \\varphi^\\eps(t,x). \n\\end{equation*}\nDuhamel's\nformula for $w^\\eps$ reads\n\\begin{align*}\n w^\\eps(t) &= -i\\eps^{3\/2}\\int_{-\\infty}^t U^\\eps_V(t-s)\\(|\\psi^\\eps|^{2}\\psi^\\eps\n -|\\varphi^\\eps|^{2}\\varphi^\\eps\\)(s)ds\\\\\n&\\quad +i\\eps^{-1}\\int_{-\\infty}^t U^\\eps_V(t-s) \\mathcal L^\\eps(s)ds. \n\\end{align*}\nDenoting $L^a(]-\\infty,t];L^b(\\R^3))$ by $L^a_tL^b$, Strichartz\nestimates yield, for any $L^2$-admissible pair $(q_1,r_1)$,\n\\begin{equation}\\label{eq:stri-weps}\n \\eps^{1\/q_1}\\|w^\\eps\\|_{L^{q_1}_t L^{r_1}} \\lesssim\n \\eps^{3\/2-1\/q}\\left\\||\\psi^\\eps|^{2}\\psi^\\eps \n -|\\varphi^\\eps|^{2}\\varphi^\\eps\\right\\|_{L^{q'}_tL^{r'}} +\n \\frac{1}{\\eps}\\|\\mathcal L^\\eps\\|_{L^1_tL^2},\n\\end{equation}\nwhere $(q,r)$ is the admissible pair chosen in the proof of\nProposition~\\ref{prop:waveop-quant}, that is $r=2\\si+2$. Since we now\nhave $d=3$ and $\\si=1$, this means:\n\\begin{equation*}\n q=\\frac{8}{3},\\quad k=8,\n\\end{equation*}\nand \\eqref{eq:stri-weps} yields\n\\begin{equation}\\label{eq:w-presque}\n \\eps^{1\/q_1}\\|w^\\eps\\|_{L^{q_1}_t L^{r_1}} \\lesssim\n \\eps^{9\/8}\\( \\|w^\\eps\\|^2_{L^8_t L^4}+ \\|\\varphi^\\eps\\|^2_{L^8_t\n L^4}\\)\\|w^\\eps\\|_{L^{8\/3}_tL^4} +\n \\frac{1}{\\eps}\\|\\mathcal L^\\eps\\|_{L^1_tL^2}.\n\\end{equation}\nThe strategy is then to first\nobtain an a priori estimate for $w^\\eps$ in $L^8_tL^4$, and then to\nuse it in the above estimate. In order to do so, we begin by\nestimating the source term $\\mathcal L^\\eps$, in the next subsection. \n\\subsection{Estimating the source term}\n\\label{sec:estim-source-term}\n\n\\begin{proposition}\\label{prop:est-source}\n Let $d= 3$, $\\si=1$, $V$ satisfying Assumption~\\ref{hyp:V}\n with $\\mu>2$, and $u_-\\in \\Sigma^k$ for some $k\\ge\n 7$. Suppose that Assumption~\\ref{hyp:flot} is satisfied.\nLet $u\\in C(\\R;\\Sigma^k)$ given by\n Theorem~\\ref{theo:scatt-class} and \n Proposition~\\ref{prop:extra-u}. The source term $\\mathcal L^\\eps$ satisfies\n \\begin{equation*}\n \\frac{1}{\\eps} \\|\\mathcal L^\\eps(t)\\|_{L^2(\\R^3)}\\lesssim \\frac{\\sqrt\n \\eps}{\\
\\)u(t,y).\n\\end{equation*}\nIn particular,\n\\begin{equation*}\n \\frac{1}{\\eps}\\| \\mathcal L^\\eps(t)\\|_{L^2(\\R^3)} = \\|{\\mathcal\n S}^\\eps(t)\\|_{L^2(\\R^3)} ,\\quad \\frac{1}{\\eps}\\| \\mathcal\n L^\\eps(t)\\|_{L^{3\/2}(\\R^3)} = \\eps^{1\/4}\\|{\\mathcal \n S}^\\eps(t)\\|_{L^{3\/2}(\\R^3)}. \n\\end{equation*}\n Taylor's formula and Assumption~\\ref{hyp:V} yield the pointwise estimate\n \\begin{equation*}\n | {\\mathcal S}^\\eps(t,y) | \\lesssim \\sqrt\\eps |y|^3 \\int_0^1\n \\frac{1}{\\
^{\\mu+3}}d\\theta |u(t,y)|. \n \\end{equation*}\nTo simplify notations, we consider only positive times. Recall that\nfrom Assumption~\\ref{hyp:flot}, $p^+\\not =0$. Introduce, for\n$0<\\eta< |p^+|\/2$,\n\\begin{equation*}\n \\Omega = \\left\\{y\\in \\R^3,\\quad |y|\\ge \\eta\\frac{t}{\\sqrt\\eps}\\right\\}.\n\\end{equation*}\nSince $q(t)\\sim p^+ t$ as\n$t\\to \\infty$, on the complement of $\\Omega$, we can use the decay of $V$,\n\\eqref{eq:short}, to infer the pointwise estimate\n\\begin{equation}\\label{eq:Spoint}\n | {\\mathcal S}^\\eps(t,y) | \\lesssim \\sqrt\\eps |y|^3\n \\frac{1}{\\